Download as pdf or txt
Download as pdf or txt
You are on page 1of 91

Solar Energy

Impact of charge-compensated Fe and Nb co-substitution on BaTiO3: bandgap and


grain size reduction and enhanced bulk photovoltaic power of Al/BFNT/Ag solar cell
--Manuscript Draft--

Manuscript Number: SEJ-D-22-03508R1

Article Type: Research paper

Section/Category: Photovoltaic materials, cells and systems

Keywords: Bandgap; Photocurrent; Open circuit voltage; Ferroelectric Photovoltaic

Corresponding Author: Sundarakannan B, Ph.D

INDIA

First Author: L. Venkidu

Order of Authors: L. Venkidu

N. Raja

Sundarakannan B, Ph.D

Abstract: The generation of above bandgap photovoltage using bulk ferroelectric materials has
become a subject of great interest, however, their photocurrent density is limited by a
broad bandgap and poor conductivity. To overcome this limitation, we replaced
aliovalent metal ions (Fe3+ and Nb5+) at the B-site of robust ferroelectric BaTiO3 and
fabricated an Al/BaTi 1-2x Fe x Nb x O3/Ag photovoltaic device. Both the experimental
and the theoretical studies showed that bandgap was lowered to ~2.55 eV and hence
absorption of wide energy range of the solar spectrum was attained. An apt top
electrode, reduced bandgap and domain size resulted in greater photocurrent density
of 1.46 μA/cm2 and photovoltage of 8.31 volts for Al/0.075BFNT/Ag solar cell in
unpoled condition. This research suggest that reduced band gap, mixed structural
phases and nano-sized domains suffices greatest PV power output while the large
polarization and poling are not necessary prerequisites.

Suggested Reviewers: Tansir Ahamad


King Saud University
tahamed@ksu.edu.sa

Duraisamy Kumaresan
Amrita Vishwa Vidyapeetham Amrita School of Engineering
k_duraisamy@cb.amrita.edu

Yunbin He
Hubei Key Laboratory of Polymer Materials
ybhe@hubu.edu.cn

Sarp Kaya
Koç University
sarpkaya@ku.edu.tr

Powered by Editorial Manager® and ProduXion Manager® from Aries Systems Corporation
Cover Letter

30.12.2022
From
Dr. B. Sundarakannan
Professor
Department of Physics
Manonmaniam Sundaranar University
Tirunelveli -627 012, Tamil Nadu
India
E-mail:sundarakannan@msuni.ac.in

To
Ranga Pitchumani
Editor-in-Chief ,
Virginia Tech,
Falls Church, Virginia,
United States of America.

Sub.: Submission of a manuscript to Solar Energy -Reg.,


I am pleased to submit an original research manuscript with the title “Impact of
charge-compensated Fe and Nb co-substitution on BaTiO3: bandgap and grain size
reduction and enhanced of bulk photovoltaic power of Al/BFNT/Ag solar cell” by L.
Venkidu, N.Raja and B.Sundarakannan for publication in the “Solar Energy”. On behalf of all
the authors, I ensure that this manuscript is unpublished and not being considered for
publication elsewhere. All the authors have seen and approved the manuscript.

In this manuscript, we present the following; (i). Fe3+ and Nb5+ co-doped ferroelectric
photovoltaic solar cell does not exist experimentally. Therefore, we have selected wide
bandgap ferroelectric BaTiO3 ceramics and reduced its bandgap using aliovalent Fe3+ and Nb5+
ions substitution at B-site to fabricate a bulk ferroelectric photovoltaic device in the
Al/BFNT/Ag configuration. (ii) In particular, the bandgap of the Al/0.075BFNT/Ag device
was reduced to ~2.55 eV, which was confirmed by experimental and theoretical DFT
calculations. (iii) This study suggests that small band gap and reduced grain size are sufficient
for maximum bulk photovoltaic power generation (~ 12 μW/cm2), whereas substantial
polarization and poling are not essential factors.
Up to the author’s knowledge, enhancement of photovoltaic properties using Al/ BaTi1-
2xFexNbxO3/Ag device configuration has not yet been published. Under these circumstances,
we believe that this manuscript is suitable for publication in the "Solar Energy". In view of our
present new findings and the suitability of our conclusions to the scope of the Journal, we do
appeal to you to arrange for the review of the manuscript. Furthermore, we checked the
grammar and plagiarism in the manuscript using Grammarly and Turnitin software.

Thanking You
Yours faithfully
B. Sundarakannan
Detailed Response to Reviewers

Detailed response to the reviewers

To

Dr. Giribabu Lingamallu, Ph.D


Associate Editor
Solar Energy

Dear Sir,

Ref: SEJ-D-22-03508

Title of the manuscript: "Impact of charge-compensated Fe and Nb co-substitution on

BaTiO3: bandgap and grain size reduction and enhanced bulk photovoltaic power of

Al/BFNT/Ag solar cell"

Thank you so much for considering our manuscript and providing an opportunity for improving

the manuscript based on the important suggestions from the reviewers. We have revised the

manuscript as per the suggestions of the reviewers and also provided the required additional

details by the reviewers.

Herein the point-by-point responses to the comments raised by reviewers are provided.

Accordingly, the changes are carried out in the manuscript and are highlighted (yellow color)

in the revised manuscript.

Response to the Reviewer- 1

1. Literature survey reported in the manuscript on the (bulk) photovoltaic effect of ferroelectric

BaTiO3 is insufficient. The following key papers must be included.

i). A. Zenkevich et al. PHYSICAL REVIEW B 90, 161409(R) (2014), Doi:

10.1103/PhysRevB.90.161409

ii). Y. Liu et al. Phys. Chem. Lett. 2022, 13, 48, 11071-11075. Doi:
10.1021/acs.jpclett.2c03194

iii) G. Baiju et al. Solar Energy, 2021, 224,93-101. DoI: 10.1016/j.solener.2021.05.063

iv) V. M. Fridkin, Ferroelectrics, 2015, 484, 1-13. DOI: 10.1080/00150193.2015.1059151

Response: Thank you for your suggestion. As per the reviewer’s suggestion, the above

papers were added in the revised manuscript and as reference number:7, 8, 9, and 18

respectively and are highlighted in the revised manuscript.

2. In the experimental section, a short and complete description of each experimental set-up

must be provided. In XPS and Raman analyses, ensure that the photon energy used in the

measurement is given. Give the source, detector, and resolution details. See that the fitting

parameters are given in detail, including the background. The choice of line shapes, line widths,

and deconvolution should be documented well. The type of analyzer and the geometry of the

measurement should be described. Finally, the software used in the analysis should be listed.

Errors in fitting parameters must be present.

Response: Thank you for your suggestion. As per the reviewer’s suggestion the details

of both the XPS and the Raman analyses were added in the experimental section. For

readability, the complete Experimental section is provided below and highlighted sentences are

newly introduced in the revised manuscript as per the reviewer’s suggestion.

Experimental section:

The bulk ceramics powder with the composition Ba(Ti1-2xFexNbx)O3 (BFNT) (x=0.0,

0.025, 0.05, 0.075) was fabricated by solid-state reaction method (SSR) and represented as BT,

0.025BFNT, 0.05BFNT and 0.075BFNT. The starting materials of BaCO3 (99.9%), TiO2

(99%), Fe2O3 (99.9%), and Nb2O5 (99.99%) were weighted according to the stoichiometry of

each composition. The mixed powder was ground using a mortar and pestle for 30 min. The

mixture was calcined at 950oC/5 hr in the air atmosphere for all compositions. The calcined

powder was mixed with a binder of PVA, (polyvinyl alcohol) by manual granulation in a
mortar. The uniaxially compressed pellets are 13 mm in diameter and 1 mm thick under 7-ton

weight. The pellets were sintered at 300oC/1hr to remove the PVA and then sintered at

1300oC/2hr in a furnace. Room temperature X-ray diffraction measurements were carried out

using λ= 0.9654 Å the angle-dispersive X-ray diffraction (ADXRD) at the beamline-12

installed in Indus-2 synchrotron source RRCAT Indore, India. The FT-Raman spectrometer

(Bruker, RFS 27) was used to probe the structural phases of the samples. The instrument is

equipped with an Nd: YAG laser (1064 nm) as a source and an InGaAs detector (1024 × 256).

It has a resolution of 2 cm-1. The measured data were fitted with the Lorentz function using

Origin software (version:8.5) and fitting parameters such as peak position and line width were

added in supporting information. To determine the bandgap of the samples, a diffuse

reflectance spectrophotometer (DRS) (UV 2400 Schimadzu) was used. The silver paste was

applied on both sides of the polished pellets for measuring the room temperature polarization-

electric field (P-E loop) hysteresis loops using a Precision ferroelectric tester system (Radian

technologies P-E loop tracer). The PHI - VERSAPROBE III is an advanced instrument that

was used to obtain X-ray photoelectronic spectra. It features a multichannel detector and

monochromatic Al-Kα source with an energy of 1486.6 eV, providing excellent resolution

ranging from 0.1 to 40 eV for sample analysis and characterization. The peaks background

were subtracted for more accuracy in peak fitting, then the Gaussian function was used to fit

each elemental XPS peak with Origin software (version: 8.5) and the parameters were included

in supporting information. The Energy Dispersive X-ray Spectrometer Quantax 200 with X

Flash® 6130 was used for EDX and mapping studies. The photovoltaic measurements were

done using Sol3A Class AAA solar simulator with Keithley electrometer (2400) measuring

setup.

3. In XRD, Fig 1 (a), Each peak index must be visible clearly, and the font size must be large

enough to avoid obscurity


Response: Thank you for your suggestion. As per the reviewer’s suggestion, the peak index

and font size of the text in fig.1 were increased to 14 from 10 changes for clear visibility in

the manuscript.

4. On Page 9, line 29, where denotes the <Ti-O-Ti bond angle calculated… must be defined

unambiguously in words

Response: Thank you for your suggestion. As per the reviewer’s suggestion. This point is

addressed in the revised manuscript in the following way:

180−𝜃
𝜃𝑡𝑖𝑙𝑡 = 2

where 𝜃 is the angle between two Ti4+ ions in adjacent unit cells bonded with a face-

centered oxygen ion (Ti4+-O2--Ti4+).

5. In Equation 2, the symbols do not have clear superscripts and subscripts, making it vague

Response: Thank you for your suggestion. As per the reviewer’s suggestion the

equation (2) is rewritten as given below in the revised manuscript.


6
1 𝛿𝑛 − < 𝛿 > 2
∆𝑜𝑐𝑡 = ∑ [ ]
6 <𝛿>
𝑖=1

where 𝛿𝑛 and < 𝛿 > denote the individual and average interatomic metal-oxygen bond

length.

6. In the XPS survey scan spectra of samples, a significant shift in the C1s peaks is seen. Why?

Response: Thank you for your question we provide the following answer. The presence

of a ubiquitous carbon (C1s) peak in carbon-free samples is due to atmospheric carbon

contamination on sample surfaces. As per the reviewer’s suggestion, in the magnified view

(attached here) C-1s peak was observed at ~283.07, ~283.14 and ~ 283.2 for BT, 0.025BFNT

and 0.075BFNT samples respectively which is consistent with recent research publication

(https://doi.org/10.1007/s10854-020-03900-y). However, no significant shift was observed in


the peak position of C-1s but there was a noticeable shoulder peak associated with barium

carbonate species (https://doi.org/10.1007/s10854-020-03900-y).

7. On page 15, line 16., short-circuit current density increased under increasing the substitution

and achieved 8.31 V …..The large Voc (i.eVoc>Eg) justifies that it is an anomalous

photovoltaic effect…..This is not an uncommon observation, as claimed. There has been a

wealth of reports available on this effect.

Response: Thank you for your suggestion. We agree that many ferroelectric materials

exhibit photovoltage greater than bandgap and for example in the following reports, the

photovoltage is greater. We provide the following answer. Many bulk ferroelectric materials

exhibited an abnormal photovoltaic effect, which is frequently reported

(https://doi.org/10.1063/1.5088635, https://doi.org/10.1103/PhysRevApplied.11.044007,

https://doi.org/10.1038/s41598-018-26205-x). However, as mentioned in the introduction


section, the main difficulty of bulk ferroelectric photovoltaics (BFPV) is that they achieve a

very low output short-circuit current density in the range of pA/cm2 to nA/cm2

(https://doi.org/10.1038/s41598-018-26205-x, https://dx.doi.org/10.1021/acs.jpcc.0c10655,

https://doi.org/10.1063/1.5088635). Also, the key focus of the present work is to improve both

the Jsc and Voc of the BFPV device simultaneously, and we have achieved an impressive result

that the Jsc reached micro Ampere (μA/cm2) while a large open circuit voltage was retained.

Moreover, from fig. S6, the output power (Jsc×Voc) of the device was substantially higher than

that of several BFPV reports (Table S7).

8. An EDX mapping/spectrum of grains observed in SEM images must be provided in the

microscopic analysis

Response: Thank you for your suggestion. As per the reviewer’s suggestion, the EDX

mapping and spectrum were taken for all the samples and included in the supplementary as fig.

S6 (a, b, c, d) and discussed in the microstructure analysis portion of the supporting

information. As per the reviewer’s suggestion, this point is addressed in the manuscript in the

following way:

This study utilized EDX spectrum and elemental mapping analysis to confirm the presence of

Ba, Ti, and O elements in pure BT as shown in fig. S6 (a). Further, the expected amount of Ba,

Ti, Fe, Nb and O were found to be present in aliovalent substituted samples as illustrated by

fig. S6 (b-d), with atomic percentage for each element demonstrated by table S4 . The EDX

mapping results indicated a homogeneous distribution of all the ions across all samples. It was

also noted that substituted ions such as Fe3+ and Nb5+ are also homogeneously spread out in

0.05BFNT and 0.075BFNT.


Response to the Reviewer- 2

1. I'm wondering how did you relax your structures? Was that stepwise or +U correction and

SOC effect were included from beginning of all relaxations? It seems for 136 atoms it most be

very costly.

Response: Thank you for your question and suggestion to improve the clarity of the

manuscript. We start the DFT calculations by using the experimental cell parameters of the

cubic BFNT. A 3×3×3 supercell was used, with four Ti atoms replaced randomly by two Fe-

Nb pairs to represent Ba(Ti1-2xFexNbx)O3 (BFNT) (x=0.075). To take into account the strong

correlation of Fe atom, we have carried out the ionic relaxation calculation by using GGA + U

method with U= 1-5 eV (U is on-site Coulomb interaction), and the SOC effect was also

considered from the beginning for all the calculations. The choice of this methodology is based

on the previously published reports (https://doi.org/10.1103/PhysRevB.77.085122;

https://doi.org/10.1103/PhysRevB.87.214110

https://doi.org/10.1103/PhysRevB.57.1505).

Later the possible combinations of the dopants sites based on the dopant-dopant distance

(explained in Q. No 2) were identified and the ionic relaxation calculations were carried out as

explained above. Even though the computation is expensive, the quality of the results is reliable

(The calculations were carried out by using HEMERA 4 clusters, HZDR (Helmholtz-Zentrum

Dresden-Rossendorf), Germany, and NANDADEVI clusters in Institute of Mathematical

Science, Taramani, Chennai, Tamilnadu, India.

2. For doping possibilities I think it's more accurate to run a simple first and second neighbor

analysis of X-sites to figure out number of possibilities and then reduce them by doping-doping

distance criteria. I've checked also the figures of replaced sites which were not really clear how
they are only possible choices. I recommend to add a convincing paragraph that those

replacement are only possible configurations.

Response: We begin the calculations by taking the experimental cell parameters of the

cubic BFNT and replacing four Ti atoms with two Fe-Nb pairs to reflect the composition,

Ba(Ti1-2xFexNbx)O3 (BFNT) (x=0.075). We are well aware that larger supercells are needed for

calculations of this kind to take into consideration substituted systems, but as the number of

atoms in these supercells rises, the size of the supercell can be restricted to a 3×3×3 supercell.

Therefore we limit ourselves to only considering the first and second neighbor distance as the

criteria, to find the minimum energy configuration of the composition. Within the 3×3×3

supercell, twelve different configurations of particularly substituted systems are feasible

according to permutations.

𝑛!
𝑃𝑒𝑟𝑚𝑢𝑡𝑎𝑡𝑖𝑜𝑛, 𝑛𝑃𝑟 =
(𝑛 − 𝑟)!

where n and r represents the total number of lattice sites used for doping and number of type

of atoms replaced.

Initially, the dopants were randomly placed in the nearest neighbor sites of the supercell

and it was discovered that the lattice sites around the dopants were slightly distorted around

the dopant cites, exhibiting tetragonal symmetry rather than cubic symmetry. Among those,

the non-symmetrical arrangements cause very large tetragonal distortions within and around

the dopant sites, such as two Fe ions and one Nb ion in the ac-plane and only one Nb ion along

the b-axis. According to previously published literature results, such larger distortions lead to

widening the energy bandgap of the BFNT (https://dx.doi.org/10.1021/acs.jpcc.0c10655).

Taking this into consideration, four similar configurations were ignored in the calculations.

Therefore, the remaining eight symmetrically non-degenerate configurations, as illustrated in

Fig. S3 (Supporting Information), were taken only considered to reflect the composition,

Ba(Ti1-2xFexNbx)O3 (BFNT) (x=0.075). It turns out that configuration 2 (as shown in Fig. S3
and Table S5), with the two Fe dopant ions positioned as the first nearest and Fe-Nb positioned

as the second nearest neighbor, is the lowest energy or ground state. Further, the DOS

calculation was done only for that configuration 2 by using HSE hybrid functional.

3. In section 3.5, regrading PDOS results why really by additional Fe, contribution of O-2p

reduces so drastically at VB and Ti-3d contribution increases? Furthermore, does that might

cause decrease of polarity you have reported? I recommend to add a short description about

this here.

Response: The author would like to thank the reviewer for the suggestion regarding

section 3.5 (fig. 5(d)), where the green solid line was incorrectly used for Fe-3d and the blue

solid line for O-2p. After checking the plots again, it is found that these labels had been

mistakenly interchanged and this mistake is now rectified in a revised version of Figure 5(d)

(attached below), which has also been added to the revised manuscript. We ensured that all the

graphs are correct now after rechecking. From fig. 5(d), the PDOS analysis shows that Ti-3d

is much lower in VB than O-2p upon metal substitution. This is consistent with pure BT.
As per the reviewer’s suggestion, the contribution of Fe & Nb in VB & CB are

addressed in the manuscript in the following way: For the sake of the readability for the

reviewer, the whole paragraph is given below where newly added lines are highlighted. “It is

remarkable to notice from fig. 5(d) that on metal substitution at B-site, the Ti-3d and O-2p

orbital still dominate the CB minimum and VB maximum of the BFNT respectively. With

increasing substitution level to x=0.075, the formation of new 3d states of Fe and Nb was

observed at both the valence and conduction bands as shown in fig. 5(d). The PDOS in Fig.

5(d) shows the hybridization of Fe ion 3d-states with BT host electronic states, close to the

valence band maximum. The incorporation of Fe-3d states localized in VB has resulted in

stronger hybridization with the O-2p state near the valance band edge than that observed for

Ti–O hybridization. HSE calculations clearly show the formation of new states at the top of the
valence band edge with predominant Fe 3d character, which primarily decreases bandgap.

Additionally, strong d–d interaction between Ti–3d and Fe/Nb – d state at top of VB has led to

the accumulation of carriers near bottom CB as well as top VB edges. Also, the contribution

of lower Fe-3d (~ -7.2 to -6.7 eV) state increases the width of VB from ~ -4.75 eV (i.e. BT) to

~ -8 eV as shown in fig. S4. Such enhancement is beneficial for the mobility of the

photogenerated charge carriers within the material system as from an earlier report

(https://doi.org/10.1155/2012/823498). On the other hand, fig. 5(d), the substitution of Fe and

Nb in BFNT results in the hybridization of Nb–O and Fe-O at the conduction band, which

effectively separates the photogenerated electron-hole pair

(https://doi.org/10.1080/00150193.2018.1528958). Furthermore, this hybridization between

Fe-3d and Nb-3d with O-2p pushes the CB minimum towards lower energy levels resulting in

the significant reduction of the conduction band edge from 3.26 to 2.49 eV, as shown in Fig.

5(d) in the reason for minimizing the photoexcitation energy of charge carriers. The red shift

of CBM combined with an increase in mobility of photogenerated charge carriers along VB

due to Fe–O & Nb–O in BFNT leads to the expectation of improvement in photocurrent

density”.

Furthermore, does that might cause decrease of polarity you have reported? I recommend to

add a short description about this here.

Response: The study shows that the ferroelectric polarization and bandgap of the

samples are directly connected with the BO6 (Ti/Fe/Nb-O6) octahedra. The octahedral tilt was

suppressed due to metal doping, leading to a cubic phase. According to the theoretical

investigation, hybridization among Ti-3d0 and O-2p is responsible for off-center ion

displacement which results in larger ferroelectric polarization and wider bandgap in pure BT.

Furthermore, B cation off-center displacements have been reported to enhance the antibonding

character of conduction bands (CB), thus widening the band gap in the tetragonal phase as per
earlier reports (http://dx.doi.org/10.1063/1 .4871707). Metal substitution leads to stronger

hybridization of non-d0-Fe3+ and O 2p resulting in reduced tilt angle along with decreased

antibonding character which leads to narrow bandgap; also it reduces the off-center

displacement due to localized Fe orbital resulting in lower polarization observed in BFNT

samples.

4. Regarding the last point, Ti-d doesn't need to be U correction? As its d orbital contribution

at VB becomes important.

Response: Thank you for your question. Based on the earlier estimations in the

literature, we didn’t consider the U correction for Ti ion (

https://doi.org/10.1103/PhysRevB.87.214110;

https://doi.org/10.1039/D1TC01868J;

DOI: 10.1209/0295-5075/124/27005). Further for DOS calculations, hybrid functional

calculations were carried out, since it is well known to overcome the shortcomings of the GGA

pseudopotentials (https://doi.org/10.1103/PhysRevB.78.104116;

https://doi.org/10.1021/acsenergylett.8b00492). As per the Fe-ion is concerned one needs to

find the ground state accurately to find the band gap value in the match with the experimental

results so the U term is only considered for Fe ion.

5.Separated Figures and captions were hard to reach. Figures caption are also needed to be

completed, please include all details that one can understand them independently. The quality

of figures, at least in this version, is not good enough

Response: Thank you for your suggestion. As per the reviewer’s suggestion, the quality

of figure are improved and the detailed figure captions were added to the manuscript as well

as in the supporting information.


Marked Revised Manuscript Click here to view linked References

Impact of charge-compensated Fe and Nb co-substitution on BaTiO3: bandgap and grain

size reduction and enhanced of bulk photovoltaic power of Al/BFNT/Ag solar cell

L.Venkidua, N. Rajaa&b and B.Sundarakannana*

a Department of Physics, Manonmaniam Sundaranar University, Abishekapatti, Tirunelveli-

627012, Tamil Nadu, India.

b The Institute of Mathematical Sciences, C.I.T. Campus, Taramani, Chennai 600113, India.

Corresponding author: sundarakannan@msuniv.ac.in

Abstract

The generation of above bandgap photovoltage using bulk ferroelectric materials has

become a subject of great interest, however, their photocurrent density is limited by a broad

bandgap and poor conductivity. To overcome this limitation, we replaced aliovalent metal ions

(Fe3+ and Nb5+) at the B-site of robust ferroelectric BaTiO3 and fabricated an Al/BaTi1-

2xFexNbxO3/Ag photovoltaic device. Both the experimental and the theoretical studies showed

that bandgap was lowered to ~2.55 eV and hence absorption of wide energy range of the solar

spectrum was attained. An apt top electrode, reduced bandgap and domain size resulted in

greater photocurrent density of 1.46 μA/cm2 and photovoltage of 8.31 volts for

Al/0.075BFNT/Ag solar cell in unpoled condition. This research suggest that reduced band

gap, mixed structural phases and nano-sized domains suffices greatest PV power output while

the large polarization and poling are not necessary prerequisites.

Keywords: Bandgap, Photocurrent, Open circuit voltage, Ferroelectric Photovoltaic

1 Introduction:

Photovoltaic solar energy is obtained by converting sunlight into electricity using a p-

n junction based on the semiconductors [1]. Photovoltaic (PV) technology is playing a


significant role in modern industry to alleviate the energy crisis. Traditional junction-based

solar cell technology is widely used, though the raw material of silicon (Si) is relatively high

price and the fabrication process requires so sophisticated technology. The output voltage in

semiconductor based material is limited by their bandgap and the power conversion efficiency

(PCE) depends on the complicated interfaces of PV devices. Conversely, non-centrosymmetric

bulk ferroelectric materials proficiently separate electron-hole pairs through their internal

electric field originating from the spontaneous polarization represented as the bulk photovoltaic

effect (BPV) [2, 3]. BPV devices are not limited by the bandgap and therefore produce a larger

photovoltage (anomalous photovoltaic effect) than traditional junction-based PV devices [2‒

4]. In recent times, bulk ferroelectric polycrystalline photovoltaics have attracted special

attention because of their low cost, good stability, facile fabrication of the device, and high

photovoltage [2‒5]. The main drawback of ferroelectric ceramics is their low power

conversation efficiency (PCE) mainly due to their low photocurrent density under light which

is related to their low conductivity [2, 3, 5]. Therefore, how to simultaneously enhance the

photocurrent and photovoltage of bulk ferroelectric ceramics and thereby power conversion

efficiency becomes an important issue to pushing forward the practically realizable application

of the ferroelectric photovoltaic device. Recent research focused on the BPV effect in various

ferroelectric systems such as ferroelectric ceramics [2-5], single crystals [6, 7] thin films [8, 9,

10], polymers [11], organic and inorganic ferroelectric composites [12]. Recent research has

proposed explanations for the BPV with various models such as domain wall theory [13],

ballistic model [2], shift-current model [14], interface effect (Schottky- junction effect) [15]

electrode, and depolarization field effect [16]. Considering the above models, ferroelectric

materials with relatively small bandgaps and large polarizations are promising candidates for

photovoltaic applications because the materials can absorb large amounts of the solar spectrum

while simultaneously separating charge carriers.


In this regard, BaTiO3 (BT) is considered one of the promising eco-friendly

ferroelectric perovskite materials with off-center distortions from its centrosymmetric position

and large spontaneous polarization arising from the strong hybridization among the lowest

unoccupied Ti4+-d0 state and O-2p state at room temperature [17, 2]. However, pristine

ferroelectric BT can only be able to absorb the ultraviolet (UV) region of the solar spectrum

due to its wide bandgap (3.2 eV) [14, 17, 18]. Only 3.5% of solar radiation is encompassed by

UV light (wavelength 200‒390 nm) yet visible light accounts for 40% of solar energy

(wavelength 390-700 nm) [5]. Therefore, constructing a ferroelectric BPV device with a

narrow bandgap (Eg) is an essential factor for absorbing abundant light and maintaining

ferroelectricity. Recent studies have focused on substituting ions at A-site and/or partially

occupied non-d0 ions at the B-site of BT, as a way to reduce its bandgap as well as to control

ferroelectric properties. In this regard, a large short-circuit current density of Jsc=7.51 μA/cm2

was obtained with A-site Sr-doped BT, but it didn’t induce anomalous photovoltage (Voc<1.5

V i.e. Eg = 3.04 eV) [19]. However, the maximum Voc of about 16 V was attained by BT co-

doping with Bi and Li ions at the A-site, but the determined photocurrent was within the range

of nano ampere [14]. Kola et al. achieved a large Voc of 16 V with a single substitution on B-

site as Ag/BaTi1-xSnxO3/Ag PV device based on the shift current model, but the current density

was very low around ~2 nA/cm2 range [2]. However, there have been only a few reports on the

effect of co-substitution on the B-site of bulk ferroelectric photovoltaic materials up to now.

For example, L.Wu et al. reported that Ni2+ and Nb5+ co-doped (0.9)BaTiO3‐

(0.1)Ba(Ni1/2Nb1/2)O2.75 provided a decreased bandgap of ~1.5 eV and maintained remnant

polarization ~5 μC/cm2 with obtained steady-state current of 8 nA/cm2 [20]. In contrast to other

transition metal pair substitutions (such as Mg3+-Nb-5+, Co3+-Nb5+, and Sc3+-Nb5+), the Fe3+-

Nb5+ pair at the B-site of BT has little effect on the bandgap reduction, and the significant
suppression in ferroelectricity is be expected to reduce the photovoltaic properties of BT and

decrease the device performance [21, 22]. However, there is a lack of experimental evidence.

Also, it has been noted that in conventional electrode/ferroelectric/electrode

architecture, the net build-in electric field (Ebi) of the top and bottom metal electrodes can form

two back-to-back Schottky barriers that deplete the electrode-ferroelectric interface [15, 23].

From the prior reports, lowering the height of the Schottky barrier with a low-work function

metal electrode can induce a large photovoltaic effect of the device [15, 23]. Therefore, by

choosing a suitable configuration of electrodes in contact with ferroelectric materials, the

photocurrents and photo-generated voltage of the device may also be enhanced. In this regard,

aluminium (Al) and silver (Ag) electrodes have been selected as the top and bottom electrodes

respectively, for our PV device. Considering that as a scope, we used a ferroelectric material

such as BaTiO3 to provide the bandgap reduction, enhance photocurrent, photovoltage and

maintaining ferroelectricity through charge-neutral chemical co-substitution of aliovalent Fe3+

and Nb5+ at B-site with an asymmetrical electrode architecture. Both Nb5+ and Fe3+ perform a

typical charge compensator and bandgap reducers respectively, with strong photo response,

high electron mobility, reduced recombination loss, easy crystallization, controlled oxygen

vacancies and good candidates in the ferroelectric photovoltaic system. In particular, the

reduction of the bandgap from 3.2 eV to 2.55 eV and enhanced photocurrent density of 1.46

µA/cm2 have been achieved with Al/0.075BFNT/Ag device, which is larger than that of the

Ag/BBLT/Ag [17] and Ag/BST/Ag photovoltaic cells [2]. We perform theoretical

computations to further envisage our experimental results and to find the reason for the

bandgap reduction.

2 Experimental procedure
The bulk ceramics powder with the composition Ba(Ti1-2xFexNbx)O3 (BFNT) (x=0.0

0.025, 0.05, 0.075) was fabricated by solid-state reaction method (SSR) and represented as BT,

0.025BFNT, 0.05BFNT and 0.075BFNT. The starting materials of BaCO3 (99.9%), TiO2

(99%), Fe2O3 (99.9%), and Nb2O5 (99.99%) were weighted according to the stoichiometry of

each composition. The mixed powder was ground using a mortar and pestle for 30 min. The

mixture was calcined at 950oC/5 hr in the air atmosphere for all compositions. The calcined

powder was mixed with a binder of PVA, (polyvinyl alcohol) by manual granulation in a

mortar. The uniaxially compressed pellets are 13 mm in diameter and 1 mm thick under 7-ton

weight. The pellets were sintered at 300oC/1hr to remove the PVA and then sintered at

1300oC/2hr in a furnace. Room temperature X-ray diffraction measurements were carried out

using λ= 0.9654 Å the angle-dispersive X-ray diffraction (ADXRD) at the beamline-12

installed in Indus-2 synchrotron source RRCAT Indore, India. The FT-Raman spectrometer

(Bruker, RFS 27) was used to probe the structural phases of the samples. The instrument is

equipped with a Nd:YAG laser (1064 nm) as a source and an InGaAs detector (1024 × 256). It

has a resolution of 2 cm-1. The measured data was fitted with Lorentz function using Origin

software (version:8.5) and fitting parameters such as peak position and line width were added

in supporting information. To determine the bandgap of the samples, a diffuse reflectance

spectrophotometer (DRS) (UV-2400 Schimadzu) was used. The silver paste was applied on

both sides of the polished pellets for measuring the room temperature polarization-electric field

(P-E loop) hysteresis loops using a Precision ferroelectric tester system (Radian technologies

P-E loop tracer). The PHI - VERSAPROBE III is an advanced instrument that was used to

obtain X-ray photoelectronic spectra. It features a multichannel detector and monochromatic

Al-Kα source with energy of 1486.6 eV, providing excellent resolution ranging from 0.1 to 40

eV for sample analysis and characterization. The peaks background were subtracted for more

accuracy in peak fitting, then the Gaussian function was used to fit each elemental XPS peaks
with Origin software (version: 8.5) and the parameters were included in supporting

information. The Energy Dispersive X-ray Spectrometer Quantax 200 with X Flash® 6130 was

used for EDX and mapping studies. The photovoltaic measurements were done using Sol3A

Class AAA solar simulator with Keithley electrometer (2400) measuring setup.

2.1 Device fabrication

The PV studies were conducted on the cell configuration of Al/BFNT/Ag, where Al

and Ag are the top and bottom electrodes, respectively. BFNT pellets were mirror polished and

thickness of 0.5mm, diameter of 11mm and area of pellets of ~0.95 cm 2. For photovoltaic

measurements, Aluminium (Al) electrode was coated on the top surface of the sample (0.1-mm

thickness of shadow mask) and bottom surface was covered with silver (Ag) electrode using a

thermal evaporation technique with heating the samples at 400oC/30min. Ag paste was used to

connect the electrodes and copper wire for the photovoltaic experiment.

2.2 Computational Details

Density functional theory (DFT) simulations were performed to understand the

substantial decrease in the measured bandgap value of BFNT samples. Calculations were

carried out on a 3×3×3 supercell of 135 atoms using the projector augmented wave (PAW)

[24] method as implemented in the Vienna Ab Initio Simulation Package (VASP) [25]. The

generalized gradient approximation (GGA) [26] was used to depict the exchange and

correlations of electrons. A Monkhorst pack scheme [27] k-points mesh was set to 2×2×2, and

a plane wave energy cut-off of 520 eV was chosen. To account for any possible strong

correlation GGA +U approach was employed, with U = 4 eV for the 3d-orbitals of the Fe-ions

[28]. Since GGA functionals tend to underestimate the bandgap values for most materials, HSE

hybrid functional is used on optimized structures to compute the density of states (DOS) [22,

32]. DOS was sampled with 3×3×3 k-mesh and Gaussian smearing of 0.1 eV.
3 Results and discussion:

3.1 Synchrotron Angle Dispersive X-ray diffraction (ADXRD)

The synchrotron X-ray diffraction patterns of BT, 0.025BFNT, 0.05BFNT and

0.075BFNT samples were measured at room temperature and are shown in fig. 1(a). No

impurity phase was observed in all samples. Peak indexed for ceramics samples agrees with

recently reported BFNT [30, 31], which has a perovskite cubic phase with space group Pm3̅m

because of an increase in co-substitution. The gradual decrement in the (200)pc peak splitting

is evidence as seen in the enlarged view in fig 1(b). The two-phase model was confirmed for

BT samples from synchrotron ADXRD is consistent with laboratory X-ray diffraction

described in our prior study [31, 32], the two-phases of the BT sample are tetragonal (P4mm)

and orthorhombic (Amm2).

The Rietveld analysis program Jana2006 package [33] was used for crystal structure

refinement. Rietveld refinement parameters for specified space groups are the lattice

parameters, atomic coordinates, atomic occupancies, scale factor, 2θ zero shift, 36-degree

Legendre polynomials for background and pseudo-Voigt line shape with asymmetry. As per

the nominal composition of the perovskite ABO3 structure, the A site atomic occupancy is Ba,

and the B site atomic occupancy is Ti/(Fe & Nb). Figure 1 (c) and 1 (d) show the Rietveld

refinement fit of the synchrotron ADXRD pattern of pure BT using a single-phase model,

tetragonal (P4mm) and two-phase model tetragonal (P4mm) + orthorhombic (Amm2)

respectively. The goodness of fit (χ2) parameters for the P4mm and P4mm+Amm2 models are

χ2=5.71 and 4.71, respectively. The poor fit or misfits (Bragg position, peak shape, intensities)

could not be improved satisfactorily, as shown in the insets of fig. 1(c) [indicated by circle]. In

contrast, P4mm and Amm2 two-phase models could well account for the Bragg profiles, peak

position, and intensities. Fig. 1(d) illustrates the excellent fit in an enlarged view of (111)pc and
(200)pc for the two-phase model. Synchrotron ADXRD patterns of co-substituted BT, depicted

in fig. 1(e)-(g), were refined with P4mm (T) and Pm3̅m (C) structural models which conform

with earlier reports [30, 31]. Rietveld refined structural parameters such as lattice parameter,

volume, c/a, atomic position, R-factor (wRp), phase fraction (%wt), and goodness of fit (χ2) of

BFNT (x=0.0, 0.025, 0.05, and 0.75) samples are listed in table S1. The ratios of phase

percentage of P4mm and Pm3̅m crystal structure are 85.3:14.7, 76.7:23.3 and 41.5:58.5 for x=

0.025, 0.05, and 0.075 respectively. The mixed phase structures of BT and BFNT samples were

then verified by Raman spectroscopy as shown in Fig. S1(a–d), and a detailed description is

given in the Supporting Information.

According to Vegard’s law [34], the lattice constant of the cubic unit cell increases with

the concentration of substitution ions, as shown in table S1. Also, the volume of the cubic

lattice increases with x due to larger ionic radii (rFe3+=64.5 pm and rNb5+= 64 pm) in weight

compared to the host radius Ti (rTi4+= 60.5 pm). The c/a ratio of the tetragonal unit cell

constantly drops with increasing x, which can be attributed to the aliovalent replacement ions

such as Fe3+ and Nb5+ at the Ti4+ site that induces an internal stress field in BT ceramics. The

structural phase transition originated from this internal stress field by the broken Ti-O

hybridization. The change in lattice parameters and the volume of the unit cell is associated

with oxygen-octahedral tilting and distortion of octahedra. The change in lattice parameters

and the volume of the unit cell is associated with the oxygen-octahedral tilting and distortion

of octahedra. The oxygen-octahedral tilt can be determined by the equation given below [35,

36]:

180−𝜃
𝜃𝑡𝑖𝑙𝑡 = (1)
2

where 𝜃 is the angle between two Ti4+ ions in adjacent unit cells bonded with a face-

centered oxygen ion (Ti4+-O2--Ti4+). The octahedral tilting of BT compared with


0.075BFNT is illustrated in fig.2(a) and (b). Also, the octahedral distortion can be estimated

by the following relation [35]:


6
1 𝛿𝑛 − < 𝛿 > 2
∆𝑜𝑐𝑡 = ∑ [ ]
6 <𝛿>
𝑖=1

(2)

where 𝛿𝑛 and < 𝛿 > denote the individual and average interatomic metal-oxygen bond length.

With increasing substitution at the B-site, the structure of BFNT ceramics is closer to pseudo-

cubic with reducing octahedral tilting and distortion from table S2. The decreasing octahedra

tilt could affect the bandgap of the samples.

3.2 X-ray photoelectron spectroscopy

To investigate the electronic states of each element in BaTi1-2xFexNbxO3, the X-ray

photoelectronic spectra were recorded. The detailed survey and core-level spectra of Ba and Ti

for BT, 0.025BFNT and 0.075BFNT are discussed in the supporting information (fig. S2). Fig.

3(a) denotes that the binding energy of the Ti-2p spectrum shifted into a lower energy level

with increasing substitution. This can be attributed to the local electric field. This local electric

field induced to decrease in the bandgap of the ferroelectric material [37]. As a result, the

lowering bandgap influenced the intermediate charge transfer, thus can suppress the resistivity

of the material.

Fig. 3(b) illustrates the asymmetric nature of core-level Fe-3d spectra of 0.025BFNT

and 0.075BFNT and it can be split into Fe-2p3/2 and Fe2p1/2 spin states. For 0.025BFNT, the

XPS spectra can be deconvoluted into the characteristic peaks of Fe2+ and Fe3+ assigned to ~

708.8 eV, ~713.7 eV and ~ 710.8 eV, 722.06 eV, respectively. The absence of Fe2+ peaks in

0.075BFNT indicates the suppression of the oxygen vacancies. Additionally, the satellite peaks

around ~718.12 eV and 720.1 eV were also found in fig 3(b). These results are matched with
previous literature [37, 38]. The absence of doublet peaks around 719.9 eV and 706.7 eV

indicates that the substituted iron (Fe) is ionized and not metallic [39]. The photoelectronic

spectra around ~204.3 eV and ~208.3 eV represent the Nb-3d5/2 and Nb-3d3/2 states

respectively, were seen in fig. 3(c) for 0.025BFNT and 0.075BFNT samples. The oxidized state

of Nb is 5+ for these peaks.

The core-level spectra of O-1s of BT, 0.025BFNT and 0.075BFNT are shown in fig.

3(d). The O(I), O(II), and O(III) spectra represent the lattice oxygen bonding with Ti/Fe/Nb,

oxygen vacancies, and adsorbed oxygen is assigned at ~529 eV, ~531 eV, and ~533 eV

respectively. The binding energy and area under the fitting curve of O(I), O(II), and O(III) are

listed in table 1, where the area of the oxygen vacancy tends to decrease with increasing

substitution. This results from the simultaneous doping of donor (Nb5+) and acceptor (Fe3+) at

the B-site can neutralize the charge carriers and avoid the oxygen vacancies [40]. The reaction

can be written using Kroger-Vink notation as follows;

′ •
𝐹𝑒2 𝑂3 + 𝑁𝑏2 𝑂3 → 2𝐹𝑒𝑇𝑖 + 2𝑁𝑏𝑇𝑖 + 6𝑂𝑜 (3)

Hence, the samples are obeying the following condition of charge neutrality [41]

∑𝑛𝑖=1 𝑥𝐴𝑖 𝑛𝐴𝑖 + ∑𝑛𝑘=1 𝑥𝐵𝑖 𝑛𝐵𝑖 = 6 (4)

where, 𝑥𝐴𝑖 , 𝑥𝐵𝑖 (for x= 0.0, 0.025, 0.05 and 0.075) are the fractions of ions and 𝑛𝐴𝑖 , 𝑛𝐵𝑖 are the

valances of ions (Ba-2+, Ti-4+, Fe-3+, Nb-5+) at the A & B sites, respectively.

3.3 Hysteresis loop

To find the impact on ferroelectricity under B-site substitution of BT ceramics, the

Polarization-Electric field (P-E) hysteresis loop was measured at room temperature with a 30

kV/cm applied field, as shown in fig. 4. The shape of the P-E hysteresis loop tends to be

unsaturated with increasing metal substitution. The parameters such as remnant polarization
(Pr), spontaneous polarization (Ps), and coercive field (Ec) were decreased with increasing x as

in fig.4. It was suggested in the literature that the reduction of ferroelectricity upon metal

doping due to the formation of oxygen vacancies leads to an increase in the leakage current in

ferroelectric samples [32]. However, in the current study, the production of oxygen vacancies

is strongly restricted by the charge compensator Nb5+ which is further validated by the XPS

results of the Fe-2p and O-1s spectra. Therefore, the ferroelectricity in BT ceramics was

suppressed by the following features. The origin of local polarization in BaTiO3 ceramics can

be attributed to the off-center displacement of d0 Ti-ion about the oxygen octahedral cage. The

inclusion of bigger-size ions helps reduce the octahedral distortion and leads to suppressing the

Ti-O off-center displacement of BT ceramics. According to modern ferroelectric theories, the

partially occupied transition metal d-ion replaced at Ti-site supports centrosymmetric and

therefore simultaneously reduces the off-center distortion needed for ferroelectricity [42, 43].

The following theoretical calculations demonstrate the proof that the non-d0 Fe3+ state perturbs

both the valence and conduction band of BFNT materials. Therefore, the ferroelectricity of BT

was destabilized by the alterations in the hybridization of Ti-3d to O-2p orbital. These results

are consistent with the structural discussions and indicate that the ferroelectricity was

suppressed by co-substitution.

3.4 Bandgap

The DRS was used to investigate the optical characteristics of BT and metal-substituted

BT in the wavelength region of 200-900 nm as shown in fig. 5 (a). Fig 5 (a) illustrates the

absorbance spectra of the samples; in which all the substituted samples exhibit a shoulder

around ~400-500 nm (encircled region) and its absorbance increases along with x and it

broadens toward lower energy ranges. The bandgap energy of BT and BFNT samples can be

estimated using the Tauc plot as shown in fig. 5(b) and listed in table 2. The optical bandgap

of pure BT (3.25 eV) decreases with increasing metal substitution at the B-site. According to
previous studies, the valence band (VB) is assigned to the O-2p orbital and that of the

conduction band (CB) to the Ti-3d0 orbital in pure BaTiO3 [44], which generates a bandgap

energy of 3.25 eV. Meanwhile, non-d0 metals like Fe3+ replaced on the B-site might affect the

electronic band structure due to the anisotropic crystal fields of the octahedron created by the

metal-oxygen interactions [45]. In this context, the Fe/Ti/Nb could be treated as a point charge,

and their spatial arrangement affects the energy of the d-orbital of the central metal atom.

According to crystal field theory, a partially occupied 3d orbital in an octahedral crystal field

is divided into two eg orbitals and three t2g orbitals [2]. This field causes eg to evolve into one

high energy state (dx2-dy2) and one low energy state (dz2), whereas t2g evolves into one high

energy state (dxy) and two low energy levels (dzy and dzx) [45, 46]. As a result, the existence of

the 3d state between CB and VB with Fe3+ ions is indicated by the inter-band transition from

dz2 to dx2-dy2. Previous reports suggested that the large electronegativity difference between

metal and oxygen can induce a larger bandgap in perovskite materials [46, 47]. Here, the

electronegativity difference of Nb-O (1.84) and Fe-O (1.61) is much less than that of Ti-O

(1.9), therefore the reduced bandgap of BFNT largely depends on Fe3+ metal ion.

The d-orbital metal ions and the p-orbital of oxygen ions can be defined using the linear

combination of atomic orbitals model (LCAO). If the density of the state is deemed constant,

the following equations show that the probability of an average d-orbital mixing (rd) in the

valence band can be calculated based on the bandgap (Eg) and bandwidth (EB) [48, 40];

1 𝜂
𝑟𝑑 = 2 + 2(1−𝜂) ln(𝜂) × 100 (5)

where 𝜂 means the covalent mixing and it lies between 0 and 1. The value of 𝜂 can be

determined by the following equation:

𝐸𝑔
𝜂 = 2𝐸 (6)
𝐵
From table 2 𝜂 values tend to be small with increasing metal substitution indicating the strong

covalent mixing.

The value of EB is found from the equation below:

𝐸𝑔 2
𝐸𝐵 = √( 2 ) + 8(𝑝𝑑𝜋)2 (7)

For perovskite samples, the value of 𝑝𝑑𝜋 =1 eV [45] and the calculated𝐸𝐵 , 𝜂 and 𝑟𝑑 are listed

in table 2. The results obtained here are evidence that the mixing of d-orbitals into valence

bands is a function of bandgap (𝐸𝑔 ). The decreasing bandgap suggests an increased probability

of d-orbitals mixing into the valence band, which is verified by the DFT calculations as shown

below.

3.5 Density functional theory

A 3×3×3 supercell was constructed using the experimental lattice parameters of cubic

BFNT. Four Ti ions in the supercell were replaced by two Fe and two Nb ions to match the

experimental doping concentration (x=0.075). Based on the dopant-dopant distance, eight

distinct configurations were identified as illustrated in fig.S3 and the ground state energy

calculations were performed to understand the effect of the dopant ion interaction over the

ground state energy. It is observed that the calculated ground state energies differ relatively, as

tabulated in table S5 and the optimized structure of the minimum energy configuration-2 is

used to calculate DOS.

Figure 5 (c, d) shows the computed electronic density of states corresponding to the

tetragonal-BT and cubic-0.075BFNT structures in their minimal energy configuration. The

calculated bandgap values for pure BT and 0.075BFNT were 3.26 and 2.49 eV, respectively.

These values are in line with the experimental results of 3.2 and 2.55 eV. The valence band of

pure BT is mostly occupied by the O-2p state, whereas the conduction band consists primarily
of the Ti-3d state, as seen in Fig. 5(c). The estimated valence bandwidth is -4.75 eV, which is

consistent with the DFT results from previous papers [49, 50].

It is remarkable to notice from fig. 5(d) that on metal substitution at B-site, the Ti-3d

and O-2p orbital still dominate the CB minimum and VB maximum of the BFNT respectively.

With increasing substitution level to x=0.075, the formation of new 3d states of Fe and Nb was

observed at both the valence and conduction bands as shown in fig. 5(d). The PDOS in Fig.

5(d) shows the hybridization of Fe ion 3d-states with BT host electronic states, close to the

valence band maximum. The incorporation of Fe-3d states localized in VB has resulted in

stronger hybridization with the O-2p state near the valance band edge than that observed for

Ti–O hybridization. HSE calculations clearly show the formation of new states at the top of the

valence band edge with predominant Fe 3d character, which primarily decreases bandgap.

Additionally, strong d–d interaction between Ti–3d and Fe/Nb – d state at top of VB has led to

the accumulation of carriers near bottom CB as well as top VB edges. Also, the contribution

of lower Fe-3d (~ -7.2 to -6.7 eV) state increases the width of VB from ~ -4.75 eV (i.e. BT) to

~ -8 eV as shown in fig. S4. Such enhancement is beneficial for the mobility of the

photogenerated charge carriers within the material system as from an earlier report [51]. On

the other hand, fig. 5(d), the substitution of Fe and Nb in BFNT results in the hybridization of

Nb–O and Fe-O at the conduction band, which effectively separates the photogenerated

electron-hole pair [52]. Furthermore, this hybridization between Fe-3d and Nb-3d with O-2p

pushes the CB minimum towards lower energy levels resulting in the significant reduction of

the conduction band edge from 3.26 to 2.49 eV, as shown in Fig. 5(d) in the reason for

minimizing the photoexcitation energy of charge carriers. The red shift of CBM combined with

an increase in mobility of photogenerated charge carriers along VB due to Fe–O & Nb–O in

BFNT leads to the expectation of improvement in photocurrent density.


3.6 Photovoltaic characteristics:

To evaluate the photovoltaic (PV) response of BT ceramics with Fe and Nb substitution,

the short-circuit current density (Jsc) vs open-circuit voltage (Voc) curve were measured under

both dark and light irradiation of 100 mW/cm2 and the results are illustrated in fig. 6(a).

Surprisingly, all the BT samples exhibit bulk photovoltaic characteristics with Jsc in

microampere and large Voc as listed in table 3. It was noted that the open-circuit voltage and

short-circuit current density increased under increasing the substitution and achieved 8.31 V

and 1.46 µA/cm2 respectively, for 0.075BFNT. The large Voc (i.eVoc>Eg) justifies that it is an

anomalous photovoltaic effect. Further, the following relation can be used to calculate the

photoresponsivity (R) and specific detectivity (D) of the device [53]:

𝐽𝑠𝑐
𝑅= (8)
𝐼

𝑅
𝐷 = [(2𝑞𝐽 1/2 ] (9)
𝑑𝑎𝑟𝑘 )

where I means the light intensity (100 mW/cm2), q refers the electron charge (1.6 × 10-19 C)

and Jsc, Jdark stands for the short-circuit current density and dark current density of the device,

respectively. According to table 3 the maximum R and D values of 1.46× 10-2 mA/W and 3.57

× 108 Jones were obtained for Al/0.075BFNT/Ag device, respectively, which is remarkably

larger than prior reports [53, 54].

The time dependent photocurrent density curve for all samples was measured under 100

mW/cm2 light irradiation with an ON/OFF switching time of 60 s and a bias of 10 V, as shown

in Fig. 6(b). It was found that the photocurrent density of all ceramic samples increased when

the light was turned on, then decayed and returned to its original state once the light source was

turned off. This clearly shows that all of the devices have a good ON/OFF switching response.

Fig. 6(c) displayed the performance of time resolved photocurrent density of BT and
0.075BFNT device. The photocurrent of both devices increased quickly when the light source

was turned on, followed by a modest improvement in Jsc, until achieving saturation. After the

light source is switched off, the photocurrent was rapidly falls for both the devices, then slowly

decays until returning to the initial condition of Jsc (dark current). Such rise and decay responses

of both devices were detected using exponential fitting of the experimental data using the

relation below;

−𝑡 −𝑡
𝐼 = 𝐼0 + [𝐶 𝑒𝑥𝑝 ( 𝜏 )] + [𝐷 𝑒𝑥𝑝 ( 𝜏 )] (10)
1 2

where I0 represents the steady state photocurrent, C and D are constants, t is time and 𝜏1 and 𝜏2

are the relaxation time constants [55]. The excellent fitted photo response curves of BT and

0.075BFNT samples as presented in fig. 6 (c) where 𝜏𝑟 and 𝜏𝑑 are time constants for rising and

decay edges, respectively. The estimated time constants of BT are: 𝜏𝑟1 = 7.5 s, 𝜏𝑟2 = 17.3 s,

𝜏𝑑1 = 10.1 s, 𝜏𝑑2 = 12.3 s and 0.075BFNT are: 𝜏𝑟1 = 7.0 s, 𝜏𝑟2 = 8.6 s, 𝜏𝑑1 = 10.6 s, 𝜏𝑑2 = 13

s. The rising constants are associated with the carrier creation process, whereas the decay

constants are associated with the carrier recombination process. As a result, BT has a longer

duration for carrier generation than 0.075BFNT, yet the decay constants of both devices are

comparable.

The PV properties of bulk BT ceramics depend on various factors such as material

thickness [56], the work function of the electrode material [23], light intensity [14], applied

poling field [17], electrical conduction and behavior of domain walls [13], ferroelectric

polarization [17] and electrode/ferroelectric/electrode interface [15, 23]. However, the bulk

photovoltaic effect is largely dependent on the interface of the top electrode/ferroelectric

material/bottom electrode architecture. According to Fan et al. the comparative study of the

bulk photovoltaic effect on Ti/BTO/SRO and Pt/BTO/SRO combinations revealed a short-

circuit current(Isc) of about 1.2 nA and 0.3 nA, respectively [57]. This result suggested that the
suppression of photocurrent in Pt/BTO/SRO interface was due to the large work function of

the Pt electrode (φPt = 5.5 eV) than Ti electrode (φTi = 4.33 eV). A similar PV study was

conducted for Mg/PLZT/ITO, Pt/PLZT/ITO, and Ag/PLZT/ITO cell configurations, where the

open-circuit voltage and short circuit current density increased in the order of large work

function metal electrodes to less work function metal electrodes [19]. Moreover, Subhajit Pal

et al. reported that the pure BT ceramics with a cell configuration of Ag/BT/Ag provide a Voc

of 2 V and Jsc of 3 nA/cm2 under the illumination of 160 mW/ cm2 with the device thickness

of about 0.5 mm [14]. Although in our case, Al/BT/Ag provides Jsc of 0.65 μA/cm2 and Voc of

3.4 V even under 100 mW/cm2 illumination with a similar thickness as pure BT. Here, the

height of the Schottky barrier (φB) can be determined by the difference between the work

function of the metal electrode (φM) and the electron affinity (χ) of FE material [58].

Considering that the work functions of Al and Ag are 4.125 eV and 4.48 eV respectively and

the electron affinity of BT is 3.8 eV [59], then φB= 0.325 eV (if Al is a top electrode) and φB=

0.68 eV (if Ag is a top electrode). Comparatively, the barrier height is much smaller when Al

works as the top electrode of the device than Ag. In addition, the enhancement of the PV

response in pure BT can originate from the coexistence of the tetragonal and orthorhombic

phases. According to the shift current theory, the enhancement of PV properties in KNbO3 is

achieved in excess in mixed of orthorhombic and rhombohedral phases compared to that of the

tetragonal phase [60]. Moreover, Kola et al. reported that maximum Voc was attained in B-site

doped BT at a larger orthorhombic phase than that of the tetragonal single phase [2]. Therefore,

the greater bulk PV response in pure BT in this study than in Subhajit Pal et al. [14] result is

based on the structural symmetry and low work function metal electrode.

The greatest variation in PV outcomes from metal substitution depends on the bandgap,

grain size, domain size, and dielectric and ferroelectric characteristics [13]. The existence of

an aliovalent Fe3+-Nb5+ pair in BaTiO3 greatly alters the properties of both the valence and
conduction bands, according to theoretical DFT calculation. The bandgap of the ferroelectric

material decreased considerably from 3.25 eV to 2.55 eV when B-site replacement increased.

This reduction in bandgap potentially absorbs the maximum visible region of the light source

as shows in fig. 5(a). Bandgap reduction is an important component covering a wide area of

the light source and is beneficial to increase the current density of the device. Additionally, the

XPS findings of fig. 3(b) demonstrated that the formation of minor oxygen vacancies was

obtained in 0.025BFNT, however in the 0.075BFNT sample, the oxygen vacancies were

substantially controlled by the lack of Fe2+ state. The oxygen vacancy suppression in

0.075BFNT increases the Jsc of the device more sharply than that in 0.025BFNT by reducing

the recombination between oxygen vacancies and hopping electrons.

Meanwhile, the open-circuit voltage of the device can be expressed as [61]

𝑉𝑜𝑐 = 𝐸𝑝𝑣 . 𝑑 (11)

where Epv denotes the photovoltaic field and d is the distance between the metal electrodes.

This equation suggests an anomalous photovoltaic effect, in which the open-circuit voltage of

the device is substantially higher than the bandgap of the material and is not constrained by the

bandgap of bulk ferroelectrics. For instance, A. M. Glass et al. in Fe-doped LiNbO3 [62]

attained the anomalous photovoltage of around 1,000 volts. Generally, the PV response of bulk

ferroelectrics depends on the grain size and domain size effect. In an earlier study, we described

the microstructure of Fe and Nb-modified BaTiO3. Currently, the microstructure of

0.075BFNT was analyzed and the results are shown in Fig. S5. Additionally, elemental analysis

of all samples was conducted using EDX mapping & spectra as depicted in Fig. S6; further

details regarding microstructure analysis can be found in Supporting Information. According

to Frank et al. it has been proposed that the generated high photovoltage in ceramics is

proportional to the number of grains, i.e., high photovoltage is generated from the summation
of individual photo-emf across the grains [63]. Consistent with this result, the number of grains

in smaller grain sizes is greater than in larger ones, resulting in higher output photovoltage with

increasing metal substitution. In addition, the grain size of BaTiO3 is co-related with the

domain size by the simple relationship as follows [64]

𝑑 ∝ 𝑔1/2 (12)

where g and d denote the grain and domain size of the ceramics, respectively. The domain size

will also be affected by changes in grain size. Earlier reports stated that the open-circuit voltage

of ferroelectrics was induced by the addition of an electrostatic potential difference between

the domain walls [13]. Here the grain size of BaTiO3 ceramics was decreased upon metal

substitution in the order of 3.6 μm, 0.68 μm and 0.51 μm for x=0.0, 0.025, and 0.05

respectively, from our previous findings [31] and fig. S5 reveals the grain size of x=0.075 was

0.14 μm. Thus, the smaller grain size results in smaller domain size and the smaller size of

domains have a larger number of domain walls and achieve a higher open-circuit voltage of

the device. As a result, the reduced grain size increased the photovoltage of the Al/BFNT/Ag

device.

The aforementioned findings demonstrated an inventive strategy to enhance both the

photocurrent and photovoltage of BFNT ceramics. Numerous investigations have

indicated that the large ferroelectric polarization is a crucial component for the anomalous

photovoltaic effect, in which the photovoltage of the ferroelectric material increases

substantially above the bandgap of the material meanwhile the generated photocurrent values

are very low in the range of nano to pico ampere [17, 24, 25]. Even for the pure BT sample,

the anomalous photovoltage of around 3.4 V was attained due to the larger polarization and

non-centrosymmetry orthorhombic + tetragonal phase structure, as confirmed by P-E loop and

ADXRD analysis respectively. However, the large anti-bonding property between Ti-3d and
O-2p results in a wide bandgap (Eg=3.25 eV), hence the generated photocurrent was quite low

(Jsc= 0.65 μA/cm2). According to the first-principle calculations based on the shift current

mechanism, maximum polarization is not a necessary condition for the BPV response, although

inversion symmetry must be broken [14, 65]. According to the above mechanism, the non-

centrosymmetric ferroelectric tetragonal phase coexists with the paraelectric cubic phase in

metal-substituted BT, as proven by ADXRD and Raman investigations, where the remnant

polarization was reduced compared to pure BT, but the photovoltaic parameters and PV power

(Voc × Jsc) of the 0.075BFNT device were remarkably improved and comparable with recent

studies, as shown in fig. 6(d) and table S7. Therefore, we concluded from the overall

experimental results and theoretical DFT calculations that the high photoresponsivity, large

detectivity of solar spectrum with concurrent enhancement of both photocurrent and

photovoltage in bulk ferroelectric BT is possible through appropriate co-substitution of

aliovalent Fe3+-Nb5+ pair, which is successfully tuning the bandgap to absorb abundant regions

of the solar spectrum, controlling oxygen vacancies to reduce the e-h recombination rate,

minimizing the grain size and using a low work function top electrode.

4. Conclusion:

In summary, the structural, optical, electrical and photovoltaic properties of BaTi1-

2xFexNbxO3 (x=0.0, 0.025, 0.05 and 0.075) were discussed. The synchrotron ADXRD data

combined with Raman spectra confirmed that the pure BT possesses the mixed phase of

tetragonal + orthorhombic at RT. Increasing the Fe3+ and Nb5+ substitution at B-site modifies

the structure to tetragonal + cubic symmetry, reducing octahedral tilt, and distortion along with

increasing internal stress field contributed to tuning the bandgap into a lower energy state of

2.55 eV and thus can efficiently absorb the larger area of the solar spectrum (~ 500 nm). To

comprehend the drastic decrement in the bandgap of BFNT samples from pristine BT, DFT

studies were carried out. It was found that the band gap value decreases drastically from 3.28
eV to 2.49 eV because of the aliovalent Fe3+-Nb5+ ions contributions to the valance and

conduction band. The tuned bandgap develops the metal-oxygen covalent character and

enhances the charge carrier mobility of the device. Further, XPS analysis depicted a clear

picture about suppression of oxygen vacancies in 0.075BFNT through Fe-2p and O-1s spectra.

Meanwhile, the grain size reduction is responsible for the continuous growth of Voc. From time

resolved J-t fitted curves, the slow response rise time constant was improved for 0.075BFNT

than BT device. Here, the simultaneous enhancement of Jsc and Voc does not follow the

maximum polarization concept. Finally, an apt top electrode, reduced bandgap, small grain size

and suppression of oxygen vacancies are contributed to enhance the PV power of

Al/0.075BFNT/Ag device to ~12 μW/cm2.

Acknowledgment:

This study was financially funded by the University Grant Commission-Basic Science

Research Fellowship scheme (F.No:25-1/2014-15(BSR)/7-305/2010/(BSR), and L. Venkidu is

thankful to the government of India for funding this study. The author also thanks Dr. A.K

Sinhan, Dr. Archana Sagdeo, and Mr. Mandvendra Narayanan Singh, Raja Ramanna Centre

for Advanced Technology, Indore, India, for their help in Angle Dispersive X-ray Diffraction

(ADXRD) beamline-12 (BL-12) at Indus-2 synchrotron radian source. The fabrication of the

mask (Al/BFNT/Ag) was carried out at the Center for Nano Science and Engineering (CeNSE),

Indian Insitute of Science, Bangaluru, under the INUP program funded by the Ministry of

Education (MoE), Ministry of Electronics and Information Technology (MeitY), and

Nanomission, Department of Science and Technology (DST), Government of India. The author

L.Venkidu sincerely thanks Ms. Adithi for their help during the mask fabrication. SERB grant

(CRG/2019/006990) is also acknowledged for the P-E loop tester. N. Raja gratefully

acknowledges C. Kamal, Raja Ramanna Centre for Advanced Technology, Indore, India for
the fruitful discussions and help in calculations. N. Raja also acknowledges Rajesh Ravindran

and The Institute of Mathematical Sciences, Chennai, India.

Author Contributions

The corresponding author, Dr. B. Sundarakannan is responsible for ensuring that the

descriptions are accurate and agreed by all authors. The role(s) of all authors is listed as follows,

L.Venkidu: Conceptualization, Methodology, Software, Writing - Original Draft

N. Raja: DFT Validation and Investigation

B. Sundarakannan: Conceptualization, Methodology, Validation, Investigation, Writing -

Review & Editing

Data availability

The datasets generated and analyzed during the current study are available from the

corresponding author upon reasonable request.

References:

1. ArturoMorales-Acevedo, Variable band-gap semiconductors as the basis of new solar cells,

solar energy, 83 2009 1466-14771, https://doi.org/10.1016/j.solener.2009.04.004

2. L.Kola, D.Murali, S.Pal, B.R.K.Nanda, and P.Murugavel, Enhanced bulk photovoltaic

response in Sn doped BaTiO3 through composition dependent structural transformation, Appl.

Phys. Lett. 114 (2019) 183901, https://doi.org/10.1063/1.5088635

3. V.M. Fridkin, Bulk Photovoltaic Effect in Noncentrosymmetric Crystals, Crystallogr.

Rep. 46 (2001) 654–658, https://doi.org/10.1134/1.1387133


4. A. B. Swain, D. Murali, B.R.K. Nanda, and P. Murugavel, Large bulk photovoltaic response

by symmetry-breaking structural transformation in ferroelectric [Ba(Zr0.2Ti0.8)O]0.5[(BaCa-

)TiO3]0.5 , Phys. Rev. Applied 11 (2019) 044007,

https://doi.org/10.1103/PhysRevApplied.11.044007

5. X.He, C.Chen, C.Li, H.Zeng, and Z.Yi, Ferroelectric, Photoelectric and photovoltaic

performance of silver niobate ceramics, Adv. Funct. Mater. 29 (2019) 1900918,

https://doi.org/10.1002/adfm.201900918

6. Y. Noguchi and H. Matsuo, Ferroelectric photovoltaic tensor in visible-light-active Fe-doped

BaTiO3 single crystals, jpn. J. Appl. Phys. 60 (2021) SFFA01,https://doi.org/10.35848

/1347-4065/ac0c6c

7. Y. Liu, X. Wang, F. Fan, and C. Li, Bulk Photovoltage Effect in Ferroelectric BaTiO3, J.

Phys. Chem. Lett. (2022), 13, 48, 11071-11075, https://doi.org/10.1021/acs.jpclett.2c03194

8. A. Zenkevich, Yu. Matveyev, K. Maksimova, R. Gaynutdinov, A. Tolstikhina, and V.

Fridkin, Giant bulk photovoltaic effect in thin ferroelectric BaTiO3 films, Phys. Rev. B, 90

(2014) 161409, https://doi.org/10.1103/PhysRevB.90.161409

9. V. M. Fridkin, Ferroelectricity and Giant Bulk Photovoltaic Effect in BaTiO3 Films at the

Nanoscale, Ferroelectrics, 484:1, 1-13, https://doi.org/10.1080/00150193.2015.1059151

10. Z.Fan, K.Yao,and J.Wang, Photovoltaic effect in an indium-tin-oxide/ZnO/BiFeO3/Pt

heterostructure, Appl. Phys. Lett. 105 (2014) 162903, https://doi.org/10.1063/1.4899146

11. Q. Liu, I.Khatri, R.Ishikawa, A.Fujimori, K.Ueno, K.Manabe, H.Nishino,and H.Shirai,

Improved photovoltaic performance of crystalline-Si/organic Schottky junction solar cells

using ferroelectric polymers, Appl. Phys. Lett. 103 (2013) 163503,

https://doi.org/10.1063/1.4826323
12. Z.Fan, K.Sun and J.Wang, Perovskites for photovoltaics: a combined review of organic–

inorganic halide perovskites and ferroelectric oxide perovskites, J. Mater. Chem. A, 2015 3: 18809-

18828, https://doi.org/10.1039/C5TA04235F

13. H. Matsuo, Y. Kitanaka, R. Inoue, Y. Noguchi, M. Miyayama, T. Kiguchi, and T. J. Konno,

Bulk and domain-wall effects in ferroelectric photovoltaics, Phys. Rev. B 94 (2016): 214111,

https://doi.org/10.1103/PhysRevB.94.214111

14. S. Pal, A. B. Swain, P. P. Biswas, D .Murali, A. Pal, B. R. K. Nanda and P. Murugavel,

Giant photovoltaic response in band engineered ferroelectric perovskite, Sci Rep 8 (2018)

8005, https://doi.org/10.1038/s41598-018-26205-x

15. P. Zhang, D. Cao, C. Wang, M. Shen, X. Su, L. Fang, W. Dong, F. Zheng , Enhanced

photocurrent in Pb(Zr0.2Ti0.8)O3 ferroelectric film by artificially introducing asymmetrical

interface Schottky barriers, Mater Chem and Phys 135 (2012) 304-308,

https://doi.org/10.1016/j.matchemphys.2012.04.041

16. R .Gaoa, W. Caia, G. Chena, X. Deng, X. Caoa, C. Fua, Enhanced ferroelectric photovoltaic

effect based on converging depolarization field, Mater Res Bull 84 (2016) 93–98,

http://dx.doi.org/10.1016/j.materresbull.2016.07.031

17. P. Pal, K. Rudrapal, P. Maji, A. R. Chaudhuri, and D. Choudhury, Toward an Enhanced

Room-Temperature Photovoltaic Effect in Ferroelectric Bismuth and Iron Co-doped BaTiO3,

J. Phys. Chem. C 125 (2021) 5315−5326, https://dx.doi.org/10.1021/acs.jpcc.0c10655

18. K. G. Baiju, B. Murali, D. Kumaresan, Ferroelectric barium titanate microspheres with

superior light-scattering ability for the performance enhancements of flexible polymer dye

sensitized solar cells and photodetectors, Solar Energy, 224 (2021) 93–101,

https://doi.org/10.1016/j.solener.2021.05.063

19. Z. Bai, Q. Zhang, Y. Zhang, Z. Huang, Y. Gao, J. Liu, X. Wang, Enhanced photocurrent

of self-powered ultraviolet photodetectors based on Ba1−xSrxTiO3 ceramics via ferroelectric


polarization, J Alloys compd 885 (2021) 161177,

https://doi.org/10.1016/j.jallcom.2021.161177

20. L. Wu, A. R. Akbashev, A. A. Podpirka, J.E. Spanier, P.K. Davies, Infrared‐ to‐ ultraviolet

light‐ absorbing BaTiO3‐ based ferroelectric photovoltaic materials, J Am Ceram Soc. 102

(2019)1–12, https://doi.org/10.1111/jace.16307

21. S. Hao, M. Yao, G. Vitali-Derrien, P. Gemeiner, M. Otonicˇar, P. Ruello, H. Bouyanfif,

Pierre-Eymeric Janolin, B. Dkhil and C. Paillard, Optical absorption by design in a

ferroelectric: co-doping in BaTiO3, J. Mater. Chem. C, 10 (2022) 227-234,

https://doi.org/10.1039/D1TC04250E

22. S. Das, S. Ghara, P. Mahadevan, A. Sundaresan, J. Gopalakrishnan, and D.D. Sarma,

Designing a Lower Band Gap Bulk Ferroelectric Material with a Sizable Polarization at Room

Temperature, ACS Energy Lett. 3 (2018) 1176−1182,

https://doi.org/10.1021/acsenergylett.8b00492

23. J. Zhang, X. Su, M. Shen, ZhihuaDai, L. Zhang, XiyunHe, W. Cheng, M.Cao1 and G. Zou,

Enlarging photovoltaic effect: combination of classic photoelectric and ferroelectric

photovoltaic effects, Sci Rep 3 (2013) 2109, https://doi.org/10.1038/srep02109

24. B. Hammer, L. B. Hansen and J. K. Nørskov, Improved adsorption energetics within

density-functional theory using revised Perdew-Burke-Ernzerhof functionals, Phys. Rev. B, ,

59 (1999) 7413-7421, https://doi.org/10.1103/PhysRevB.59.7413.

25. G. Kresse and D. Joubert, From Ultrasoft Pseudopotentials to the Projector Augmented-

Wave Method Phys. Rev. B, 1999 59: 1758-1775,

http://dx.doi.org/10.1103/PhysRevB.59.1758.
26. J. P. Perdew, K. Burke and M. Ernzerhof, Generalized Gradient Approximation Made

Simple, Phys. Rev. Lett., 77 (1996) 3865-3868, https://doi.org/10.1103/PhysRevLett.77.3865

27. H. J. Monkhorst and J. D. Pack, Special points for Brillouin-zone integrations,

Phys. Rev. B, 13 (1976) 5188-5192, https://doi.org/10.1103/PhysRevB.13.5188.

28. H. K. Chandra, K. Gupta, A. K. Nandy and P. Mahadevan, Ferroelectric distortions in doped

ferroelectrics: BaTiO3:M (M=V−Fe), Phys. Rev. B, 87 (2013) 214110,

https://doi.org/10.1103/PhysRevB.87.214110.

29. R. Wahl, D. Vogtenhuber and G. Kresse, SrTiO3 and BaTiO3 revisited using the projector

augmented wave method: Performance of hybrid and semilocal functionals, Phys. Rev. B, 78 (2008)

104116, https://doi.org/10.1103/PhysRevB.78.104116.

30. Zhuo Wang, Xiang Ming Chen,Evolution from relaxor-like dielectric to ferroelectric in

Ba[(Fe0.5Nb0.5)1−xTix]O3, Solid State Commun, 151 (2011) 708-711,

https://doi.org/10.1016/j.ssc.2011.02.015.

31. Venkidu L, V.G.Babu M, E.Rubavathi P, Bagyalakshmi B, Sundarakannan B, Structure,

microstructure, magnetic and magnetodielectric investigations on

BaTi(1-x-y)FexNbyO3 ceramics, Ceram Int, 44 (2018) 8161–8165,

https://doi.org/10.1016/j.ceramint.2018.01.263.

32. P.E. Rubavathi, L. Venkidu, M.V.G. Babu, R.V. Raman, B. Bagyalakshmi, S.M.A. Kader,

K. Baskar, M. Muneeswaran, N.V. Giridharan, B. Sundarakannan, Structure, morphology and

magnetodielectric investigations of BaTi1−xFexO3−δ ceramics, J Mater Sci: Mater Electron 30

(2019) 5706–5717. https://doi.org/10.1007/s10854-019-00864-6.

33. Petricek,V., Dusek, M.Palatinus, L.Z.Kristallogr, Crystallographic Computing System

JANA2006: General features, Crystalline Mater. 229 (2014) 345-352,

https://doi.org/10.1515/zkri-2014-1737.

34. L. Vegard, Die Konstitution der Mischkristalle und die Raumfüllung der Atome,
Z. Phys. 1921 5: 17–26.

35. J.A.Alonso, M.J.Martı´nez-Lope, and M.T.Casais, Evolution of the Jahn-Teller Distortion

of MnO6 Octahedra in RMnO3 Perovskites (R ) Pr, Nd, Dy, Tb, Ho, Er, Y): A Neutron

Diffraction Study, Inorg. Chem. 39 (2000) 917-923, https://doi.org/10.1021/ic990921e

36. L. Venkidu, A.B. Athanas, S. Kalaiyar, B. Sundarakannan, Fabrication of DSSCs using

ferroelectric photoanodes of co-substituted (1-x)BiFeO3-(x)BaFe1/2Nb1/2O3: Structural

correlation to bandgap reduction and its impact on power conversion efficiency, Mater Letter

311 (2022) 131473, https://doi.org/10.1016/j.matlet.2021.131473

37. B. Luo, X. Wang, E. Tian, H. Song, Q. Zhao, Z.Cai, W. Feng, L. Li, Giant permittivity and

low dielectric loss of Fe doped BaTiO3 ceramics: Experimental and first-principles

calculations, J Eur Ceramic Soc, 38 (2018) 1562-1568,

http://dx.doi.org/10.1016/j.jeurceramsoc.2017.10.014

38. L. Srisombat, S. Ananta, B. Singhana, T. RandallLee, R. Yimnirun, Chemical investigation

of Fe3+/Nb5+-doped barium titanate ceramics, Ceram Inter 39 (2013) S591–S594,

https://doi.org/10.1016/j.ceramint.2012.10.142

39. B. Luo, X. Wang, E. Tian, H. Song, Q .Zhao, Z. Cai, W. Feng, L. Li, Giant permittivity

and low dielectric loss of Fe doped BaTiO3 ceramics: Experimental and first-principles

calculations, J Eur Ceramic Soc, 38 (2018)1562-1568,

http://dx.doi.org/10.1016/j.jeurceramsoc.2017.10.014

40. M.M. Ismail, Ferroelectric characteristics of Fe/Nb co-doped BaTiO3, Modern Physics

Letters B, 33 (2019) 1950261, https://doi.org/10.1142/S0217984919502610

41. Heywang, W., Lubitz, K. and Wersing, W. eds., 2008. Piezoelectricity: evolution and

future of a technology (Vol. 114). Springer Science & Business Media.


42. P. Pal, K. Rudrapa, S. Mahana, S. Yadav, T. Paramanik, S. Mishra, K. Singh, G. Sheet, D.

Topwal, A.R. Chaudhuri, and D. Choudhury, Origin and tuning of room-temperature

multiferroicity in Fe-doped BaTiO3, Phys. Rev. B 101 (2020) 064409,

https://doi.org/10.1103/PhysRevB.101.064409

43. N.A. Hill, Why Are There so Few Magnetic Ferroelectrics?, J. Phys. Chem. B 104 (2000)

6694-6709, https://doi.org/10.1021/jp000114x

44. D. Phuyal, S. Mukherjee, S. Jana, F. Denoel, M.V. Kamalakar, S.M. Butorin, A.

Kalaboukhov, H. Rensmo, and O. Karis, Ferroelectric properties of BaTiO3 thin films co-doped

with Mn and Nb, AIP Advances 9 (2019) 095207 https://doi.org/10.1063/1.5118869

45. M.S. Alkathy, M.H. Lente, J.A. Eiras, Bandgap narrowing of Ba0.92Na0.04Bi0.04TiO3

ferroelectric ceramics by transition metals doping for photovoltaic applications, Mater Chem

and Phys 257 (2021): 123791, https://doi.org/10.1016/j.matchemphys.2020.123791

46. M.S. Alkathy, J.A .Eiras, F.L. Zabotto, K.C.J. Raju, Structural, optical, dielectric, and

multiferroic properties of sodium and nickel co‑ substituted barium titanate ceramics, J Mater

Sci: Mater Electron 32 (2021) 12828–12840 https://doi.org/10.1007/s10854-020-03900-y

47. F. Wang, I. Grinberg, and A.M. Rappe, Band gap engineering strategy via polarization

rotation in perovskite ferroelectrics, Appl. Phys. Lett., 104 (2014) 152903,

https://doi.org/10.1063/1.4871707

48. T. Wolfram, S. Ellialtioglu, Electronic and Optical Properties of D-Band

Perovskites, Cambridge University Press, New York, 2006.

49. S. Dahbi, N. Tahiri, O. El Bounagui, H. Ez-Zahraouy, Electronic, optical, and

thermoelectric properties of perovskite BaTiO3 compound under the effect of compressive

strain, Chemical Physics, 544 (2021) 111105,

https://doi.org/10.1016/j.chemphys.2021.111105
50. H. Pinto, A. Stashans, Computational study of self-trapped hole polarons in tetragonal

BaTiO3, Phys. Rev. B 65 (2002) 134304, https://doi.org/10.1103/PhysRevB.65.134304

51. Hsuan-Chung Wu, Sheng-Hong Li, and Syuan-Wei Lin, Effect of Fe Concentration on Fe-

Doped Anatase TiO2 from GGA + U Calculations, International Journal of Photoenergy, 2012

(2012), 823498, https://doi.org/10.1155/2012/823498

52. Y. Q. Wang, Y. Liu, M. X. Zhang & F. F. Min (2018) Electronic, magnetic and optical

properties of charge-compensated (Nb, TM=Fe, Cr)-codoped SrTiO3 from first principles,

Ferroelectrics, 537:1, 68-78, https://doi.org/10.1080/00150193.2018.1528958

53. N. Ma, and Y. Yang, Boosted Photocurrent via Cooling Ferroelectric BaTiO3 Materials for

Self-Powered 405 nm Light Detection, Nano energy, 60 (2019) 95-102,

https://doi.org/10.1016/j.nanoen.2019.03.036

54. N. Ma, K . Zhang, Y. Yang, Photovoltaic-Pyroelectric Coupled Effect Induced Electricity

for Self-Powered Photodetector System, Adv. Mater., 29 (2017) 1703694,

https://doi.org/10.1002/adma.201703694

55. Z. Jin, L. Gao, Q. Zhou, J. Wang, High-performance flexible ultraviolet photoconductors

based on solution-processed ultrathin ZnO/Au nanoparticle composite films, Sci Rep, 4 (2014)

4268, https://doi.org/10.1038/srep04268

56. Y. Lu, Y. Liu and P. Chen, High efficient photovoltaics in BaTiO3 thin film, Mater. Res.

Express 6 (2019) 116217, https://doi.org/10.1088/2053-1591/ab39a6

57. H. Fan, C. Chen, Z. Fan, L. Zhang, Z. Tan, P. Li, Z. Huang, J. Yao, G. Tian, Q.Luo, Z.Li,

X. Song, D. Chen, M. Zeng, J. Gao, X. Lu, Y. Zhao, X. Gao, and J.M. Liu, Resistive switching

and photovoltaic effects in ferroelectric BaTiO3-based capacitors with Ti and Pt top electrodes,

Appl. Phys. Lett., 111 (2017) 252901, https://doi.org/10.1063/1.4999982


58. D. A. Neamen D 2012 Semiconductor Physics and Devices Basic Principles 4th Edition,

McGraw-Hill Education, New York 2011

59. S. Y. Wang, M. Li, W. F. Liu, J. Gao, Resistive switching behavior of

BaTiO3/La0.8Ca0.2MnO3heterostructures, Phys Lette A, 379 (2015) 1288-1292,

https://doi.org/10.1016/j.physleta.2015.02.037

60. F. Wang and A. M. Rappe, First-principles calculation of the bulk photovoltaic effect in

KNbO3 and (K,Ba)(Ni,Nb)O3−δ, Phys. Rev. B 91 (2015) 165124,

https://doi.org/10.1103/PhysRevB.91.165124

61. A. P´erez-Tom´as, A. Mingorance, D. Tanenbaum, M. Lira-Cantú, Metal Oxides in

Photovoltaics: All-Oxide, Ferroic, and Perovskite Solar Cells In The Future of Semiconductor

Oxides in Next-Generation Solar Cells, Barcelona, Spain, (2018) 267–356,

https://doi.org/10.1016/B978-0-12-811165- 9.00008-9

62. A. M. Glass, D. von der Linde, and T. J. Negran, Highvoltage bulk photovoltaic effect and

the photorefractive process in LiNbO3, Appl. Phys. Lett. 25 (1974) 233,

http://dx.doi.org/10.1063/1.1655453

63. Brody, P.S., Crowne, F. Mechanism for the high voltage photovoltaic effect in ceramic

ferroelectrics. J. Electron. Mater. 4 (1975) 955–971 https://doi.org/10.1007/BF02660182

64. X. Lv, X-X.Zhang, and J. Wu, Nano-Domains in Lead-Free Piezoceramics: A Review, J.

Mater. Chem. A, 8 (2020) 10026-10073, https://doi.org/10.1039/D0TA03201H

65. S. Pal, S. Muthukrishnan, B. Sadhukhan, Sarath N. V., D. Murali, and P. Murugavel, Bulk

photovoltaic effect in BaTiO3-based ferroelectric oxides: An experimental and theoretical

study, J. Appl. Phys. 129 (2021) 084106, https://doi.org/10.1063/5.0036488

66. H. Xiang, F. Zhang, Z. Yi, M. Ma, Y. Gu, F. Liu, Y. Li, and Z. Liu, Non-stoichiometry

induced switching behavior of ferroelectric photovoltaic effect in BaTiO3 ceramics, Phys.

Status Solidi RRL, 13 (2019)1900074, https://doi.org/10.1002/pssr.201900074


67. Wen-Yuan Pan, Yu-Cheng Tang, Y. Yin, Ai-Zhen Song, Jing-Ru Yu, S. Ye, Bo-Ping

Zhang , Jing-Feng Li, Ferroelectric and photovoltaic properties of (Ba, Ca)(Ti, Sn, Zr)O3

perovskite ceramics, Ceram Inter, 47 (2021) 23453-23462

https://doi.org/10.1016/j.ceramint.2021.05.061

68. R. Su, D. Zhang, M. Wu, Fa-tang Li, Y. Liu, Z. Wang, C. Xu, X. Lou, Q. Yu, Y. Yang,

Plasmonic-enhanced ferroelectric photovoltaic effect in 0-3 type BaTiO3-Au ceramics, J Alloys

Compd 785 (2019) 584-589, https://doi.org/10.1016/j.jallcom.2019.01.223


Figure captions

Fig.1. (a) ADXRD patterns of BT and BFNT samples at RT. (b) Enlarged (002), (200)

reflections from BT and BFNT samples. Rietveld refinement of pure BT refined with (c)

P4mm (d) P4mm + Amm2 space groups (e-g) Rietveld refinement ADXRD patterns of

0.025BFNT, 0.05BFNT and 0.075BFNT ceramics.

Fig.2. Enlarged view of two nearby linked TiO6 octahedra with octahedral tilting of (a) BT and

(b) 0.075BFNT samples.

Fig.3. (a) XPS binding energy spectra of Ti-peak shifting and high-resolution core-level

spectra of (b) Fe-2p, (c) Nb-3d and (d) O-1s elements.

Fig.4. Polarization- Electric field (P-E) hysteresis loop of BT and BFNT samples at RT.

Fig.5. Absorption spectra of BT and BFNT samples as a function of (a) wavelength, and (b)

Tauc’s plot. The total density of states (TDOS) and projected density of states (PDOS) for the

structures (c) tetragonal BT and (d) cubic 0.075BFNT. (In both graphs, the Fermi level is set

to zero.)

Fig.6. (a) J-V characteristics of BT and BFNT samples. (b) Time-dependent photocurrent

density (J-t) curve. (c) Time-resolved photocurrent density rise and decay curves of BT and

0.075BFNT device (d) Graph comparing Voc, Jsc and PV power of BFNT with other BT-based

ferroelectric photovoltaics.

Table captions:

1. Deconvolution of O-1s peak parameters for BT and BFNT samples obtained from XPS

measurement.
2. Bandgap (Eg), Bandwidth (EB), 𝜂, and probability of d-orbital mixing obtained for BT and

BNFT samples.

3. Photovoltaic parameters determined for BT and BFNT samples.

Table 1

Sample O(I) O(II) (%) O(III)(%)

Binding Peak Binding Peak Binding Peak

energy area energy area energy area

(eV) (%) (eV) (%) (eV) (%)

BT 528.17 28.27 533.59 67.52 537.83 4.21

0.025BFNT 527.48 26.59 531.41 70.15 534.06 3.24

0.075BFNT 527.64 33.79 530.43 55.73 532.17 10.46

Table 2

Sample Eg (eV) EB (eV) 𝜂 rd (%)

BT 3.25 3.26 0.49816 15

0.025BFNT 2.68 3.14 0.42814 18

0.05BFNT 2.64 3.12 0.42290 18.5

0.075BFNT 2.55 3.10 0.37958 19


Table 3

Sample Jsc Voc FF η R D

(μA/cm2) (Volts) (%) (mA/W) (× 108

Jones)

BT 0.69 3.40 0.290 0.0006 0.0069 1.688

0.025BFNT 0.22 5.14 0.297 0.0003 0.0022 0.538

0.05BFNT 0.97 7.68 0.307 0.0022 0.0097 2.373

0.075BFNT 1.46 8.31 0.332 0.0041 0.0146 3.572


Highlights

Highlights:

 Al/BaTi1-2xFexNbxO3/Ag is an excellent bulk ferroelectric photovoltaic device

configuration.

 Increment of aliovalent Fe3+ and Nb5+ ionic substitutions reduces the bandgap of

0.075BFNT ceramics to ~2.55 eV and broadens the visible spectrum absorption.

 The photocurrent density of Al/0.075BFNT/Ag photovoltaic cell is about ~2.2 times

higher than that of pure BT.

 A high photovoltaic power of ~12 μW/cm2 was achieved.


Unmarked Revised Manuscript for Publication Click here to view linked References

Impact of charge-compensated Fe and Nb co-substitution on BaTiO3: bandgap and grain

size reduction and enhanced bulk photovoltaic power of Al/BFNT/Ag solar cell

L.Venkidua, N. Rajaa&b and B.Sundarakannana*

a Department of Physics, Manonmaniam Sundaranar University, Abishekapatti, Tirunelveli-

627012, Tamil Nadu, India.

b The Institute of Mathematical Sciences, C.I.T. Campus, Taramani, Chennai 600113, India.

Corresponding author: sundarakannan@msuniv.ac.in

Abstract

The generation of above bandgap photovoltage using bulk ferroelectric materials has

become a subject of great interest, however, their photocurrent density is limited by a broad

bandgap and poor conductivity. To overcome this limitation, we replaced aliovalent metal ions

(Fe3+ and Nb5+) at the B-site of robust ferroelectric BaTiO3 and fabricated an Al/BaTi1-

2xFexNbxO3/Ag photovoltaic device. Both the experimental and the theoretical studies showed

that bandgap was lowered to ~2.55 eV and hence absorption of wide energy range of the solar

spectrum was attained. An apt top electrode, reduced bandgap and domain size resulted in

greater photocurrent density of 1.46 μA/cm2 and photovoltage of 8.31 volts for

Al/0.075BFNT/Ag solar cell in unpoled condition. This research suggest that reduced band

gap, mixed structural phases and nano-sized domains suffices greatest PV power output while

the large polarization and poling are not necessary prerequisites.

Keywords: Bandgap, Photocurrent, Open circuit voltage, Ferroelectric Photovoltaic

1 Introduction:

Photovoltaic solar energy is obtained by converting sunlight into electricity using a p-

n junction based on the semiconductors [1]. Photovoltaic (PV) technology is playing a


significant role in modern industry to alleviate the energy crisis. Traditional junction-based

solar cell technology is widely used, though the raw material of silicon (Si) is relatively high

price and the fabrication process requires so sophisticated technology. The output voltage in

semiconductor based material is limited by their bandgap and the power conversion efficiency

(PCE) depends on the complicated interfaces of PV devices. Conversely, non-centrosymmetric

bulk ferroelectric materials proficiently separate electron-hole pairs through their internal

electric field originating from the spontaneous polarization represented as the bulk photovoltaic

effect (BPV) [2, 3]. BPV devices are not limited by the bandgap and therefore produce a larger

photovoltage (anomalous photovoltaic effect) than traditional junction-based PV devices [2‒

4]. In recent times, bulk ferroelectric polycrystalline photovoltaics have attracted special

attention because of their low cost, good stability, facile fabrication of the device, and high

photovoltage [2‒5]. The main drawback of ferroelectric ceramics is their low power

conversation efficiency (PCE) mainly due to their low photocurrent density under light which

is related to their low conductivity [2, 3, 5]. Therefore, how to simultaneously enhance the

photocurrent and photovoltage of bulk ferroelectric ceramics and thereby power conversion

efficiency becomes an important issue to pushing forward the practically realizable application

of the ferroelectric photovoltaic device. Recent research focused on the BPV effect in various

ferroelectric systems such as ferroelectric ceramics [2-5], single crystals [6, 7] thin films [8, 9,

10], polymers [11], organic and inorganic ferroelectric composites [12]. Recent research has

proposed explanations for the BPV with various models such as domain wall theory [13],

ballistic model [2], shift-current model [14], interface effect (Schottky- junction effect) [15]

electrode, and depolarization field effect [16]. Considering the above models, ferroelectric

materials with relatively small bandgaps and large polarizations are promising candidates for

photovoltaic applications because the materials can absorb large amounts of the solar spectrum

while simultaneously separating charge carriers.


In this regard, BaTiO3 (BT) is considered one of the promising eco-friendly

ferroelectric perovskite materials with off-center distortions from its centrosymmetric position

and large spontaneous polarization arising from the strong hybridization among the lowest

unoccupied Ti4+-d0 state and O-2p state at room temperature [17, 2]. However, pristine

ferroelectric BT can only be able to absorb the ultraviolet (UV) region of the solar spectrum

due to its wide bandgap (3.2 eV) [14, 17, 18]. Only 3.5% of solar radiation is encompassed by

UV light (wavelength 200‒390 nm) yet visible light accounts for 40% of solar energy

(wavelength 390-700 nm) [5]. Therefore, constructing a ferroelectric BPV device with a

narrow bandgap (Eg) is an essential factor for absorbing abundant light and maintaining

ferroelectricity. Recent studies have focused on substituting ions at A-site and/or partially

occupied non-d0 ions at the B-site of BT, as a way to reduce its bandgap as well as to control

ferroelectric properties. In this regard, a large short-circuit current density of Jsc=7.51 μA/cm2

was obtained with A-site Sr-doped BT, but it didn’t induce anomalous photovoltage (Voc<1.5

V i.e. Eg = 3.04 eV) [19]. However, the maximum Voc of about 16 V was attained by BT co-

doping with Bi and Li ions at the A-site, but the determined photocurrent was within the range

of nano ampere [14]. Kola et al. achieved a large Voc of 16 V with a single substitution on B-

site as Ag/BaTi1-xSnxO3/Ag PV device based on the shift current model, but the current density

was very low around ~2 nA/cm2 range [2]. However, there have been only a few reports on the

effect of co-substitution on the B-site of bulk ferroelectric photovoltaic materials up to now.

For example, L.Wu et al. reported that Ni2+ and Nb5+ co-doped (0.9)BaTiO3‐

(0.1)Ba(Ni1/2Nb1/2)O2.75 provided a decreased bandgap of ~1.5 eV and maintained remnant

polarization ~5 μC/cm2 with obtained steady-state current of 8 nA/cm2 [20]. In contrast to other

transition metal pair substitutions (such as Mg3+-Nb-5+, Co3+-Nb5+, and Sc3+-Nb5+), the Fe3+-

Nb5+ pair at the B-site of BT has little effect on the bandgap reduction, and the significant
suppression in ferroelectricity is be expected to reduce the photovoltaic properties of BT and

decrease the device performance [21, 22]. However, there is a lack of experimental evidence.

Also, it has been noted that in conventional electrode/ferroelectric/electrode

architecture, the net build-in electric field (Ebi) of the top and bottom metal electrodes can form

two back-to-back Schottky barriers that deplete the electrode-ferroelectric interface [15, 23].

From the prior reports, lowering the height of the Schottky barrier with a low-work function

metal electrode can induce a large photovoltaic effect of the device [15, 23]. Therefore, by

choosing a suitable configuration of electrodes in contact with ferroelectric materials, the

photocurrents and photo-generated voltage of the device may also be enhanced. In this regard,

aluminium (Al) and silver (Ag) electrodes have been selected as the top and bottom electrodes

respectively, for our PV device. Considering that as a scope, we used a ferroelectric material

such as BaTiO3 to provide the bandgap reduction, enhance photocurrent, photovoltage and

maintaining ferroelectricity through charge-neutral chemical co-substitution of aliovalent Fe3+

and Nb5+ at B-site with an asymmetrical electrode architecture. Both Nb5+ and Fe3+ perform a

typical charge compensator and bandgap reducers respectively, with strong photo response,

high electron mobility, reduced recombination loss, easy crystallization, controlled oxygen

vacancies and good candidates in the ferroelectric photovoltaic system. In particular, the

reduction of the bandgap from 3.2 eV to 2.55 eV and enhanced photocurrent density of 1.46

µA/cm2 have been achieved with Al/0.075BFNT/Ag device, which is larger than that of the

Ag/BBLT/Ag [17] and Ag/BST/Ag photovoltaic cells [2]. We perform theoretical

computations to further envisage our experimental results and to find the reason for the

bandgap reduction.

2 Experimental procedure
The bulk ceramics powder with the composition Ba(Ti1-2xFexNbx)O3 (BFNT) (x=0.0

0.025, 0.05, 0.075) was fabricated by solid-state reaction method (SSR) and represented as BT,

0.025BFNT, 0.05BFNT and 0.075BFNT. The starting materials of BaCO3 (99.9%), TiO2

(99%), Fe2O3 (99.9%), and Nb2O5 (99.99%) were weighted according to the stoichiometry of

each composition. The mixed powder was ground using a mortar and pestle for 30 min. The

mixture was calcined at 950oC/5 hr in the air atmosphere for all compositions. The calcined

powder was mixed with a binder of PVA, (polyvinyl alcohol) by manual granulation in a

mortar. The uniaxially compressed pellets are 13 mm in diameter and 1 mm thick under 7-ton

weight. The pellets were sintered at 300oC/1hr to remove the PVA and then sintered at

1300oC/2hr in a furnace. Room temperature X-ray diffraction measurements were carried out

using λ= 0.9654 Å the angle-dispersive X-ray diffraction (ADXRD) at the beamline-12

installed in Indus-2 synchrotron source RRCAT Indore, India. The FT-Raman spectrometer

(Bruker, RFS 27) was used to probe the structural phases of the samples. The instrument is

equipped with a Nd:YAG laser (1064 nm) as a source and an InGaAs detector (1024 × 256). It

has a resolution of 2 cm-1. The measured data was fitted with Lorentz function using Origin

software (version:8.5) and fitting parameters such as peak position and line width were added

in supporting information. To determine the bandgap of the samples, a diffuse reflectance

spectrophotometer (DRS) (UV-2400 Schimadzu) was used. The silver paste was applied on

both sides of the polished pellets for measuring the room temperature polarization-electric field

(P-E loop) hysteresis loops using a Precision ferroelectric tester system (Radian technologies

P-E loop tracer). The PHI - VERSAPROBE III is an advanced instrument that was used to

obtain X-ray photoelectronic spectra. It features a multichannel detector and monochromatic

Al-Kα source with energy of 1486.6 eV, providing excellent resolution ranging from 0.1 to 40

eV for sample analysis and characterization. The peaks background were subtracted for more

accuracy in peak fitting, then the Gaussian function was used to fit each elemental XPS peaks
with Origin software (version: 8.5) and the parameters were included in supporting

information. The Energy Dispersive X-ray Spectrometer Quantax 200 with X Flash® 6130 was

used for EDX and mapping studies. The photovoltaic measurements were done using Sol3A

Class AAA solar simulator with Keithley electrometer (2400) measuring setup.

2.1 Device fabrication

The PV studies were conducted on the cell configuration of Al/BFNT/Ag, where Al

and Ag are the top and bottom electrodes, respectively. BFNT pellets were mirror polished and

thickness of 0.5mm, diameter of 11mm and area of pellets of ~0.95 cm2. For photovoltaic

measurements, Aluminium (Al) electrode was coated on the top surface of the sample (0.1-mm

thickness of shadow mask) and bottom surface was covered with silver (Ag) electrode using a

thermal evaporation technique with heating the samples at 400oC/30min. Ag paste was used to

connect the electrodes and copper wire for the photovoltaic experiment.

2.2 Computational Details

Density functional theory (DFT) simulations were performed to understand the

substantial decrease in the measured bandgap value of BFNT samples. Calculations were

carried out on a 3×3×3 supercell of 135 atoms using the projector augmented wave (PAW)

[24] method as implemented in the Vienna Ab Initio Simulation Package (VASP) [25]. The

generalized gradient approximation (GGA) [26] was used to depict the exchange and

correlations of electrons. A Monkhorst pack scheme [27] k-points mesh was set to 2×2×2, and

a plane wave energy cut-off of 520 eV was chosen. To account for any possible strong

correlation GGA +U approach was employed, with U = 4 eV for the 3d-orbitals of the Fe-ions

[28]. Since GGA functionals tend to underestimate the bandgap values for most materials, HSE

hybrid functional is used on optimized structures to compute the density of states (DOS) [22,

32]. DOS was sampled with 3×3×3 k-mesh and Gaussian smearing of 0.1 eV.
3 Results and discussion:

3.1 Synchrotron Angle Dispersive X-ray diffraction (ADXRD)

The synchrotron X-ray diffraction patterns of BT, 0.025BFNT, 0.05BFNT and

0.075BFNT samples were measured at room temperature and are shown in fig. 1(a). No

impurity phase was observed in all samples. Peak indexed for ceramics samples agrees with

recently reported BFNT [30, 31], which has a perovskite cubic phase with space group Pm3̅m

because of an increase in co-substitution. The gradual decrement in the (200)pc peak splitting

is evidence as seen in the enlarged view in fig 1(b). The two-phase model was confirmed for

BT samples from synchrotron ADXRD is consistent with laboratory X-ray diffraction

described in our prior study [31, 32], the two-phases of the BT sample are tetragonal (P4mm)

and orthorhombic (Amm2).

The Rietveld analysis program Jana2006 package [33] was used for crystal structure

refinement. Rietveld refinement parameters for specified space groups are the lattice

parameters, atomic coordinates, atomic occupancies, scale factor, 2θ zero shift, 36-degree

Legendre polynomials for background and pseudo-Voigt line shape with asymmetry. As per

the nominal composition of the perovskite ABO3 structure, the A site atomic occupancy is Ba,

and the B site atomic occupancy is Ti/(Fe & Nb). Figure 1 (c) and 1 (d) show the Rietveld

refinement fit of the synchrotron ADXRD pattern of pure BT using a single-phase model,

tetragonal (P4mm) and two-phase model tetragonal (P4mm) + orthorhombic (Amm2)

respectively. The goodness of fit (χ2) parameters for the P4mm and P4mm+Amm2 models are

χ2=5.71 and 4.71, respectively. The poor fit or misfits (Bragg position, peak shape, intensities)

could not be improved satisfactorily, as shown in the insets of fig. 1(c) [indicated by circle]. In

contrast, P4mm and Amm2 two-phase models could well account for the Bragg profiles, peak

position, and intensities. Fig. 1(d) illustrates the excellent fit in an enlarged view of (111)pc and
(200)pc for the two-phase model. Synchrotron ADXRD patterns of co-substituted BT, depicted

in fig. 1(e)-(g), were refined with P4mm (T) and Pm3̅m (C) structural models which conform

with earlier reports [30, 31]. Rietveld refined structural parameters such as lattice parameter,

volume, c/a, atomic position, R-factor (wRp), phase fraction (%wt), and goodness of fit (χ2) of

BFNT (x=0.0, 0.025, 0.05, and 0.75) samples are listed in table S1. The ratios of phase

percentage of P4mm and Pm3̅m crystal structure are 85.3:14.7, 76.7:23.3 and 41.5:58.5 for x=

0.025, 0.05, and 0.075 respectively. The mixed phase structures of BT and BFNT samples were

then verified by Raman spectroscopy as shown in Fig. S1(a–d), and a detailed description is

given in the Supporting Information.

According to Vegard’s law [34], the lattice constant of the cubic unit cell increases with

the concentration of substitution ions, as shown in table S1. Also, the volume of the cubic

lattice increases with x due to larger ionic radii (rFe3+=64.5 pm and rNb5+= 64 pm) in weight

compared to the host radius Ti (rTi4+= 60.5 pm). The c/a ratio of the tetragonal unit cell

constantly drops with increasing x, which can be attributed to the aliovalent replacement ions

such as Fe3+ and Nb5+ at the Ti4+ site that induces an internal stress field in BT ceramics. The

structural phase transition originated from this internal stress field by the broken Ti-O

hybridization. The change in lattice parameters and the volume of the unit cell is associated

with oxygen-octahedral tilting and distortion of octahedra. The change in lattice parameters

and the volume of the unit cell is associated with the oxygen-octahedral tilting and distortion

of octahedra. The oxygen-octahedral tilt can be determined by the equation given below [35,

36]:

180−𝜃
𝜃𝑡𝑖𝑙𝑡 = (1)
2

where 𝜃 is the angle between two Ti4+ ions in adjacent unit cells bonded with a face-

centered oxygen ion (Ti4+-O2--Ti4+). The octahedral tilting of BT compared with


0.075BFNT is illustrated in fig.2(a) and (b). Also, the octahedral distortion can be estimated

by the following relation [35]:


6
1 𝛿𝑛 − < 𝛿 > 2
∆𝑜𝑐𝑡 = ∑ [ ]
6 <𝛿>
𝑖=1

(2)

where 𝛿𝑛 and < 𝛿 > denote the individual and average interatomic metal-oxygen bond length.

With increasing substitution at the B-site, the structure of BFNT ceramics is closer to pseudo-

cubic with reducing octahedral tilting and distortion from table S2. The decreasing octahedra

tilt could affect the bandgap of the samples.

3.2 X-ray photoelectron spectroscopy

To investigate the electronic states of each element in BaTi1-2xFexNbxO3, the X-ray

photoelectronic spectra were recorded. The detailed survey and core-level spectra of Ba and Ti

for BT, 0.025BFNT and 0.075BFNT are discussed in the supporting information (fig. S2). Fig.

3(a) denotes that the binding energy of the Ti-2p spectrum shifted into a lower energy level

with increasing substitution. This can be attributed to the local electric field. This local electric

field induced to decrease in the bandgap of the ferroelectric material [37]. As a result, the

lowering bandgap influenced the intermediate charge transfer, thus can suppress the resistivity

of the material.

Fig. 3(b) illustrates the asymmetric nature of core-level Fe-3d spectra of 0.025BFNT

and 0.075BFNT and it can be split into Fe-2p3/2 and Fe2p1/2 spin states. For 0.025BFNT, the

XPS spectra can be deconvoluted into the characteristic peaks of Fe2+ and Fe3+ assigned to ~

708.8 eV, ~713.7 eV and ~ 710.8 eV, 722.06 eV, respectively. The absence of Fe2+ peaks in

0.075BFNT indicates the suppression of the oxygen vacancies. Additionally, the satellite peaks

around ~718.12 eV and 720.1 eV were also found in fig 3(b). These results are matched with
previous literature [37, 38]. The absence of doublet peaks around 719.9 eV and 706.7 eV

indicates that the substituted iron (Fe) is ionized and not metallic [39]. The photoelectronic

spectra around ~204.3 eV and ~208.3 eV represent the Nb-3d5/2 and Nb-3d3/2 states

respectively, were seen in fig. 3(c) for 0.025BFNT and 0.075BFNT samples. The oxidized state

of Nb is 5+ for these peaks.

The core-level spectra of O-1s of BT, 0.025BFNT and 0.075BFNT are shown in fig.

3(d). The O(I), O(II), and O(III) spectra represent the lattice oxygen bonding with Ti/Fe/Nb,

oxygen vacancies, and adsorbed oxygen is assigned at ~529 eV, ~531 eV, and ~533 eV

respectively. The binding energy and area under the fitting curve of O(I), O(II), and O(III) are

listed in table 1, where the area of the oxygen vacancy tends to decrease with increasing

substitution. This results from the simultaneous doping of donor (Nb5+) and acceptor (Fe3+) at

the B-site can neutralize the charge carriers and avoid the oxygen vacancies [40]. The reaction

can be written using Kroger-Vink notation as follows;

′ •
𝐹𝑒2 𝑂3 + 𝑁𝑏2 𝑂3 → 2𝐹𝑒𝑇𝑖 + 2𝑁𝑏𝑇𝑖 + 6𝑂𝑜 (3)

Hence, the samples are obeying the following condition of charge neutrality [41]

∑𝑛𝑖=1 𝑥𝐴𝑖 𝑛𝐴𝑖 + ∑𝑛𝑘=1 𝑥𝐵𝑖 𝑛𝐵𝑖 = 6 (4)

where, 𝑥𝐴𝑖 , 𝑥𝐵𝑖 (for x= 0.0, 0.025, 0.05 and 0.075) are the fractions of ions and 𝑛𝐴𝑖 , 𝑛𝐵𝑖 are the

valances of ions (Ba-2+, Ti-4+, Fe-3+, Nb-5+) at the A & B sites, respectively.

3.3 Hysteresis loop

To find the impact on ferroelectricity under B-site substitution of BT ceramics, the

Polarization-Electric field (P-E) hysteresis loop was measured at room temperature with a 30

kV/cm applied field, as shown in fig. 4. The shape of the P-E hysteresis loop tends to be

unsaturated with increasing metal substitution. The parameters such as remnant polarization
(Pr), spontaneous polarization (Ps), and coercive field (Ec) were decreased with increasing x as

in fig.4. It was suggested in the literature that the reduction of ferroelectricity upon metal

doping due to the formation of oxygen vacancies leads to an increase in the leakage current in

ferroelectric samples [32]. However, in the current study, the production of oxygen vacancies

is strongly restricted by the charge compensator Nb5+ which is further validated by the XPS

results of the Fe-2p and O-1s spectra. Therefore, the ferroelectricity in BT ceramics was

suppressed by the following features. The origin of local polarization in BaTiO3 ceramics can

be attributed to the off-center displacement of d0 Ti-ion about the oxygen octahedral cage. The

inclusion of bigger-size ions helps reduce the octahedral distortion and leads to suppressing the

Ti-O off-center displacement of BT ceramics. According to modern ferroelectric theories, the

partially occupied transition metal d-ion replaced at Ti-site supports centrosymmetric and

therefore simultaneously reduces the off-center distortion needed for ferroelectricity [42, 43].

The following theoretical calculations demonstrate the proof that the non-d0 Fe3+ state perturbs

both the valence and conduction band of BFNT materials. Therefore, the ferroelectricity of BT

was destabilized by the alterations in the hybridization of Ti-3d to O-2p orbital. These results

are consistent with the structural discussions and indicate that the ferroelectricity was

suppressed by co-substitution.

3.4 Bandgap

The DRS was used to investigate the optical characteristics of BT and metal-substituted

BT in the wavelength region of 200-900 nm as shown in fig. 5 (a). Fig 5 (a) illustrates the

absorbance spectra of the samples; in which all the substituted samples exhibit a shoulder

around ~400-500 nm (encircled region) and its absorbance increases along with x and it

broadens toward lower energy ranges. The bandgap energy of BT and BFNT samples can be

estimated using the Tauc plot as shown in fig. 5(b) and listed in table 2. The optical bandgap

of pure BT (3.25 eV) decreases with increasing metal substitution at the B-site. According to
previous studies, the valence band (VB) is assigned to the O-2p orbital and that of the

conduction band (CB) to the Ti-3d0 orbital in pure BaTiO3 [44], which generates a bandgap

energy of 3.25 eV. Meanwhile, non-d0 metals like Fe3+ replaced on the B-site might affect the

electronic band structure due to the anisotropic crystal fields of the octahedron created by the

metal-oxygen interactions [45]. In this context, the Fe/Ti/Nb could be treated as a point charge,

and their spatial arrangement affects the energy of the d-orbital of the central metal atom.

According to crystal field theory, a partially occupied 3d orbital in an octahedral crystal field

is divided into two eg orbitals and three t2g orbitals [2]. This field causes eg to evolve into one

high energy state (dx2-dy2) and one low energy state (dz2), whereas t2g evolves into one high

energy state (dxy) and two low energy levels (dzy and dzx) [45, 46]. As a result, the existence of

the 3d state between CB and VB with Fe3+ ions is indicated by the inter-band transition from

dz2 to dx2-dy2. Previous reports suggested that the large electronegativity difference between

metal and oxygen can induce a larger bandgap in perovskite materials [46, 47]. Here, the

electronegativity difference of Nb-O (1.84) and Fe-O (1.61) is much less than that of Ti-O

(1.9), therefore the reduced bandgap of BFNT largely depends on Fe3+ metal ion.

The d-orbital metal ions and the p-orbital of oxygen ions can be defined using the linear

combination of atomic orbitals model (LCAO). If the density of the state is deemed constant,

the following equations show that the probability of an average d-orbital mixing (rd) in the

valence band can be calculated based on the bandgap (Eg) and bandwidth (EB) [48, 40];

1 𝜂
𝑟𝑑 = 2 + 2(1−𝜂) ln(𝜂) × 100 (5)

where 𝜂 means the covalent mixing and it lies between 0 and 1. The value of 𝜂 can be

determined by the following equation:

𝐸𝑔
𝜂 = 2𝐸 (6)
𝐵
From table 2 𝜂 values tend to be small with increasing metal substitution indicating the strong

covalent mixing.

The value of EB is found from the equation below:

𝐸𝑔 2
𝐸𝐵 = √( 2 ) + 8(𝑝𝑑𝜋)2 (7)

For perovskite samples, the value of 𝑝𝑑𝜋 =1 eV [45] and the calculated𝐸𝐵 , 𝜂 and 𝑟𝑑 are listed

in table 2. The results obtained here are evidence that the mixing of d-orbitals into valence

bands is a function of bandgap (𝐸𝑔 ). The decreasing bandgap suggests an increased probability

of d-orbitals mixing into the valence band, which is verified by the DFT calculations as shown

below.

3.5 Density functional theory

A 3×3×3 supercell was constructed using the experimental lattice parameters of cubic

BFNT. Four Ti ions in the supercell were replaced by two Fe and two Nb ions to match the

experimental doping concentration (x=0.075). Based on the dopant-dopant distance, eight

distinct configurations were identified as illustrated in fig.S3 and the ground state energy

calculations were performed to understand the effect of the dopant ion interaction over the

ground state energy. It is observed that the calculated ground state energies differ relatively, as

tabulated in table S5 and the optimized structure of the minimum energy configuration-2 is

used to calculate DOS.

Figure 5 (c, d) shows the computed electronic density of states corresponding to the

tetragonal-BT and cubic-0.075BFNT structures in their minimal energy configuration. The

calculated bandgap values for pure BT and 0.075BFNT were 3.26 and 2.49 eV, respectively.

These values are in line with the experimental results of 3.2 and 2.55 eV. The valence band of

pure BT is mostly occupied by the O-2p state, whereas the conduction band consists primarily
of the Ti-3d state, as seen in Fig. 5(c). The estimated valence bandwidth is -4.75 eV, which is

consistent with the DFT results from previous papers [49, 50].

It is remarkable to notice from fig. 5(d) that on metal substitution at B-site, the Ti-3d

and O-2p orbital still dominate the CB minimum and VB maximum of the BFNT respectively.

With increasing substitution level to x=0.075, the formation of new 3d states of Fe and Nb was

observed at both the valence and conduction bands as shown in fig. 5(d). The PDOS in Fig.

5(d) shows the hybridization of Fe ion 3d-states with BT host electronic states, close to the

valence band maximum. The incorporation of Fe-3d states localized in VB has resulted in

stronger hybridization with the O-2p state near the valance band edge than that observed for

Ti–O hybridization. HSE calculations clearly show the formation of new states at the top of the

valence band edge with predominant Fe 3d character, which primarily decreases bandgap.

Additionally, strong d–d interaction between Ti–3d and Fe/Nb – d state at top of VB has led to

the accumulation of carriers near bottom CB as well as top VB edges. Also, the contribution

of lower Fe-3d (~ -7.2 to -6.7 eV) state increases the width of VB from ~ -4.75 eV (i.e. BT) to

~ -8 eV as shown in fig. S4. Such enhancement is beneficial for the mobility of the

photogenerated charge carriers within the material system as from an earlier report [51]. On

the other hand, fig. 5(d), the substitution of Fe and Nb in BFNT results in the hybridization of

Nb–O and Fe-O at the conduction band, which effectively separates the photogenerated

electron-hole pair [52]. Furthermore, this hybridization between Fe-3d and Nb-3d with O-2p

pushes the CB minimum towards lower energy levels resulting in the significant reduction of

the conduction band edge from 3.26 to 2.49 eV, as shown in Fig. 5(d) in the reason for

minimizing the photoexcitation energy of charge carriers. The red shift of CBM combined with

an increase in mobility of photogenerated charge carriers along VB due to Fe–O & Nb–O in

BFNT leads to the expectation of improvement in photocurrent density.


3.6 Photovoltaic characteristics:

To evaluate the photovoltaic (PV) response of BT ceramics with Fe and Nb substitution,

the short-circuit current density (Jsc) vs open-circuit voltage (Voc) curve were measured under

both dark and light irradiation of 100 mW/cm2 and the results are illustrated in fig. 6(a).

Surprisingly, all the BT samples exhibit bulk photovoltaic characteristics with J sc in

microampere and large Voc as listed in table 3. It was noted that the open-circuit voltage and

short-circuit current density increased under increasing the substitution and achieved 8.31 V

and 1.46 µA/cm2 respectively, for 0.075BFNT. The large Voc (i.eVoc>Eg) justifies that it is an

anomalous photovoltaic effect. Further, the following relation can be used to calculate the

photoresponsivity (R) and specific detectivity (D) of the device [53]:

𝐽𝑠𝑐
𝑅= (8)
𝐼

𝑅
𝐷 = [(2𝑞𝐽 1/2 ] (9)
𝑑𝑎𝑟𝑘 )

where I means the light intensity (100 mW/cm2), q refers the electron charge (1.6 × 10-19 C)

and Jsc, Jdark stands for the short-circuit current density and dark current density of the device,

respectively. According to table 3 the maximum R and D values of 1.46× 10-2 mA/W and 3.57

× 108 Jones were obtained for Al/0.075BFNT/Ag device, respectively, which is remarkably

larger than prior reports [53, 54].

The time dependent photocurrent density curve for all samples was measured under 100

mW/cm2 light irradiation with an ON/OFF switching time of 60 s and a bias of 10 V, as shown

in Fig. 6(b). It was found that the photocurrent density of all ceramic samples increased when

the light was turned on, then decayed and returned to its original state once the light source was

turned off. This clearly shows that all of the devices have a good ON/OFF switching response.

Fig. 6(c) displayed the performance of time resolved photocurrent density of BT and
0.075BFNT device. The photocurrent of both devices increased quickly when the light source

was turned on, followed by a modest improvement in Jsc, until achieving saturation. After the

light source is switched off, the photocurrent was rapidly falls for both the devices, then slowly

decays until returning to the initial condition of Jsc (dark current). Such rise and decay responses

of both devices were detected using exponential fitting of the experimental data using the

relation below;

−𝑡 −𝑡
𝐼 = 𝐼0 + [𝐶 𝑒𝑥𝑝 ( 𝜏 )] + [𝐷 𝑒𝑥𝑝 ( 𝜏 )] (10)
1 2

where I0 represents the steady state photocurrent, C and D are constants, t is time and 𝜏1 and 𝜏2

are the relaxation time constants [55]. The excellent fitted photo response curves of BT and

0.075BFNT samples as presented in fig. 6 (c) where 𝜏𝑟 and 𝜏𝑑 are time constants for rising and

decay edges, respectively. The estimated time constants of BT are: 𝜏𝑟1 = 7.5 s, 𝜏𝑟2 = 17.3 s,

𝜏𝑑1 = 10.1 s, 𝜏𝑑2 = 12.3 s and 0.075BFNT are: 𝜏𝑟1 = 7.0 s, 𝜏𝑟2 = 8.6 s, 𝜏𝑑1 = 10.6 s, 𝜏𝑑2 = 13

s. The rising constants are associated with the carrier creation process, whereas the decay

constants are associated with the carrier recombination process. As a result, BT has a longer

duration for carrier generation than 0.075BFNT, yet the decay constants of both devices are

comparable.

The PV properties of bulk BT ceramics depend on various factors such as material

thickness [56], the work function of the electrode material [23], light intensity [14], applied

poling field [17], electrical conduction and behavior of domain walls [13], ferroelectric

polarization [17] and electrode/ferroelectric/electrode interface [15, 23]. However, the bulk

photovoltaic effect is largely dependent on the interface of the top electrode/ferroelectric

material/bottom electrode architecture. According to Fan et al. the comparative study of the

bulk photovoltaic effect on Ti/BTO/SRO and Pt/BTO/SRO combinations revealed a short-

circuit current(Isc) of about 1.2 nA and 0.3 nA, respectively [57]. This result suggested that the
suppression of photocurrent in Pt/BTO/SRO interface was due to the large work function of

the Pt electrode (φPt = 5.5 eV) than Ti electrode (φTi = 4.33 eV). A similar PV study was

conducted for Mg/PLZT/ITO, Pt/PLZT/ITO, and Ag/PLZT/ITO cell configurations, where the

open-circuit voltage and short circuit current density increased in the order of large work

function metal electrodes to less work function metal electrodes [19]. Moreover, Subhajit Pal

et al. reported that the pure BT ceramics with a cell configuration of Ag/BT/Ag provide a Voc

of 2 V and Jsc of 3 nA/cm2 under the illumination of 160 mW/ cm2 with the device thickness

of about 0.5 mm [14]. Although in our case, Al/BT/Ag provides Jsc of 0.65 μA/cm2 and Voc of

3.4 V even under 100 mW/cm2 illumination with a similar thickness as pure BT. Here, the

height of the Schottky barrier (φB) can be determined by the difference between the work

function of the metal electrode (φM) and the electron affinity (χ) of FE material [58].

Considering that the work functions of Al and Ag are 4.125 eV and 4.48 eV respectively and

the electron affinity of BT is 3.8 eV [59], then φB= 0.325 eV (if Al is a top electrode) and φB=

0.68 eV (if Ag is a top electrode). Comparatively, the barrier height is much smaller when Al

works as the top electrode of the device than Ag. In addition, the enhancement of the PV

response in pure BT can originate from the coexistence of the tetragonal and orthorhombic

phases. According to the shift current theory, the enhancement of PV properties in KNbO3 is

achieved in excess in mixed of orthorhombic and rhombohedral phases compared to that of the

tetragonal phase [60]. Moreover, Kola et al. reported that maximum Voc was attained in B-site

doped BT at a larger orthorhombic phase than that of the tetragonal single phase [2]. Therefore,

the greater bulk PV response in pure BT in this study than in Subhajit Pal et al. [14] result is

based on the structural symmetry and low work function metal electrode.

The greatest variation in PV outcomes from metal substitution depends on the bandgap,

grain size, domain size, and dielectric and ferroelectric characteristics [13]. The existence of

an aliovalent Fe3+-Nb5+ pair in BaTiO3 greatly alters the properties of both the valence and
conduction bands, according to theoretical DFT calculation. The bandgap of the ferroelectric

material decreased considerably from 3.25 eV to 2.55 eV when B-site replacement increased.

This reduction in bandgap potentially absorbs the maximum visible region of the light source

as shows in fig. 5(a). Bandgap reduction is an important component covering a wide area of

the light source and is beneficial to increase the current density of the device. Additionally, the

XPS findings of fig. 3(b) demonstrated that the formation of minor oxygen vacancies was

obtained in 0.025BFNT, however in the 0.075BFNT sample, the oxygen vacancies were

substantially controlled by the lack of Fe2+ state. The oxygen vacancy suppression in

0.075BFNT increases the Jsc of the device more sharply than that in 0.025BFNT by reducing

the recombination between oxygen vacancies and hopping electrons.

Meanwhile, the open-circuit voltage of the device can be expressed as [61]

𝑉𝑜𝑐 = 𝐸𝑝𝑣 . 𝑑 (11)

where Epv denotes the photovoltaic field and d is the distance between the metal electrodes.

This equation suggests an anomalous photovoltaic effect, in which the open-circuit voltage of

the device is substantially higher than the bandgap of the material and is not constrained by the

bandgap of bulk ferroelectrics. For instance, A. M. Glass et al. in Fe-doped LiNbO3 [62]

attained the anomalous photovoltage of around 1,000 volts. Generally, the PV response of bulk

ferroelectrics depends on the grain size and domain size effect. In an earlier study, we described

the microstructure of Fe and Nb-modified BaTiO3. Currently, the microstructure of

0.075BFNT was analyzed and the results are shown in Fig. S5. Additionally, elemental analysis

of all samples was conducted using EDX mapping & spectra as depicted in Fig. S6; further

details regarding microstructure analysis can be found in Supporting Information. According

to Frank et al. it has been proposed that the generated high photovoltage in ceramics is

proportional to the number of grains, i.e., high photovoltage is generated from the summation
of individual photo-emf across the grains [63]. Consistent with this result, the number of grains

in smaller grain sizes is greater than in larger ones, resulting in higher output photovoltage with

increasing metal substitution. In addition, the grain size of BaTiO3 is co-related with the

domain size by the simple relationship as follows [64]

𝑑 ∝ 𝑔1/2 (12)

where g and d denote the grain and domain size of the ceramics, respectively. The domain size

will also be affected by changes in grain size. Earlier reports stated that the open-circuit voltage

of ferroelectrics was induced by the addition of an electrostatic potential difference between

the domain walls [13]. Here the grain size of BaTiO3 ceramics was decreased upon metal

substitution in the order of 3.6 μm, 0.68 μm and 0.51 μm for x=0.0, 0.025, and 0.05

respectively, from our previous findings [31] and fig. S5 reveals the grain size of x=0.075 was

0.14 μm. Thus, the smaller grain size results in smaller domain size and the smaller size of

domains have a larger number of domain walls and achieve a higher open-circuit voltage of

the device. As a result, the reduced grain size increased the photovoltage of the Al/BFNT/Ag

device.

The aforementioned findings demonstrated an inventive strategy to enhance both the

photocurrent and photovoltage of BFNT ceramics. Numerous investigations have

indicated that the large ferroelectric polarization is a crucial component for the anomalous

photovoltaic effect, in which the photovoltage of the ferroelectric material increases

substantially above the bandgap of the material meanwhile the generated photocurrent values

are very low in the range of nano to pico ampere [17, 24, 25]. Even for the pure BT sample,

the anomalous photovoltage of around 3.4 V was attained due to the larger polarization and

non-centrosymmetry orthorhombic + tetragonal phase structure, as confirmed by P-E loop and

ADXRD analysis respectively. However, the large anti-bonding property between Ti-3d and
O-2p results in a wide bandgap (Eg=3.25 eV), hence the generated photocurrent was quite low

(Jsc= 0.65 μA/cm2). According to the first-principle calculations based on the shift current

mechanism, maximum polarization is not a necessary condition for the BPV response, although

inversion symmetry must be broken [14, 65]. According to the above mechanism, the non-

centrosymmetric ferroelectric tetragonal phase coexists with the paraelectric cubic phase in

metal-substituted BT, as proven by ADXRD and Raman investigations, where the remnant

polarization was reduced compared to pure BT, but the photovoltaic parameters and PV power

(Voc × Jsc) of the 0.075BFNT device were remarkably improved and comparable with recent

studies, as shown in fig. 6(d) and table S7. Therefore, we concluded from the overall

experimental results and theoretical DFT calculations that the high photoresponsivity, large

detectivity of solar spectrum with concurrent enhancement of both photocurrent and

photovoltage in bulk ferroelectric BT is possible through appropriate co-substitution of

aliovalent Fe3+-Nb5+ pair, which is successfully tuning the bandgap to absorb abundant regions

of the solar spectrum, controlling oxygen vacancies to reduce the e-h recombination rate,

minimizing the grain size and using a low work function top electrode.

4. Conclusion:

In summary, the structural, optical, electrical and photovoltaic properties of BaTi1-

2xFexNbxO3 (x=0.0, 0.025, 0.05 and 0.075) were discussed. The synchrotron ADXRD data

combined with Raman spectra confirmed that the pure BT possesses the mixed phase of

tetragonal + orthorhombic at RT. Increasing the Fe3+ and Nb5+ substitution at B-site modifies

the structure to tetragonal + cubic symmetry, reducing octahedral tilt, and distortion along with

increasing internal stress field contributed to tuning the bandgap into a lower energy state of

2.55 eV and thus can efficiently absorb the larger area of the solar spectrum (~ 500 nm). To

comprehend the drastic decrement in the bandgap of BFNT samples from pristine BT, DFT

studies were carried out. It was found that the band gap value decreases drastically from 3.28
eV to 2.49 eV because of the aliovalent Fe3+-Nb5+ ions contributions to the valance and

conduction band. The tuned bandgap develops the metal-oxygen covalent character and

enhances the charge carrier mobility of the device. Further, XPS analysis depicted a clear

picture about suppression of oxygen vacancies in 0.075BFNT through Fe-2p and O-1s spectra.

Meanwhile, the grain size reduction is responsible for the continuous growth of Voc. From time

resolved J-t fitted curves, the slow response rise time constant was improved for 0.075BFNT

than BT device. Here, the simultaneous enhancement of Jsc and Voc does not follow the

maximum polarization concept. Finally, an apt top electrode, reduced bandgap, small grain size

and suppression of oxygen vacancies are contributed to enhance the PV power of

Al/0.075BFNT/Ag device to ~12 μW/cm2.

Acknowledgment:

This study was financially funded by the University Grant Commission-Basic Science

Research Fellowship scheme (F.No:25-1/2014-15(BSR)/7-305/2010/(BSR), and L. Venkidu is

thankful to the government of India for funding this study. The author also thanks Dr. A.K

Sinhan, Dr. Archana Sagdeo, and Mr. Mandvendra Narayanan Singh, Raja Ramanna Centre

for Advanced Technology, Indore, India, for their help in Angle Dispersive X-ray Diffraction

(ADXRD) beamline-12 (BL-12) at Indus-2 synchrotron radian source. The fabrication of the

mask (Al/BFNT/Ag) was carried out at the Center for Nano Science and Engineering (CeNSE),

Indian Insitute of Science, Bangaluru, under the INUP program funded by the Ministry of

Education (MoE), Ministry of Electronics and Information Technology (MeitY), and

Nanomission, Department of Science and Technology (DST), Government of India. The author

L.Venkidu sincerely thanks Ms. Adithi for their help during the mask fabrication. SERB grant

(CRG/2019/006990) is also acknowledged for the P-E loop tester. N. Raja gratefully

acknowledges C. Kamal, Raja Ramanna Centre for Advanced Technology, Indore, India for
the fruitful discussions and help in calculations. N. Raja also acknowledges Rajesh Ravindran

and The Institute of Mathematical Sciences, Chennai, India.

Author Contributions

The corresponding author, Dr. B. Sundarakannan is responsible for ensuring that the

descriptions are accurate and agreed by all authors. The role(s) of all authors is listed as follows,

L.Venkidu: Conceptualization, Methodology, Software, Writing - Original Draft

N. Raja: DFT Validation and Investigation

B. Sundarakannan: Conceptualization, Methodology, Validation, Investigation, Writing -

Review & Editing

Data availability

The datasets generated and analyzed during the current study are available from the

corresponding author upon reasonable request.

References:

1. ArturoMorales-Acevedo, Variable band-gap semiconductors as the basis of new solar cells,

solar energy, 83 2009 1466-14771, https://doi.org/10.1016/j.solener.2009.04.004

2. L.Kola, D.Murali, S.Pal, B.R.K.Nanda, and P.Murugavel, Enhanced bulk photovoltaic

response in Sn doped BaTiO3 through composition dependent structural transformation, Appl.

Phys. Lett. 114 (2019) 183901, https://doi.org/10.1063/1.5088635

3. V.M. Fridkin, Bulk Photovoltaic Effect in Noncentrosymmetric Crystals, Crystallogr.

Rep. 46 (2001) 654–658, https://doi.org/10.1134/1.1387133


4. A. B. Swain, D. Murali, B.R.K. Nanda, and P. Murugavel, Large bulk photovoltaic response

by symmetry-breaking structural transformation in ferroelectric [Ba(Zr0.2Ti0.8)O]0.5[(BaCa-

)TiO3]0.5 , Phys. Rev. Applied 11 (2019) 044007,

https://doi.org/10.1103/PhysRevApplied.11.044007

5. X.He, C.Chen, C.Li, H.Zeng, and Z.Yi, Ferroelectric, Photoelectric and photovoltaic

performance of silver niobate ceramics, Adv. Funct. Mater. 29 (2019) 1900918,

https://doi.org/10.1002/adfm.201900918

6. Y. Noguchi and H. Matsuo, Ferroelectric photovoltaic tensor in visible-light-active Fe-doped

BaTiO3 single crystals, jpn. J. Appl. Phys. 60 (2021) SFFA01,https://doi.org/10.35848

/1347-4065/ac0c6c

7. Y. Liu, X. Wang, F. Fan, and C. Li, Bulk Photovoltage Effect in Ferroelectric BaTiO3, J.

Phys. Chem. Lett. (2022), 13, 48, 11071-11075, https://doi.org/10.1021/acs.jpclett.2c03194

8. A. Zenkevich, Yu. Matveyev, K. Maksimova, R. Gaynutdinov, A. Tolstikhina, and V.

Fridkin, Giant bulk photovoltaic effect in thin ferroelectric BaTiO3 films, Phys. Rev. B, 90

(2014) 161409, https://doi.org/10.1103/PhysRevB.90.161409

9. V. M. Fridkin, Ferroelectricity and Giant Bulk Photovoltaic Effect in BaTiO3 Films at the

Nanoscale, Ferroelectrics, 484:1, 1-13, https://doi.org/10.1080/00150193.2015.1059151

10. Z.Fan, K.Yao,and J.Wang, Photovoltaic effect in an indium-tin-oxide/ZnO/BiFeO3/Pt

heterostructure, Appl. Phys. Lett. 105 (2014) 162903, https://doi.org/10.1063/1.4899146

11. Q. Liu, I.Khatri, R.Ishikawa, A.Fujimori, K.Ueno, K.Manabe, H.Nishino,and H.Shirai,

Improved photovoltaic performance of crystalline-Si/organic Schottky junction solar cells

using ferroelectric polymers, Appl. Phys. Lett. 103 (2013) 163503,

https://doi.org/10.1063/1.4826323
12. Z.Fan, K.Sun and J.Wang, Perovskites for photovoltaics: a combined review of organic–

inorganic halide perovskites and ferroelectric oxide perovskites, J. Mater. Chem. A, 2015 3: 18809-

18828, https://doi.org/10.1039/C5TA04235F

13. H. Matsuo, Y. Kitanaka, R. Inoue, Y. Noguchi, M. Miyayama, T. Kiguchi, and T. J. Konno,

Bulk and domain-wall effects in ferroelectric photovoltaics, Phys. Rev. B 94 (2016): 214111,

https://doi.org/10.1103/PhysRevB.94.214111

14. S. Pal, A. B. Swain, P. P. Biswas, D .Murali, A. Pal, B. R. K. Nanda and P. Murugavel,

Giant photovoltaic response in band engineered ferroelectric perovskite, Sci Rep 8 (2018)

8005, https://doi.org/10.1038/s41598-018-26205-x

15. P. Zhang, D. Cao, C. Wang, M. Shen, X. Su, L. Fang, W. Dong, F. Zheng , Enhanced

photocurrent in Pb(Zr0.2Ti0.8)O3 ferroelectric film by artificially introducing asymmetrical

interface Schottky barriers, Mater Chem and Phys 135 (2012) 304-308,

https://doi.org/10.1016/j.matchemphys.2012.04.041

16. R .Gaoa, W. Caia, G. Chena, X. Deng, X. Caoa, C. Fua, Enhanced ferroelectric photovoltaic

effect based on converging depolarization field, Mater Res Bull 84 (2016) 93–98,

http://dx.doi.org/10.1016/j.materresbull.2016.07.031

17. P. Pal, K. Rudrapal, P. Maji, A. R. Chaudhuri, and D. Choudhury, Toward an Enhanced

Room-Temperature Photovoltaic Effect in Ferroelectric Bismuth and Iron Co-doped BaTiO3,

J. Phys. Chem. C 125 (2021) 5315−5326, https://dx.doi.org/10.1021/acs.jpcc.0c10655

18. K. G. Baiju, B. Murali, D. Kumaresan, Ferroelectric barium titanate microspheres with

superior light-scattering ability for the performance enhancements of flexible polymer dye

sensitized solar cells and photodetectors, Solar Energy, 224 (2021) 93–101,

https://doi.org/10.1016/j.solener.2021.05.063

19. Z. Bai, Q. Zhang, Y. Zhang, Z. Huang, Y. Gao, J. Liu, X. Wang, Enhanced photocurrent

of self-powered ultraviolet photodetectors based on Ba1−xSrxTiO3 ceramics via ferroelectric


polarization, J Alloys compd 885 (2021) 161177,

https://doi.org/10.1016/j.jallcom.2021.161177

20. L. Wu, A. R. Akbashev, A. A. Podpirka, J.E. Spanier, P.K. Davies, Infrared‐ to‐ ultraviolet

light‐ absorbing BaTiO3‐ based ferroelectric photovoltaic materials, J Am Ceram Soc. 102

(2019)1–12, https://doi.org/10.1111/jace.16307

21. S. Hao, M. Yao, G. Vitali-Derrien, P. Gemeiner, M. Otonicˇar, P. Ruello, H. Bouyanfif,

Pierre-Eymeric Janolin, B. Dkhil and C. Paillard, Optical absorption by design in a

ferroelectric: co-doping in BaTiO3, J. Mater. Chem. C, 10 (2022) 227-234,

https://doi.org/10.1039/D1TC04250E

22. S. Das, S. Ghara, P. Mahadevan, A. Sundaresan, J. Gopalakrishnan, and D.D. Sarma,

Designing a Lower Band Gap Bulk Ferroelectric Material with a Sizable Polarization at Room

Temperature, ACS Energy Lett. 3 (2018) 1176−1182,

https://doi.org/10.1021/acsenergylett.8b00492

23. J. Zhang, X. Su, M. Shen, ZhihuaDai, L. Zhang, XiyunHe, W. Cheng, M.Cao1 and G. Zou,

Enlarging photovoltaic effect: combination of classic photoelectric and ferroelectric

photovoltaic effects, Sci Rep 3 (2013) 2109, https://doi.org/10.1038/srep02109

24. B. Hammer, L. B. Hansen and J. K. Nørskov, Improved adsorption energetics within

density-functional theory using revised Perdew-Burke-Ernzerhof functionals, Phys. Rev. B, ,

59 (1999) 7413-7421, https://doi.org/10.1103/PhysRevB.59.7413.

25. G. Kresse and D. Joubert, From Ultrasoft Pseudopotentials to the Projector Augmented-

Wave Method Phys. Rev. B, 1999 59: 1758-1775,

http://dx.doi.org/10.1103/PhysRevB.59.1758.
26. J. P. Perdew, K. Burke and M. Ernzerhof, Generalized Gradient Approximation Made

Simple, Phys. Rev. Lett., 77 (1996) 3865-3868, https://doi.org/10.1103/PhysRevLett.77.3865

27. H. J. Monkhorst and J. D. Pack, Special points for Brillouin-zone integrations,

Phys. Rev. B, 13 (1976) 5188-5192, https://doi.org/10.1103/PhysRevB.13.5188.

28. H. K. Chandra, K. Gupta, A. K. Nandy and P. Mahadevan, Ferroelectric distortions in doped

ferroelectrics: BaTiO3:M (M=V−Fe), Phys. Rev. B, 87 (2013) 214110,

https://doi.org/10.1103/PhysRevB.87.214110.

29. R. Wahl, D. Vogtenhuber and G. Kresse, SrTiO3 and BaTiO3 revisited using the projector

augmented wave method: Performance of hybrid and semilocal functionals, Phys. Rev. B, 78 (2008)

104116, https://doi.org/10.1103/PhysRevB.78.104116.

30. Zhuo Wang, Xiang Ming Chen,Evolution from relaxor-like dielectric to ferroelectric in

Ba[(Fe0.5Nb0.5)1−xTix]O3, Solid State Commun, 151 (2011) 708-711,

https://doi.org/10.1016/j.ssc.2011.02.015.

31. Venkidu L, V.G.Babu M, E.Rubavathi P, Bagyalakshmi B, Sundarakannan B, Structure,

microstructure, magnetic and magnetodielectric investigations on

BaTi(1-x-y)FexNbyO3 ceramics, Ceram Int, 44 (2018) 8161–8165,

https://doi.org/10.1016/j.ceramint.2018.01.263.

32. P.E. Rubavathi, L. Venkidu, M.V.G. Babu, R.V. Raman, B. Bagyalakshmi, S.M.A. Kader,

K. Baskar, M. Muneeswaran, N.V. Giridharan, B. Sundarakannan, Structure, morphology and

magnetodielectric investigations of BaTi1−xFexO3−δ ceramics, J Mater Sci: Mater Electron 30

(2019) 5706–5717. https://doi.org/10.1007/s10854-019-00864-6.

33. Petricek,V., Dusek, M.Palatinus, L.Z.Kristallogr, Crystallographic Computing System

JANA2006: General features, Crystalline Mater. 229 (2014) 345-352,

https://doi.org/10.1515/zkri-2014-1737.

34. L. Vegard, Die Konstitution der Mischkristalle und die Raumfüllung der Atome,
Z. Phys. 1921 5: 17–26.

35. J.A.Alonso, M.J.Martı´nez-Lope, and M.T.Casais, Evolution of the Jahn-Teller Distortion

of MnO6 Octahedra in RMnO3 Perovskites (R ) Pr, Nd, Dy, Tb, Ho, Er, Y): A Neutron

Diffraction Study, Inorg. Chem. 39 (2000) 917-923, https://doi.org/10.1021/ic990921e

36. L. Venkidu, A.B. Athanas, S. Kalaiyar, B. Sundarakannan, Fabrication of DSSCs using

ferroelectric photoanodes of co-substituted (1-x)BiFeO3-(x)BaFe1/2Nb1/2O3: Structural

correlation to bandgap reduction and its impact on power conversion efficiency, Mater Letter

311 (2022) 131473, https://doi.org/10.1016/j.matlet.2021.131473

37. B. Luo, X. Wang, E. Tian, H. Song, Q. Zhao, Z.Cai, W. Feng, L. Li, Giant permittivity and

low dielectric loss of Fe doped BaTiO3 ceramics: Experimental and first-principles

calculations, J Eur Ceramic Soc, 38 (2018) 1562-1568,

http://dx.doi.org/10.1016/j.jeurceramsoc.2017.10.014

38. L. Srisombat, S. Ananta, B. Singhana, T. RandallLee, R. Yimnirun, Chemical investigation

of Fe3+/Nb5+-doped barium titanate ceramics, Ceram Inter 39 (2013) S591–S594,

https://doi.org/10.1016/j.ceramint.2012.10.142

39. B. Luo, X. Wang, E. Tian, H. Song, Q .Zhao, Z. Cai, W. Feng, L. Li, Giant permittivity

and low dielectric loss of Fe doped BaTiO3 ceramics: Experimental and first-principles

calculations, J Eur Ceramic Soc, 38 (2018)1562-1568,

http://dx.doi.org/10.1016/j.jeurceramsoc.2017.10.014

40. M.M. Ismail, Ferroelectric characteristics of Fe/Nb co-doped BaTiO3, Modern Physics

Letters B, 33 (2019) 1950261, https://doi.org/10.1142/S0217984919502610

41. Heywang, W., Lubitz, K. and Wersing, W. eds., 2008. Piezoelectricity: evolution and

future of a technology (Vol. 114). Springer Science & Business Media.


42. P. Pal, K. Rudrapa, S. Mahana, S. Yadav, T. Paramanik, S. Mishra, K. Singh, G. Sheet, D.

Topwal, A.R. Chaudhuri, and D. Choudhury, Origin and tuning of room-temperature

multiferroicity in Fe-doped BaTiO3, Phys. Rev. B 101 (2020) 064409,

https://doi.org/10.1103/PhysRevB.101.064409

43. N.A. Hill, Why Are There so Few Magnetic Ferroelectrics?, J. Phys. Chem. B 104 (2000)

6694-6709, https://doi.org/10.1021/jp000114x

44. D. Phuyal, S. Mukherjee, S. Jana, F. Denoel, M.V. Kamalakar, S.M. Butorin, A.

Kalaboukhov, H. Rensmo, and O. Karis, Ferroelectric properties of BaTiO3 thin films co-doped

with Mn and Nb, AIP Advances 9 (2019) 095207 https://doi.org/10.1063/1.5118869

45. M.S. Alkathy, M.H. Lente, J.A. Eiras, Bandgap narrowing of Ba0.92Na0.04Bi0.04TiO3

ferroelectric ceramics by transition metals doping for photovoltaic applications, Mater Chem

and Phys 257 (2021): 123791, https://doi.org/10.1016/j.matchemphys.2020.123791

46. M.S. Alkathy, J.A .Eiras, F.L. Zabotto, K.C.J. Raju, Structural, optical, dielectric, and

multiferroic properties of sodium and nickel co‑ substituted barium titanate ceramics, J Mater

Sci: Mater Electron 32 (2021) 12828–12840 https://doi.org/10.1007/s10854-020-03900-y

47. F. Wang, I. Grinberg, and A.M. Rappe, Band gap engineering strategy via polarization

rotation in perovskite ferroelectrics, Appl. Phys. Lett., 104 (2014) 152903,

https://doi.org/10.1063/1.4871707

48. T. Wolfram, S. Ellialtioglu, Electronic and Optical Properties of D-Band

Perovskites, Cambridge University Press, New York, 2006.

49. S. Dahbi, N. Tahiri, O. El Bounagui, H. Ez-Zahraouy, Electronic, optical, and

thermoelectric properties of perovskite BaTiO3 compound under the effect of compressive

strain, Chemical Physics, 544 (2021) 111105,

https://doi.org/10.1016/j.chemphys.2021.111105
50. H. Pinto, A. Stashans, Computational study of self-trapped hole polarons in tetragonal

BaTiO3, Phys. Rev. B 65 (2002) 134304, https://doi.org/10.1103/PhysRevB.65.134304

51. Hsuan-Chung Wu, Sheng-Hong Li, and Syuan-Wei Lin, Effect of Fe Concentration on Fe-

Doped Anatase TiO2 from GGA + U Calculations, International Journal of Photoenergy, 2012

(2012), 823498, https://doi.org/10.1155/2012/823498

52. Y. Q. Wang, Y. Liu, M. X. Zhang & F. F. Min (2018) Electronic, magnetic and optical

properties of charge-compensated (Nb, TM=Fe, Cr)-codoped SrTiO3 from first principles,

Ferroelectrics, 537:1, 68-78, https://doi.org/10.1080/00150193.2018.1528958

53. N. Ma, and Y. Yang, Boosted Photocurrent via Cooling Ferroelectric BaTiO3 Materials for

Self-Powered 405 nm Light Detection, Nano energy, 60 (2019) 95-102,

https://doi.org/10.1016/j.nanoen.2019.03.036

54. N. Ma, K . Zhang, Y. Yang, Photovoltaic-Pyroelectric Coupled Effect Induced Electricity

for Self-Powered Photodetector System, Adv. Mater., 29 (2017) 1703694,

https://doi.org/10.1002/adma.201703694

55. Z. Jin, L. Gao, Q. Zhou, J. Wang, High-performance flexible ultraviolet photoconductors

based on solution-processed ultrathin ZnO/Au nanoparticle composite films, Sci Rep, 4 (2014)

4268, https://doi.org/10.1038/srep04268

56. Y. Lu, Y. Liu and P. Chen, High efficient photovoltaics in BaTiO3 thin film, Mater. Res.

Express 6 (2019) 116217, https://doi.org/10.1088/2053-1591/ab39a6

57. H. Fan, C. Chen, Z. Fan, L. Zhang, Z. Tan, P. Li, Z. Huang, J. Yao, G. Tian, Q.Luo, Z.Li,

X. Song, D. Chen, M. Zeng, J. Gao, X. Lu, Y. Zhao, X. Gao, and J.M. Liu, Resistive switching

and photovoltaic effects in ferroelectric BaTiO3-based capacitors with Ti and Pt top electrodes,

Appl. Phys. Lett., 111 (2017) 252901, https://doi.org/10.1063/1.4999982


58. D. A. Neamen D 2012 Semiconductor Physics and Devices Basic Principles 4th Edition,

McGraw-Hill Education, New York 2011

59. S. Y. Wang, M. Li, W. F. Liu, J. Gao, Resistive switching behavior of

BaTiO3/La0.8Ca0.2MnO3heterostructures, Phys Lette A, 379 (2015) 1288-1292,

https://doi.org/10.1016/j.physleta.2015.02.037

60. F. Wang and A. M. Rappe, First-principles calculation of the bulk photovoltaic effect in

KNbO3 and (K,Ba)(Ni,Nb)O3−δ, Phys. Rev. B 91 (2015) 165124,

https://doi.org/10.1103/PhysRevB.91.165124

61. A. P´erez-Tom´as, A. Mingorance, D. Tanenbaum, M. Lira-Cantú, Metal Oxides in

Photovoltaics: All-Oxide, Ferroic, and Perovskite Solar Cells In The Future of Semiconductor

Oxides in Next-Generation Solar Cells, Barcelona, Spain, (2018) 267–356,

https://doi.org/10.1016/B978-0-12-811165- 9.00008-9

62. A. M. Glass, D. von der Linde, and T. J. Negran, Highvoltage bulk photovoltaic effect and

the photorefractive process in LiNbO3, Appl. Phys. Lett. 25 (1974) 233,

http://dx.doi.org/10.1063/1.1655453

63. Brody, P.S., Crowne, F. Mechanism for the high voltage photovoltaic effect in ceramic

ferroelectrics. J. Electron. Mater. 4 (1975) 955–971 https://doi.org/10.1007/BF02660182

64. X. Lv, X-X.Zhang, and J. Wu, Nano-Domains in Lead-Free Piezoceramics: A Review, J.

Mater. Chem. A, 8 (2020) 10026-10073, https://doi.org/10.1039/D0TA03201H

65. S. Pal, S. Muthukrishnan, B. Sadhukhan, Sarath N. V., D. Murali, and P. Murugavel, Bulk

photovoltaic effect in BaTiO3-based ferroelectric oxides: An experimental and theoretical

study, J. Appl. Phys. 129 (2021) 084106, https://doi.org/10.1063/5.0036488

66. H. Xiang, F. Zhang, Z. Yi, M. Ma, Y. Gu, F. Liu, Y. Li, and Z. Liu, Non-stoichiometry

induced switching behavior of ferroelectric photovoltaic effect in BaTiO3 ceramics, Phys.

Status Solidi RRL, 13 (2019)1900074, https://doi.org/10.1002/pssr.201900074


67. Wen-Yuan Pan, Yu-Cheng Tang, Y. Yin, Ai-Zhen Song, Jing-Ru Yu, S. Ye, Bo-Ping

Zhang , Jing-Feng Li, Ferroelectric and photovoltaic properties of (Ba, Ca)(Ti, Sn, Zr)O3

perovskite ceramics, Ceram Inter, 47 (2021) 23453-23462

https://doi.org/10.1016/j.ceramint.2021.05.061

68. R. Su, D. Zhang, M. Wu, Fa-tang Li, Y. Liu, Z. Wang, C. Xu, X. Lou, Q. Yu, Y. Yang,

Plasmonic-enhanced ferroelectric photovoltaic effect in 0-3 type BaTiO3-Au ceramics, J Alloys

Compd 785 (2019) 584-589, https://doi.org/10.1016/j.jallcom.2019.01.223


Figure captions

Fig.1. (a) ADXRD patterns of BT and BFNT samples at RT. (b) Enlarged (002), (200)

reflections from BT and BFNT samples. Rietveld refinement of pure BT refined with (c)

P4mm (d) P4mm + Amm2 space groups (e-g) Rietveld refinement ADXRD patterns of

0.025BFNT, 0.05BFNT and 0.075BFNT ceramics.

Fig.2. Enlarged view of two nearby linked TiO6 octahedra with octahedral tilting of (a) BT and

(b) 0.075BFNT samples.

Fig.3. (a) XPS binding energy spectra of Ti-peak shifting and high-resolution core-level

spectra of (b) Fe-2p, (c) Nb-3d and (d) O-1s elements.

Fig.4. Polarization- Electric field (P-E) hysteresis loop of BT and BFNT samples at RT.

Fig.5. Absorption spectra of BT and BFNT samples as a function of (a) wavelength, and (b)

Tauc’s plot. The total density of states (TDOS) and projected density of states (PDOS) for the

structures (c) tetragonal BT and (d) cubic 0.075BFNT. (In both graphs, the Fermi level is set

to zero.)

Fig.6. (a) J-V characteristics of BT and BFNT samples. (b) Time-dependent photocurrent

density (J-t) curve. (c) Time-resolved photocurrent density rise and decay curves of BT and

0.075BFNT device (d) Graph comparing Voc, Jsc and PV power of BFNT with other BT-based

ferroelectric photovoltaics.

Table captions:

1. Deconvolution of O-1s peak parameters for BT and BFNT samples obtained from XPS

measurement.
2. Bandgap (Eg), Bandwidth (EB), 𝜂, and probability of d-orbital mixing obtained for BT and

BNFT samples.

3. Photovoltaic parameters determined for BT and BFNT samples.

Table 1

Sample O(I) O(II) (%) O(III)(%)

Binding Peak Binding Peak Binding Peak

energy area energy area energy area

(eV) (%) (eV) (%) (eV) (%)

BT 528.17 28.27 533.59 67.52 537.83 4.21

0.025BFNT 527.48 26.59 531.41 70.15 534.06 3.24

0.075BFNT 527.64 33.79 530.43 55.73 532.17 10.46

Table 2

Sample Eg (eV) EB (eV) 𝜂 rd (%)

BT 3.25 3.26 0.49816 15

0.025BFNT 2.68 3.14 0.42814 18

0.05BFNT 2.64 3.12 0.42290 18.5

0.075BFNT 2.55 3.10 0.37958 19


Table 3

Sample Jsc Voc FF η R D

(μA/cm2) (Volts) (%) (mA/W) (× 108

Jones)

BT 0.69 3.40 0.290 0.0006 0.0069 1.688

0.025BFNT 0.22 5.14 0.297 0.0003 0.0022 0.538

0.05BFNT 0.97 7.68 0.307 0.0022 0.0097 2.373

0.075BFNT 1.46 8.31 0.332 0.0041 0.0146 3.572


Figure-1 Click here to access/download;Figure(s);1ab.tif
Figure-2 Click here to access/download;Figure(s);2a b tilt re.tif
Figure-3 Click here to access/download;Figure(s);3ccc (1).tiff
Figure-4 Click here to access/download;Figure(s);4.tif
Figure-5 Click here to access/download;Figure(s);5.tif
Figure-6 Click here to access/download;Figure(s);6 final.tif
Supplementary and Electronic files

Click here to access/download


Supplementary and Electronic files
supporting information.docx
Declaration of Interest Statement

Declaration of interests

√ The authors declare that they have no known competing financial interests or personal
relationships that could have appeared to influence the work reported in this paper.

☐ The authors declare the following financial interests/personal relationships which may be
considered as potential competing interests:

You might also like