Next Article in Journal
Room Temperature Ammonia Gas Sensor Based on p-Type-like V2O5 Nanosheets towards Food Spoilage Monitoring
Next Article in Special Issue
Interpretation of Reflection and Colorimetry Characteristics of Indium-Particle Films by Means of Ellipsometric Modeling
Previous Article in Journal / Special Issue
From Solid-State Cluster Compounds to Functional PMMA-Based Composites with UV and NIR Blocking Properties, and Tuned Hues
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Band Gap Tuning in Transition Metal and Rare-Earth-Ion-Doped TiO2, CeO2, and SnO2 Nanoparticles

by
Iliana Apostolova
1,
Angel Apostolov
2 and
Julia Wesselinowa
3,*
1
University of Forestry, Kl. Ohridsky Blvd. 10, 1756 Sofia, Bulgaria
2
University of Architecture, Civil Engineering and Geodesy, Hristo Smirnenski Blvd. 1, 1046 Sofia, Bulgaria
3
Sofia University “St. Kliment Ohridski”, J. Bouchier Blvd. 5, 1164 Sofia, Bulgaria
*
Author to whom correspondence should be addressed.
Nanomaterials 2023, 13(1), 145; https://doi.org/10.3390/nano13010145
Submission received: 30 November 2022 / Revised: 22 December 2022 / Accepted: 24 December 2022 / Published: 28 December 2022
(This article belongs to the Special Issue Functional Nanostructured Materials—from Synthesis to Applications)

Abstract

:
The energy gap E g between the valence and conduction bands is a key characteristic of semiconductors. Semiconductors, such as TiO 2 , SnO 2 , and CeO 2 have a relatively wide band gap E g that only allows the material to absorb UV light. Using the s-d microscopic model and the Green’s function method, we have shown two possibilities to reduce the band-gap energy E g —reducing the NP size and/or ion doping with transition metals (Co, Fe, Mn, and Cu) or rare earth (Sm, Tb, and Er) ions. Different strains appear that lead to changes in the exchange-interaction constants, and thus to a decrease in E g . Moreover, the importance of the s-d interaction, which causes room-temperature ferromagnetism and band-gap energy tuning in dilute magnetic semiconductors, is shown. We tried to clarify some discrepancies in the experimental data.

1. Introduction

Dilute magnetic semiconductors (DMS)—bulk and nanostructures, doped with transition metal (TM) ions, have been intensively investigated in recent years due to the number of unusual electronic, magnetic, and optical properties. Recently, room-temperature ferromagnetism has been observed experimentally and theoretically in pure and ion-doped semiconducting TiO 2 , SnO 2 , and CeO 2 nanostructures due to uncompensated spins or to oxygen vacancies at the surface and doping effects [1,2,3,4,5]. It is known that TiO 2 , SnO 2 , and CeO 2 have a relatively wide band gap E g (∼3.2, ∼3.6, and ∼3 eV, respectively) [6,7,8], which limits their photocatalytic activity to the ultraviolet region of light. One effective way to overcome these problems is to modify the band gap of these compounds is doping with different elements, such as metals and nonmetals, which affect E g in different ways. Transition metals, such as Co, Fe, Ni, and Mn, lead to decreasing E g in TiO 2 , SnO 2 , and CeO 2 NPs, as observed experimentally by many authors [9,10,11,12,13,14,15,16,17,18,19,20,21,22], improving the absorption of visible light. It must be mentioned that Salah et al. [23] and Venugopal et al. [24] observed an increase in E g in Mn-doped SnO 2 NPs, i.e., there are some discrepancies.
In order to reduce the electron-hole recombination and to shift the absorption wavelength to the desired visible region ( λ > 400 nm), a decrease in the band gap of rare earth (RE)-doped TiO 2 , CeO 2 , and SnO 2 bulk and nanoparticles (NPs) has been considered, which improves the absorption of visible light [25,26,27,28,29,30,31,32,33]. Let us emphasize that there are also results with enhanced E g by increasing the RE ion concentration [7,34]. Long et al. [33] obtained, in Sn-doped TiO 2 , a minimum of the band gap E g as a function of the doping concentration. The doping effects in Sn-doped TiO 2 are reported to lead to the red-shift or blue-shift of E g [35,36] as well as in La-doped TiO 2 [34,37]. Again, there are some discrepancies. First-principle studies are used to analyze the optical properties of TiO 2 , SnO 2 , and CeO 2 doped with TM [38,39,40,41,42,43] and RE ions [44,45,46,47].
The aim of the present paper is using the s d model to investigate on a microscopic level the band gap tuning of TM- and RE-doped TiO 2 , SnO 2 , and CeO 2 NPs, which are promising candidates for applications of visible-light photocatalytic activities, and to explain the discrepancies reported in the experimental data of the band gap E g values. We will show that lattice deformations, lattice parameters, and surface and doping effects are strongly correlated with E g values.

2. Model and Method

The Hamiltonian, which describes the magnetic properties of TM and RE ion-doped DMS NPs, is the s-d(f) Hamiltonian proposed for ferromagnetic semiconductors [4]
H s d = H s p + H e l + H s p e l .
H s p is the Heisenberg model of the d ( f ) -electrons in TM or RE metal ions
H s p = i , j x J i j ( S i + S j + S i S j + + 2 S i z S j z ) i D i ( S i z ) 2 .
Here, S i + , S i , and S i z are the spin-operators for the localized spins at site i, J i j is the magnetic exchange interaction between nearest neighboring sites i and j, D i is the single-site anisotropy parameter, and x is the ion-doping concentration.
The Hamiltonian of the conduction band electrons H e l is observed after development according to Wannier functions and limited to one band taking into account interactions between different Wannier states [48]:
H e l = i j σ t i j c i σ + c j σ + 1 2 i j k l , σ σ v ( i j k l ) c i σ + c j σ + c k σ c l σ ,
where t i j is the hopping integral, v is the Coulomb interaction, and c i σ + and c i σ are the Fermi-creation and -annihilation operators.
The operator H s p e l couples the two subsystems—Equations (2) and (3)—by an intra-atomic exchange interaction I i :
H s p e l = i x I i S i s i .
The spin operators s i of the conduction electrons at site i can be expressed as s i + = c i + + c i , s i z = ( c i + + c i + c i + c i ) / 2 . The important role of the s-d interaction I, taken into account in our paper, is supported by many experimental data. Room-temperature ferromagnetism may be due to s p d exchange interactions (as observed in optical spectra for DMS), which may be responsible for the ferromagnetic properties of the samples [49]. The carriers are spin-polarized to mediate the ferromagnetic ordering of the magnetic moments of transition metal ions doped into the semiconductor host lattice. Moreover, for example, Kaushik et al. [50] have shown that the optical spectra have a red shift with increasing Co in TiO 2 due to s p d exchange interactions. The shift of E g in Mn-doped SnO 2 NPs could also be attributed to the s p d exchange interactions [51].
The band-gap energy E g is defined by the difference between the valence and conduction bands:
E g = ω + ( k = 0 ) ω ( k = k σ ) .
The electronic energies
ω ± ( k ) = ϵ k σ 2 I S z + k [ v ( o ) v ( k k ) ] n k σ
are observed from the Green’s functions g ( k σ ) = c k σ ; c k σ + , σ = ± 1 . n k σ is the occupation number distribution. S z is the magnetization calculated in our previous work [4].
For the approximate calculation of the Green’s functions g ( k σ ) , we use a method proposed by Tserkovnikov [52]. After a formal integration of the equation of motion for g i j , one obtains
g i j ( t ) = i θ ( t ) [ c i ; c j + ] exp ( i ω i j ( t ) t )
where
ω i j ( t ) = ω i j i t 0 t d t t [ j i ( t ) ; j j + ( t ) ] [ c i ( t ) ; c j + ( t ) ] [ j i ( t ) ; c j + ( t ) ] [ c i ( t ) ; j j + ( t ) ] [ c i ( t ) ; c j + ( t ) ] 2
with the notation j i ( t ) = [ c i , H i n t e r a c t i o n ] . The time-independent term
ω i j = [ [ c i , H ] ; c j + ] [ c i ; c j + ]
is the spin excitation energy in the generalized Hartree–Fock approximation. The time-dependent term in Equation (8) includes damping effects.

3. Results and Discussion

For the numerical calculations of ion-doped TiO 2 , we have taken the following model parameters: J = 0.31 meV, D = −0.1 meV, I = 60 meV, v = 0.3 meV, F = 4 cm 1 , R = −0.35 cm 1 , J(Co-Co) = 44 meV [53], J(Cu-Cu) = 30 meV [54], J(Fe-Fe) = 36.5 meV [55], J(Sm-Sm) = 0.3 meV [5], and J(Tb-Tb) = 1.5 meV [56], J(Er-Er) = 6 meV [57].
For example, the TiO 2 NP has a spherical shape and the Ti spins are situated into shells numbered by n = 1 , . . . , N , where n = 1 is the central spin and n = N —the surface shell, due to the changed number of next neighbors on the surface and to the fact that the reduced symmetry J s on the surface is different from that in the bulk J b , as well as due to different strains J d in the doped states that are different to the undoped ones J b . Moreover, J is inverse-proportional to the lattice parameters.

3.1. Size Dependence of the Band-Gap Energy

The first way to narrow the band gap E g of TiO 2 , SnO 2 , and CeO 2 is to reduce the size of the pure undoped compound. In general, in TiO 2 the interaction between the Ti 4 + ions (S = 0) is diamagnetic. However, on the surface of non-doped TiO 2 NPs due to uncompensated spins, there Ti 3 + ions appear with S 0 , and thus magnetism is induced. Santara et al. [58] have observed that the reduction in the particle size leads to an increase in the lattice parameter, i.e., to an increase in tensile strain. In our model, this would lead to a decrease in the exchange interactions J s on the surface compared to the bulk value J b , i.e., J s < J b . In Figure 1, curve 1, the size dependence of E g for TiO 2 is shown. It can be seen that E g decreases with decreasing NP size, as reported experimentally by Chen et al. [29], Kalathil et al. [59], Garcia et al. [60], and Dette et al. [61]. This decrease in E g with the decrease in NP size is valid for all three compounds and is shown in Figure 1, curves 2 and 3, for CeO 2 and SnO 2 , respectively, in coincidence with the experimental data for CeO 2 NPs of Ansari et al. [8] and Tatar et al. [62] but not of Tamizhdurai et al. [63], as well as for SnO 2 NPs of Kamarulzaman et al. [64], but in disagreement with the result of Asaithambi et al. [65], who reported an increase in E g with the decrease in the size of SnO 2 NPs.

3.2. Ion-Doping Dependence of the Band-Gap Energy

3.2.1. Ion-Doping Dependence of the Band-Edge Energies

The conduction band (CB) in TiO 2 is dominated by the empty Ti d-band, whereas the valence band (VB) is composed of the occupied O p-band and Ti d-band. Fujisawa et al. [66] and Dorenbos [67] have observed the valence and conduction band energies in bulk TiO 2 to be VB = −7.25 (−7.2) eV and CB = −4.05 (−4.0) eV, respectively. For SnO 2 , the reported values are for VB = −7.76 [67] (−8.2 eV [68]) and for CB = −4.14 [67] (−4.6 eV [68]). However, it must be noted that the data are not consistent. Ion doping can move the band edges of the VB and/or the CB. Next, we will study the influence of the doping with a TM, for example, Fe, and an RE ion, for example, Sm, on the VB and CB energies in bulk TiO 2 and SnO 2 . The transition metal ions form a dopant level above the VB band of TiO 2 [10,37,69]. Fe 3 + is known to create shallow trapping sites at the donor and acceptor levels [37]. This leads to a decrease in the band gap of Fe-doped TiO 2 . In Figure 2, curve 1, the calculation from Equation (6), using the s-d model, VB energy as a function of the Fe concentration, is presented. It can be seen that there is an increase in the VB edge energy. This increase is due to the contribution from the lower Fe 3d band, which benefits the hole mobility in VB. As a result of Fe-doping, the electron transition energy from the VB to the CB decreases, which may induce a red shift at the edge of the optical-absorption range. A broadening of the VB in Fe-doped TiO 2 is reported by Wu et al. [11]. The incorporation of RE ions, for example, Sm, into the TiO 2 host modifies the band gap of TiO 2 with sub-band-gap energy levels of RE ions under the CB. These energy levels offer an electronic transition from the TiO 2 valence band to the empty RE ion sub-band-gap energy levels. From Figure 2, curve 2, a reduction in the calculated CB edge energy in bulk Sm-doped TiO 2 can be seen, as observed experimentally by Wei et al. [70], too. We also observed similar behavior for the dependence of the VB and CB energies of Fe (curve 3) and Sm (curve 4) doped bulk SnO 2 .

3.2.2. Transition-Metal-Ion (Co, Fe, Mn, and Cu) Doping Effect on the Band-Gap Energy

Next, we will study the band-gap energy E g in a TM-doped TiO 2 NP, Ti 1 x TM x O 2 , for example, with Fe. E g is calculated with our model from Equation (5). The difference of the ionic radius of Fe 3 + (0.64 A ˙ ) to that of Ti 4 + (0.68 A ˙ ) [71] means that the doped ion has a smaller ionic radius than the host ion, i.e., compressive strain appears. Our calculations show a decrease in the band-gap energy E g with increasing Fe-doping concentration x in a TiO 2 NP (see Figure 3, curve 2), which is in agreement with other results [37,38,41,43,72]. The photochemical studies of George et al. [9] showed that band-gap energy E g in Fe-doped TiO 2 was reciprocally tuned proportional to the Fe content. Shortly, we will discuss the importance of the s-d interaction. It must be noted that for simplification we only consider the s electrons and the s d interaction. Due to the compressive strain, the exchange-interaction constants between the Ti 3 and the Ti 3 + -Fe 3 + ions lead to a small magnetization, which increases with increasing Fe-doping concentration. In order to explain the data strong decrease in the band gap by TM ion doping observed in the experimental data, we must take the large s-d interaction (see Equation (6)) into account, which strongly reduces the band-gap energy E g .
In Figure 3 (curves 1 and 4), the band gap reduction in TiO 2 NP with other TM ions is demonstrated, for example, with Co 2 + (0.745 A ˙ ) [50] and Cu 2 + (0.87 A ˙ ) [65], whose ionic radii are larger than the radius of the host Ti ion, i.e., unlike in the Fe ion doping, tensile strain appears. We have to use the relation J d < J b . This means that the exchange interaction between the Ti 3 + ions in the doped states would decrease compared to the undoped ones, and we would observe a small magnetization M that would slowly decrease with increasing Co or Cu dopants. It must be noted that the double exchange interaction energy between these TM ions and the Ti 3 + ion is ferromagnetic and stronger than that between the Ti 3 + -Ti 3 + ions; together with the strong s d interaction, it can change the behavior of the magnetization M, M, and it can increase when the concentration x is raised. Thus, we again obtain a decrease in the band-gap energy E g as observed in [10,50,73,74,75]. Let us emphasize that similar behavior is also reported in TM-doped CeO 2 [13,72,76,77,78] and SnO 2 NPs [12,65,79]. In order to clarify the discrepancies by Mn ion doping, let us emphasize that, with our model, we obtain a reduced band gap in a Mn 2 + (r = 0.8 A ˙ )-doped SnO 2 NP using J(Mn-Mn) = 20.24 meV [12] (see Figure 3, curve 3). The ionic radius of the doped Mn 2 + ion is smaller than that of the host Sn 4 + ion (0.83 A ˙ ), i.e., again there appears a compressive strain; we have J d > J b . The substitution of Sn 4 + by Mn 2 + would require the formation of oxygen vacancies for charge balance, which is important for the RTFM and the band-gap reducing. It is advantageous to the photocatalytic activity, too [80]. Our result is in good qualitative agreement with that of Chatterjee et al. [81] for Mn-doped CeO 2 NPs, as well as with the behavior in Mn-doped TiO 2 [17,37] and Mn-doped SnO 2 NPs [14,82], but in disagreement with Refs. [23,24,83], which reported an increase in the band-gap energy E g in Mn-doped SnO 2 NPs despite the observed compressive strain. In our opinion, these discrepancies are due to the experimental methodology of synthesis and of growth, to the method of doping, and to the method of annealing.

3.2.3. Rare Earth (Sm, Tb, and Er) Ion Doping Effects on the Band-Gap Energy

Next, we will study the observed decrease in the band-gap energy E g in a RE (for example, Sm, Tb, and Er)-doped TiO 2 , Ti 1 x RE x O 2 NP. The ionic radius of Ti 4 + is smaller than that of Sm 3 + = 1.09 A ˙ , and it is smaller than the ionic radius of the most RE ions. This means that there is a tensile strain. In Figure 4, we have calculated the band-gap energy E g as a function of the ion-doping concentration for different RE ions. It can be seen that E g decreases with increasing x. This is due to the reduction in the CB edge energy (see Figure 2, curve 2), reported also by Wei et al. [70]. Let us emphasize that the role of the intra-atomic s d exchange interaction is very important here. The situation is in analogy with the doping with the TM ions Co and Cu. The tensile strain leads to changes in the exchange-interaction constants J d between the Ti 3 + ions in the doped states and the undoped ones J b , i.e., we have to use J d < J b . It must be noted that if we only take this interaction into account, then we will observe a small decrease in the magnetization [5] and of the band gap E g . However, taking into account the strong s d interaction and the interaction between Ti and the doping ion, the magnetization M changes its behavior, and it increases with increasing doping concentration x. An increase in M with increasing Sm and Nd ion doping is reported in [84]. Therefore, the band-gap energy E g decreases (see Figure 4, curve 1) in coincidence with [70,85] but not with [20]. The reduction in the band-gap energy of Sm-doped TiO 2 indicates a red shift of the light adsorption. A decrease in E g in Sm-doped CeO 2 NPs is also observed in [86,87,88], which can increase the photocatalytic activity. The decrease in E g with increasing Tb (r = 1.06 A ˙ ) or Er (r = 1.03 A ˙ ) doping concentration is presented in Figure 4, curves 2 and 3, in coincidence with Lee et al. for Er-doped TiO 2 thin films [89]. We obtain within our s d model a decrease in E g in TiO 2 , CeO 2 , and SnO 2 with the most RE ions. Moreover, the band-gap energy E g decreases with the increase in the s d interaction I. Let us emphasize that there are some discrepancies by the RE ion doping. A decrease in E g for small doping concentrations of different RE ions (for example, Sm, La, Dy, Nd, Eu, Er, and Tb) is also reported in TiO 2 , CeO 2 , or SnO 2 NPs [70,85,90,91,92,93,94]. However, some authors have observed an increase in the band-gap energy E g in Sm [95] and Yb, Sc [45]-doped TiO 2 NPs, and Pr-doped CeO 2 NPs [96].
It must be mentioned that pure bulk TiO 2 without defects and impurities is diamagnetic due to the presence of Ti 4 + , which has no unpaired electrons, making it diamagnetic. Doping TiO 2 with TM or RE ions in order to induce unpaired spins introduces the required magnetic ordering to this compound. Due to the exchange interactions between the substituted doping ions on the Ti sites, there appears a magnetic moment, a spontaneous magnetization M different from zero, which increases with the increase in the doping concentration x (see Figure 5). The observed RTFM in ion-doped bulk TiO 2 is reported in [43,46,69]. Moreover, E g in TM- and RE-doped bulk TiO 2 decreases (see Figure 6) in good qualitative agreement with the experimental data [11,38,43]. This is also valid for the other two compounds CeO 2 and SnO 2 .

4. Conclusions

In conclusion, narrowing the optical band gap of TiO 2 , SnO 2 , and CeO 2 NPs is essential for visible light applications. Using the s d microscopic model, we have shown on a microscopic level two possibilities to reduce the band-gap energy E g of CeO 2 , TiO 2 , or SnO 2 NPs, reducing the NP size and/or ion doping with TM (Co, Fe, Mn, and Cu) or RE (Sm, Tb, and Er) ions, which can improve the photocatalytic activity. Different strains appear due to the different ionic radii of the TM or RE ions and the host ions, which lead to lattice deformations. Furthermore, the discrepancies by the E g behavior in Mn-doped and RE-doped TiO 2 , CeO 2 , and SnO 2 NPs are discussed. We have shown that the decrease in the band-gap energy E g by RE doping is much smaller than that by TM doping. Therefore, one can conclude that the doping with TM ions is a more effective method than the doping with RE ions in order to reduce the band width E g in CeO 2 , TiO 2 , and SnO 2 NPs and thus to enhance the photocatalytic activity.

Author Contributions

All authors contributed equally to this work. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The raw data that support the findings of this study are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sundaresan, A.; Rao, C.N.R. Ferromagnetism as a Universal Feature of Inorganic Nanoparticles. Nano Today 2009, 4, 96. [Google Scholar] [CrossRef]
  2. Chang, G.S.; Forrest, J.; Kurmaev, E.Z.; Morozovska, A.N.; Glinchuk, M.D.; McLeod, J.A.; Moewes, A.; Surkova, T.P.; Hong, N.H. Oxygen-vacancy-induced ferromagnetism in undoped SnO2 thin films. Phys. Rev. B 2012, 85, 165319. [Google Scholar] [CrossRef]
  3. Kumar, S.; Kim, Y.J.; Koo, B.H.; Lee, C.G. Structural and Magnetic Properties of Co Doped CeO2 Nano-Particles. J. Nanosci. Nanotechnol. 2010, 10, 7204–7207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Apostolov, A.T.; Apostolova, I.N.; Wesselinowa, J.M. A comparative study of the magnetization in transition metal ion-doped CeO2, TiO2 and SnO2 nanoparticles. Phys. E 2018, 99, 202–207. [Google Scholar] [CrossRef]
  5. Apostolov, A.T.; Apostolova, I.N.; Wesselinowa, J.M. Magnetic Properties of Rare Earth Doped SnO2, TiO2 and CeO2 Nanoparticles. Phys. Status Solidi B 2018, 255, 1800179. [Google Scholar] [CrossRef]
  6. Zhang, Y.; Xu, X. Machine Learning Band Gaps of Doped-TiO2 Photocatalysts from Structural and Morphological Parameters. ACS Omega 2020, 25, 15344–15352. [Google Scholar] [CrossRef]
  7. Nguyen, T.B.; Le, T.T.B.; Nguyen, N.L. The preparation of SnO2 and SnO2:Sb nanopowders by a hydrothermal method. Adv. Nat. Sci. Nanosci. Nanotechnol. 2010, 1, 025002. [Google Scholar] [CrossRef] [Green Version]
  8. Ansari, S.A.; Khan, M.M.; Ansari, M.O.; Lee, J.; Cho, M.H. Band Gap Engineering of CeO2 Nanostructure by Electrochemically Active Biofilm for Visible Light Applications. RSC Adv. 2014, 4, 16782–16791. [Google Scholar] [CrossRef]
  9. George, S.; Pokhrel, S.; Ji, Z.; Henderson, B.L.; Xia, T.; Li, L.J.; Zink, J.I.; Nel, A.E.; Maedler, L. Role of Fe Doping in Tuning the Band Gap of TiO2 for the Photo-Oxidation-Induced Cytotoxicity Paradigm. J. Am. Chem. Soc. 2011, 133, 11270–11278. [Google Scholar] [CrossRef] [Green Version]
  10. Nebi, M.; Peker, D.; Temel, S. Deposition of Co doped TiO2 films using sol gel spin coating technique and investigation of band gap. AIP Conf. Proc. 2018, 1935, 150004. [Google Scholar]
  11. Wu, T.S.; Chen, Y.W.; Weng, S.C.; Lin, C.N.; Chang, C.; Soo, Y.L. Dramatic band gap reduction incurred by dopant coordination rearrangement in Co-doped nanocrystals of CeO2. Sci. Rep. 2017, 7, 4715. [Google Scholar] [CrossRef] [PubMed]
  12. Sharma, A.; Varshney, M.; Kumar, S.; Verma, K.D.; Kumar, R. Magnetic properties of Fe and Ni doped SnO2 nanoparticles. Nanomater. Nanotechnol. 2011, 1, 29–33. [Google Scholar] [CrossRef]
  13. Kumari, K.; Aljawfi, R.N.; Vij, A.; Chae, K.H.; Hashim, M.; Alvi, P.A.; Kumar, S. Band gap engineering, electronic state and local atomic structure of Ni doped CeO2 nanoparticles. J. Mater. Sc. Mater. Electr. 2019, 30, 4562–4571. [Google Scholar] [CrossRef]
  14. Ungureanu, A.; Jitaru, I.; Gosnea, F. Mn-doped SnO2 prepared by a sol-gel method. UPB Sci. Bull. Series B 2013, 75, 43. [Google Scholar]
  15. Nakaruk, A.; Lin, C.Y.W.; Channei, D.; Koshy, P.; Sorrell, C.C. Fe-doped and Mn-doped titanium dioxide thin films. J. Sol-Gel Sci. Techn. 2012, 61, 175–178. [Google Scholar] [CrossRef]
  16. Deng, Q.R.; Xia, X.H.; Guo, M.L.; Gao, Y.; Shao, G. Mn-doped TiO2 nanopowders with remarkable visible light photocatalytic activity. Mater. Lett. 2011, 65, 2051–2054. [Google Scholar] [CrossRef]
  17. Chauhan, R.; Kumar, A.; Chaudhary, R.P. Structural and photocatalytic studies of Mn-doped TiO2 nanoparticles. Spectrochim. Acta Part A 2012, 98, 256–264. [Google Scholar] [CrossRef]
  18. Sagadevan, S.; Chowdhury, Z.Z.; Johan, M.R.B.; Aziz, F.A.; Roselin, L.S.; Podder, J.; Lett, J.A.; Selvin, R. Cu-Doped SnO2 Nanoparticles: Synthesis and Properties. J. Nanosci. Nanotechnol. 2019, 19, 7139–7148. [Google Scholar] [CrossRef]
  19. Munir, S.; Shah, S.M.; Hussain, H.; Khan, R.A. Effect of carrier concentration on the optical band gap of TiO2 nanoparticles. Mater. Des. 2016, 92, 64–72. [Google Scholar] [CrossRef]
  20. Babu, M.H.; Dev, B.C.; Podder, J. Texture coefficient and band gap tailoring of Fe-doped SnO2 nanoparticles via thermal spray pyrolysis. Rare Metals 2022, 41, 1332–1341. [Google Scholar] [CrossRef]
  21. Poudel, M.B.; Ojhab, G.P.; Akim, A.; Kim, H.J. Manganese-doped tungsten disulfide microcones as binder-free electrode for high performance asymmetric supercapacitor. J. Energy Storage 2022, 47, 103674. [Google Scholar] [CrossRef]
  22. Poudel, M.B.; Yu, C.; Kim, H.J. Synthesis of Conducting Bifunctional Polyaniline/Mn-TiO2 Nanocomposites for Supercapacitor Electrode and Visible Light Driven Photocatalysis. Catalysts 2020, 10, 546. [Google Scholar] [CrossRef]
  23. Salah, N.; Habib, S.; Azam, A.; Ansari, M.S.; Al-Shawafi, W.M. Formation of Mn-Doped SnO2 Nanoparticles Via the Microwave Technique: Structural, Optical and Electrical Properties. Nanomater. Nanotechnol. 2016, 6, 1. [Google Scholar] [CrossRef] [Green Version]
  24. Venugopal, B.; Nandan, L.; Ayyachamy, A.; Balaji, V.; Amirthapandian, S.; Panigrahi, B.K.; Paramasivam, T. Influence of manganese ions in the band gap of tin oxide nanoparticles: Structure, microstructure and optical studies. RSC Adv. 2014, 4, 6141. [Google Scholar] [CrossRef]
  25. Mehtab, A.; Ahmed, J.; Alshehri, S.M.; Mao, Y.; Ahmad, T. Rare earth doped metal oxide nanoparticles for photocatalysis: A perspective. Nanotechn. 2022, 33, 142001. [Google Scholar] [CrossRef]
  26. Rehani, D.; Saxena, M.; Solanki, P.R.; Sharma, S.N. Transition Metal and Rare-Earth Metal Doping in SnO2 Nanoparticles. J. Supercond. Novel Magn. 2022, 35, 2573–2581. [Google Scholar] [CrossRef]
  27. Saqib, N.U.; Adnan, R.; Shah, I. A mini-review on rare earth metal-doped TiO2 for photocatalytic remediation of wastewater. Environ. Sci. Pollut. Res. Int. 2016, 23, 15941–15951. [Google Scholar] [CrossRef]
  28. Chandran, D.; Nair, L.S.; Balachandran, S.; Babu, K.R.; Deepa, M. Band gap narrowing and photocatalytic studies of Nd3+ ion-doped SnO2 NPs using solar energy. Bull. Mater. Sci. 2016, 39, 27–33. [Google Scholar] [CrossRef] [Green Version]
  29. Chen, X.; Luo, W. Optical spectroscopy of rare earth ion-doped TiO2 nanophosphors. J. Nanosci. Nanotechnol. 2010, 10, 1482–1494. [Google Scholar] [CrossRef]
  30. Priyanka, K.P.; Tresa, S.A.; Jaseentha, O.P.; Varghese, T. Cerium doped nanotitania-extended spectral response for enhanced photocatalysis. Mater. Res. Express 2014, 1, 015003. [Google Scholar] [CrossRef]
  31. Song, L.; Zhao, X.; Cao, L.; Moon, J.-W.; Gu, B.; Wang, W. Synthesis Rare Earth Doped TiO2 Nanorods and Their Application in the Photocatalytic Degradation of Lignin. Nanoscale 2015, 7, 16695–16703. [Google Scholar] [CrossRef] [PubMed]
  32. Amritha, A.; Sundararajan, M.; Rejith, R.G.; Mohammed-Aslam, A. La-Ce doped TiO2 nanocrystals: A review on synthesis, characterization and photocatalytic activity. SN Appl. Sci. 2019, 1, 1441. [Google Scholar] [CrossRef]
  33. Long, R.; Dai, Y.; Huang, B. Geometric and Electronic Properties of Sn-Doped TiO2 from First-Principles Calculations. J. Phys. Chem. C 2009, 113, 650–653. [Google Scholar] [CrossRef]
  34. Priyanka, K.P.; Revathy, V.R.; Rosmin, P.; Thrivedu, B.; Elsa, K.M.; Nimmymol, J.; Balakrishna, K.M.; Varghese, T. Influence of La doping on structural and optical properties of TiO2 nanocrystals. Mater. Character. 2016, 113, 144–151. [Google Scholar] [CrossRef]
  35. Mahanty, S.; Roy, S.; Sen, S. Effect of Sn doping on the structural and optical properties of sol-gel TiO2 thin films. J. Cryst. Growth 2004, 261, 77–81. [Google Scholar] [CrossRef]
  36. Lin, J.; Yu, J.C.; Lo, D.; Lam, S.K. Photocatalytic Activity of Rutile Ti1-xSnxO2 Solid Solutions. J. Catal. 1999, 183, 368–372. [Google Scholar] [CrossRef]
  37. Lin, C.Y.W.; Channei, D.; Koshy, P.; Nakaruk, A.; Sorrell, C.C. Multivalent Mn-doped TiO2 thin films. Phys. E 2012, 44, 1969–1972. [Google Scholar] [CrossRef]
  38. Wang, Y.; Zhang, R.; Li, J.; Li, J.; Lin, S. First-principles study on transition metal-doped anatase TiO2. Nanoscale Res. Lett. 2014, 9, 46. [Google Scholar] [CrossRef] [Green Version]
  39. Zhou, Z.; Li, M.; Guo, L. A first-principles theoretical simulation on the electronic structures and optical absorption properties for O vacancy and Ni impurity in TiO2 photocatalysts. J. Phys. Chem. Solids 2010, 71, 1707–1712. [Google Scholar] [CrossRef]
  40. Cai, X.; Zhang, P.; Wei, S.-H. Revisit of the band gaps of rutile SnO(2 and TiO2: A first-principles study. J. Semicond. 2019, 40, 092101. [Google Scholar] [CrossRef]
  41. Mendez-Galvan, M.; Celaya, C.A.; Jaramillo-Quintero, O.; Muniz, J.; Diaz, G.; Lara-Garcia, H.A. Tuning the band gap of M-doped titanate nanotubes (M = Fe, Co, Ni, and Cu): An experimental and theoretical study. Nanoscale Adv. 2021, 3, 1382–1391. [Google Scholar] [CrossRef] [PubMed]
  42. Chetri, P.; Saikia, B.; Choudhury, A. Exploring the structural and magnetic properties of TiO2/SnO2 core/shell nanocomposite: An experimental and density functional study. J. Appl. Phys. 2013, 113, 233514. [Google Scholar] [CrossRef]
  43. Begna, W.B.; Gurmesa, G.S.; Zhang, Q.; Geffe, C.A. A DFT+U study of site dependent Fe-doped TiO2 diluted magnetic semiconductor material: Room-temperature ferromagnetism and improved semiconducting properties. AIP Adv. 2022, 12, 025002. [Google Scholar] [CrossRef]
  44. Shi, H.; Hussain, T.; Ahuja, R.; Kang, T.W.; Luo, W. Role of vacancies, light elements and rare-earth metals doping in CeO2. Sci. Rep. 2016, 6, 31345. [Google Scholar] [CrossRef] [PubMed]
  45. Bian, L.; Song, M.; Suker, Z.; Zhao, X.; Dai, Q. Band gap calculation and photo catalytic activity of rare earths doped rutile TiO2. J. Rare Earths 2009, 27, 461–468. [Google Scholar] [CrossRef]
  46. Lamrani, A.F. Rare-earth-doped TiO2 rutile as a promising ferromagnetic alloy for visible light absorption in solar cells: First principle insights. RSC Adv. 2020, 10, 35505. [Google Scholar] [CrossRef]
  47. Xie, K.; Jia, Q.; Wang, Y.; Zhang, W.; Xu, J. The Electronic Structure and Optical Properties of Anatase TiO2 with Rare Earth Metal Dopants from First-Principles Calculations. Materials 2018, 11, 179. [Google Scholar] [CrossRef] [Green Version]
  48. Elk, K.; Gasser, W. Green’s Function Method in the Solid State Physics; Akademie: Berlin, Germany, 1979; p. 61. [Google Scholar]
  49. Ohtsuki, T.; Chainani, A.; Eguchi, R.; Matsunami, M.; Takata, Y.; Taguchi, M.; Nishino, Y.; Tamasaku, K.; Yabashi, M.; Ishikawa, T.; et al. Role of Ti 3 d Carriers in Mediating the Ferromagnetism of Co∶ TiO 2 Anatase Thin Films. Phys. Rev. Lett. 2011, 106, 047602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Kaushik, A.; Dalela, B.; Kumar, S.; Alvi, P.A.; Dalela, S. Role of Co doping on structural, optical and magnetic properties of TiO2. J. Alloy. Compd. 2013, 552, 274. [Google Scholar] [CrossRef]
  51. Sabri, N.S.; Deni, M.S.M.; Zakaria, A.; Talari, M.K. Effect of Mn Doping on Structural and Optical Properties of SnO2 Nanoparticles Prepared by Mechanochemical Processing. Phys. Procedia 2012, 25, 233–239. [Google Scholar] [CrossRef] [Green Version]
  52. Tserkovnikov, Y.A. Decoupling of chains of equations for two-time Green’s functions. Teor. Mat. Fiz. 1971, 7, 250. [Google Scholar] [CrossRef]
  53. Scarrozza, M.; Vindigni, A.; Barone, P.; Sessoli, R.; Picozzi, S. A combined first-principles and thermodynamic approach to M-Nitronyl Nitroxide (M=Co, Mn) spin helices. Phys. Rev. B 2015, 91, 144422. [Google Scholar] [CrossRef]
  54. Cheng, J.; Zhou, J.; Li, J.; Dong, P.; Wu, Z. Magnetism in Cu-Doped Rutile SnO2 Semiconductor Induced by the RKKY Interaction. J. Supercond. Nov. Mat. 2014, 27, 581. [Google Scholar] [CrossRef]
  55. Adhikari, R.; Das, A.K.; Karmakar, D.; Rao, T.V.C.; Ghatak, J. Structure and magnetism of Fe-doped SnO2 nanoparticles. Phys. Rev. B 2008, 78, 024404. [Google Scholar] [CrossRef]
  56. Peters, L.; Ghosh, S.; Sanyal, B.; van Dijk, C.; Bowlan, J.; de Heer, W.; Delin, A.; Marco, I.D.; Eriksson, O.; Katsnelson, M.I.; et al. Magnetism and exchange interaction of small rare-earth clusters; Tb as a representative. Sci. Rep. 2016, 6, 19676. [Google Scholar] [CrossRef] [Green Version]
  57. Maeda, A.; Sugimoto, H. Intramolecular exchange interaction between two rare-earth ions, RE = Sm3+-Yb3+ and Y3+. J. Chem. Soc. Faraday Trans. 1986, 82, 2019–2030. [Google Scholar] [CrossRef]
  58. Santara, B.; Giri, P.K.; Imakita, K.; Fujii, M. Microscopic origin of lattice contraction and expansion in undoped rutile TiO2 nanostructures. J. Phys. D 2014, 47, 215302. [Google Scholar] [CrossRef] [Green Version]
  59. Kalathil, S.; Khan, M.M.; Ansari, S.A.; Lee, J.; Cho, M.H. Band gap narrowing of titanium dioxide (TiO2) nanocrystals by electrochemically active biofilms and their visible light activity. Nanoscale 2013, 5, 6323–6326. [Google Scholar] [CrossRef]
  60. Morales-Garcia, A.; Escatllar, A.M.; Illas, F.; Bromley, S.T. Understanding the interplay between size, morphology and energy gap in photoactive TiO2 nanoparticles. Nanoscale 2019, 11, 9032–9041. [Google Scholar] [CrossRef]
  61. Dette, C.; Perez-Osorio, M.A.; Kley, C.S.; Punke, P.; Patrick, C.E.; Jacobson, P.; Giustino, F.; Jung, S.J.; Kern, K. TiO2 anatase with a bandgap in the visible region. Nano Lett. 2014, 14, 6533–6538. [Google Scholar] [CrossRef]
  62. Tatar, B.; Sam, E.D.; Kutlu, K.; Uergen, M. Synthesis and optical properties of CeO2 nanocrystalline films grown by pulsed electron beam deposition. J. Mater. Sci. 2008, 43, 5102–5108. [Google Scholar] [CrossRef]
  63. Tamizhdurai, P.; Sakthinathan, S.; Chen, S.-M.; Shanthi, K.; Sivasanker, S.; Sangeetha, P. Environmentally friendly synthesis of CeO2 nanoparticles for the catalytic oxidation of benzyl alcohol to benzaldehyde and selective detection of nitrite. Sci. Rep. 2017, 7, 46372. [Google Scholar] [CrossRef]
  64. Kamarulzaman, N.; Aziz, N.D.A.; Kasim, M.F.; Chayed, N.F.; Subban, R.H.Y.; Badar, N. Anomalies in wide band gap SnO2 nanostructures. J. Solid State Chem. 2019, 277, 271–280. [Google Scholar] [CrossRef]
  65. Asaithambi, S.; Murugan, R.; Sakthivel, P.; Karuppaiah, M.; Rajendran, S.; Ravi, G. Influence of Ni Doping in SnO2 Nanoparticles with Enhanced Visible Light Photocatalytic Activity for Degradation of Methylene Blue Dye. J. Nanosci. Nanotechn. 2019, 19, 1–9. [Google Scholar] [CrossRef]
  66. Fujisawa, J.; Eda, T.; Hanaya, M. Comparative study of conduction-band and valence-band edges of TiO2, SrTiO3, and BaTiO3 by ionization potential measurements. Chem. Phys. Lett. 2017, 685, 23–26. [Google Scholar] [CrossRef]
  67. Dorenbos, P. The Electronic Structure of Lanthanide Impurities in TiO2, ZnO, SnO2, and Related Compounds. ECS J. Solid State Sci. Techn. 2014, 3, R19–R24. [Google Scholar] [CrossRef] [Green Version]
  68. Miyauchi, M.; Nakajima, A.; Watanabe, T.; Hashimoto, K. Photocatalysis and Photoinduced Hydrophilicity of Various Metal Oxide Thin Films. Chem. Mater. 2002, 14, 2812. [Google Scholar] [CrossRef]
  69. Kikoin, K.; Fleurov, V. On the nature of ferromagnetism in non-stoichiometric TiO2 doped with transition metals. J. Magn. Magn. Mater. 2007, 310, 2097–2098. [Google Scholar] [CrossRef]
  70. Wei, L.; Yang, Y.; Xia, X.; Fan, R.; Su, T.; Shi, Y.; Yu, J.; Lia, L.; Jiang, Y. Band edge movement in dye sensitized Sm-doped TiO2 solar cells: A study by variable temperature spectroelectrochemistry. RSC Adv. 2015, 5, 70512–70521. [Google Scholar] [CrossRef]
  71. Zheng, X.; Li, Y.; You, W.; Lei, G.; Cao, Y.; Zhang, Y.; Jiang, L. Construction of Fe-doped TiO2-x ultrathin nanosheets with rich oxygen vacancies for highly efficient oxidation of H2S. Chem. Eng. J. 2022, 430 Pt 2, 132917. [Google Scholar] [CrossRef]
  72. Wu, H.-C.; Li, S.-H.; Lin, S.-W. Effect of Fe Concentration on Fe-Doped Anatase TiO2 from GGA+U Calculations. Intern. J. Photoen. 2012, 2012, 823498. [Google Scholar] [CrossRef] [Green Version]
  73. Akshay, V.R.; Arun, B.; Dash, S.; Patra, A.K.; Mandal, G.; Mutta, G.R.; Chanda, A.; Vasundhara, M. Defect mediated mechanism in undoped, Cu and Zn-doped TiO2 nanocrystals for tailoring the band gap and magnetic properties. RSC Adv. 2018, 8, 41994. [Google Scholar] [CrossRef] [PubMed]
  74. Mathew, S.; Ganguly, P.; Rhatigan, S.; Kumaravel, V.; Byrne, C.; Hinder, S.J.; Bartlett, J.; Nolan, M.; Pillai, S.C. Cu-Doped TiO2: Visible Light Assisted Photocatalytic Antimicrobial Activity. Appl. Sci. 2018, 8, 2067. [Google Scholar] [CrossRef] [Green Version]
  75. Khlyustova, A.; Sirotkin, N.; Kusova, T.; Kraev, A.; Titov, V.; Agafonov, A. Doped TiO2: The effect of doping elements on photocatalytic activity. Mater. Adv. 2020, 1, 1193. [Google Scholar] [CrossRef]
  76. Soni, S.; Vats, V.S.; Kumar, S.; Dalela, B.; Mishra, M.; Meena, R.S.; Gupta, G.; Alvi, P.A.; Dalela, S. Structural, optical and magnetic properties of Fe-doped CeO2 samples probed using X-ray photoelectron spectroscopy. J. Mater. Sci. Mater. Electr. 2018, 29, 10141–10153. [Google Scholar] [CrossRef]
  77. Xue, M.; Liu, X.; Wu, H.; Yu, B.; Meng, F. Construction of Cu2+-doped CeO2 nanocrystals hierarchical hollow structure and its enhanced photocatalytic performance. J. Mater. Sci. Mater. Electr. 2021, 32, 27576–27586. [Google Scholar] [CrossRef]
  78. Ranjith, K.S.; Dong, C.-L.; Lu, Y.-R.; Huang, Y.-C.; Chen, C.-L.; Saravanan, P.; Asokan, K.; Kumar, R.T.R. Evolution of Visible Photocatalytic Properties of Cu-Doped CeO2 Nanoparticles: Role of Cu2+-Mediated Oxygen Vacancies and the Mixed-Valence States of Ce Ions. ACS Sust. Chem. Eng. 2018, 6, 8536–8546. [Google Scholar] [CrossRef]
  79. Arif, M.; Shah, M.Z.U.; Ahmad, S.A.; Shah, M.S.; Ali, Z.; Ullah, A.; Idrees, M.; Zeb, J.; Song, P.; Huang, T.; et al. High photocatalytic performance of copper-doped SnO2 nanoparticles in degradation of Rhodamine B dye. Optical Mater. A 2022, 134, 113135. [Google Scholar] [CrossRef]
  80. Hanaor, D.A.H.; Sorrell, C.C. Review of the anatase to rutile phase transformation. J. Mater. Sci. 2011, 46, 855. [Google Scholar] [CrossRef] [Green Version]
  81. Chatterjee, P.; Mukherjee, D.; Sarkar, A.; Chakraborty, A.K. Mn-doped CeO2-CNT nanohybrid for removal of water soluble organic dyes. Appl. Nanosci. 2022, 12, 3031–3043. [Google Scholar] [CrossRef]
  82. Mohanta, D.; Barman, K.; Jasimuddin, S.; Ahmaruzzaman, M. MnO doped SnO2 nanocatalysts: Activation of wide band gap semiconducting nanomaterials towards visible light induced photoelectrocatalytic water oxidation. J. Colloid. Interface Sci. 2017, 505, 756–762. [Google Scholar] [CrossRef] [PubMed]
  83. Azam, A.; Ahmed, A.S.; Habib, S.S.; Naqvi, A.H. Effect of Mn doping on the structural and optical properties of SnO2 nanoparticles. J. Alloy. Compd. 2012, 523, 83–87. [Google Scholar] [CrossRef]
  84. Dimri, M.C.; Khanduri, H.; Kooskora, H.; Subbi, J.; Heinmaa, I.; Mere, A.; Krustok, J.; Stern, R. Ferromagnetism in rare earth doped cerium oxide bulk samples. Phys. Status Solidi A 2012, 209, 353. [Google Scholar] [CrossRef]
  85. Dinkar, V.A.; Shridhar, S.J.; Madhukar, E.N.; Anil, E.A.; Nitin, H.K. Sm-Doped TiO2 Nanoparticles with High Photocatalytic Activity for ARS Dye Under Visible Light Synthesized by Ultrasonic Assisted Sol-Gel Method. Orient J. Chem. 2016, 32, 933. [Google Scholar] [CrossRef] [Green Version]
  86. Mandal, B.; Mondal, A.; Ray, S.S.; Kundu, A. Sm-doped mesoporous CeO2 nanocrystals: Aqueous solution-based surfactant assisted low temperature synthesis, characterization and their improved autocatalytic activity. Dalton Trans. 2016, 45, 1679–1692. [Google Scholar] [CrossRef] [PubMed]
  87. Xia, X.; Li, J.-Q.; Chen, C.-Y.; Lan, Y.; Mao, X.; Bai, F. Optimal RE (RE = La, Y and Sm) doping conditions and the enhanced mechanism for photocatalytic application of ceria nanorods. Nanotechn. 2021, 32, 195708. [Google Scholar] [CrossRef]
  88. Xia, X.; Li, J.; Chen, C.; Lan, Y.-P.; Mao, X.; Chu, Z.; Ning, D.; Zhang, J.; Liu, F. Collaborative influence of morphology tuning and RE (La, Y, and Sm) doping on photocatalytic performance of nanoceria. Environ. Sci. Pollut. Res. Int. 2022, 29, 88866–88881. [Google Scholar] [CrossRef]
  89. Lee, D.Y.; Kim, J.-T.; Park, J.-H.; Kim, Y.-H.; Lee, I.-K.; Lee, M.-H.; Kim, B.-Y. Effect of Er doping on optical band-gap energy of TiO2 thin films prepared by spin coating. Current. Appl. Phys. 2013, 13, 1301–1305. [Google Scholar] [CrossRef]
  90. Butt, K.M.; Farrukh, M.A.; Muneer, I. Influence of lanthanum doping via hydrothermal and reflux methods on the SnO2-TiO2 nanoparticles prepared by sol-gel method and their catalytic properties. J. Mater. Sci. Mater. Electron. 2016, 27, 8493–8498. [Google Scholar] [CrossRef]
  91. Nabi, G.; Ali, W.; Majid, A.; Alharbi, T.; Saeed, S.; Albedah, M.A. Controlled growth of Bi-Functional La-doped CeO2 nanorods for photocatalytic H2 production and supercapacitor applications. Int. J. Hydrogen Energy 2022, 47, 15480–15490. [Google Scholar] [CrossRef]
  92. Tian, A.; Ma, T.; Shi, X.; Wang, D.; Wu, W.; Liu, C.; Pei, W. Synergistic Improvement in Coating with UV Aging Resistance and Anti-Corrosion via La-Doped CeO2 Powders. Coatings 2021, 11, 1095. [Google Scholar] [CrossRef]
  93. Fan, Y.; Jiao, J.; Zhao, L.; Tang, J. Preparation of Lanthanide-doped (La, Nd, and Eu) polystyrene/CeO2 abrasives and investigation of slurry stability and photochemical mechanical polishing performance. Colloids Surf. A Physicochem. Eng. Asp. 2023, 656, 130508. [Google Scholar] [CrossRef]
  94. Joseph, J.; Mathew, V.; Abraham, K.E. Physical properties of Dy and La-doped SnO2 thin films prepared by a cost effective vapour deposition technique. Cryst. Res. Techn. 2006, 41, 1020–1026. [Google Scholar] [CrossRef]
  95. Babu, K.R.V.; Renuka, C.G.; Zikriya, M. Band Gap and Photo luminescence studies of Sm3+ doped TiO2 NP for wLED. Int. Adv. Res. J. Sci. Eng. Techn. 2021, 8, 24. [Google Scholar] [CrossRef]
  96. Bharathi, R.N.; Sankar, S. Structural, optical and magnetic properties of Pr doped CeO2 nanoparticles synthesized by citrate–nitrate auto combustion method. J. Mater. Sci. Mater. Electr. 2018, 29, 6679–6691. [Google Scholar] [CrossRef]
Figure 1. Size dependence of the band-gap energy E g for J s = 0.8 J b , T = 300 K for a (1) TiO 2 , (2) CeO 2 , and (3) SnO 2 NP.
Figure 1. Size dependence of the band-gap energy E g for J s = 0.8 J b , T = 300 K for a (1) TiO 2 , (2) CeO 2 , and (3) SnO 2 NP.
Nanomaterials 13 00145 g001
Figure 2. Dependence of the (1,3) VB edge energy of Fe-doped and (2,4) of the CB edge energy of Sm-doped bulk TiO 2 and SnO 2 , respectively, on the ion-doping concentration x.
Figure 2. Dependence of the (1,3) VB edge energy of Fe-doped and (2,4) of the CB edge energy of Sm-doped bulk TiO 2 and SnO 2 , respectively, on the ion-doping concentration x.
Nanomaterials 13 00145 g002
Figure 3. Dependence of the band-gap energy E g for T = 300 K on the ion-doping concentration x in a TiO 2 NP (N = 20 shells) for (1) Co, (2) Fe, (3) Mn, and (4) Cu.
Figure 3. Dependence of the band-gap energy E g for T = 300 K on the ion-doping concentration x in a TiO 2 NP (N = 20 shells) for (1) Co, (2) Fe, (3) Mn, and (4) Cu.
Nanomaterials 13 00145 g003
Figure 4. Dependence of the band-gap energy E g for T = 300 K on the ion-doping concentration x in a TiO 2 NP (N = 20 shells) for (1) Sm, (2) Tb, and (3) Er.
Figure 4. Dependence of the band-gap energy E g for T = 300 K on the ion-doping concentration x in a TiO 2 NP (N = 20 shells) for (1) Sm, (2) Tb, and (3) Er.
Nanomaterials 13 00145 g004
Figure 5. Dependence of the magnetization M on the ion-doping concentration (1) Fe and (2) Sm for J s = 1.2 J b , T = 300 K, for bulk TiO 2 .
Figure 5. Dependence of the magnetization M on the ion-doping concentration (1) Fe and (2) Sm for J s = 1.2 J b , T = 300 K, for bulk TiO 2 .
Nanomaterials 13 00145 g005
Figure 6. Dependence of the band-gap energy E g on the doping concentration x in bulk TiO 2 : (1) Co, (2) Fe, and (3) Sm, for J s = 1.2 J b , T = 300 K.
Figure 6. Dependence of the band-gap energy E g on the doping concentration x in bulk TiO 2 : (1) Co, (2) Fe, and (3) Sm, for J s = 1.2 J b , T = 300 K.
Nanomaterials 13 00145 g006
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Apostolova, I.; Apostolov, A.; Wesselinowa, J. Band Gap Tuning in Transition Metal and Rare-Earth-Ion-Doped TiO2, CeO2, and SnO2 Nanoparticles. Nanomaterials 2023, 13, 145. https://doi.org/10.3390/nano13010145

AMA Style

Apostolova I, Apostolov A, Wesselinowa J. Band Gap Tuning in Transition Metal and Rare-Earth-Ion-Doped TiO2, CeO2, and SnO2 Nanoparticles. Nanomaterials. 2023; 13(1):145. https://doi.org/10.3390/nano13010145

Chicago/Turabian Style

Apostolova, Iliana, Angel Apostolov, and Julia Wesselinowa. 2023. "Band Gap Tuning in Transition Metal and Rare-Earth-Ion-Doped TiO2, CeO2, and SnO2 Nanoparticles" Nanomaterials 13, no. 1: 145. https://doi.org/10.3390/nano13010145

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop