The magnetic and optical properties of 3d transition metal doped SnO2 nanosheets

Yong Feng, Wei-Xiao Ji, Bao-Jun Huang, Xin-lian Chen, Feng Li, Ping Li, Chang-wen Zhang and Pei-Ji Wang*
School of Physics, University of Jinan, Jinan 250022, People’s Republic of China. E-mail: ss_wangpj@ujn.edu.cn; Tel: +86 531 82765965

Received 14th January 2015 , Accepted 25th February 2015

First published on 25th February 2015


Abstract

Based on first-principles calculations, we study the electronic structure, magnetic properties and optical properties of transition metal (TM) doped SnO2NSs. Computational results indicate that pristine SnO2NSs is a direct gap semiconductor with nonmagnetic states. Cr, Mn, Fe atom doping can induce 2μB, −3μB and 2μB magnetic moment, respectively, while Ni atom doped SnO2NSs keeps the nonmagnetic states. More interestingly, Fe doped SnO2NSs becomes an indirect gap semiconductor, and the Cr, Mn and Ni atom doping maintain the character of direct gap semiconductor. For optical properties, the optical absorption edge shows red shift phenomenon for a TM atom (Cr, Mn, Fe or Ni) doped SnO2NSs. In addition, the tensity of absorption, reflection and refraction coefficient are enhanced significantly in the visible light region, which may be very useful for the design of solar cells, photoelectronic devices and photocatalysts.


1. Introduction

Dilute magnetic semiconductors (DMSs) have attracted extensive attention in experimental and theoretical studies because of their potential application in spintronic devices.1–5 For practical applications, it is required that DMSs are half-metallic (HM) or ferromagnetic (FM) around room temperature. However, DMSs usually have a rather low Curie temperature (TC), so it is desirable find a new method to improve the TC of DMSs. Recently, the reported DMSs can obtain high TC by doping with transition metal (TM) elements, such as ZnO,6–8 In2O3,9 TiO2,10 CeO2 (ref. 11) and so on.

As a typical wide-gap semiconductor with band gap Eg = 3.6 eV, SnO2 has received considerable interest for its wide variety of practical applications, such as photovoltaic devices, transparent conducting electrodes, gas sensors, solar cells, panel displays and ferroelectric transparent thin-film transistors.12–17 More recently, TM-doped SnO2 has become one of the hottest research topics in spintronics. Room temperature (RT) ferromagnetism has been observed in V,18 Cr,19,20 Mn,21 Fe,22,23 Co24 and Ni25 doped SnO2. Research in this field now mainly concentrates on SnO2 bulk while seldom reports on SnO2 nanosheets theoretical calculations. The nanosheets structure may have many excellent properties because of its larger specific surface area. SnO2 nanosheets have higher lithium storage properties with high reversible capacities and good cycling performance due to a large specific surface area.26 Compared to SnO2 bulk, SnO2NSs with larger band gap are more suitable for tuning band gap to obtain better optical properties. And high optical transmittance,27 good gas-sensing optical properties are found in transition metal doped SnO2 film. More interestingly, SnO2 nanosheets (NSs) were by fabricated Wang et al.28 in experiments, and the TC was estimated to be about 300 °C, indicated its potential value in spintronic devices. The results provide an opportunity to study two-dimensional (2D) SnO2NSs. Gul Rahman et al.29 reported intrinsic magnetism in nanosheets of SnO2 and found that it is easier to create vacancies, which are magnetic, at the surface of the sheets for SnO2NSs different thicknesses. Research by Luan et al.30 found a strong localized magnetic moment in Co-doped SnO2NSs. For optical properties, Huang31 reported that the capacity to absorb becomes stronger in the visible region for SnO2NSs doped with increasing Ag concentrations.

In the present work, we perform first principles calculations to study the electronic, magnetic and optical properties of TM element doped SnO2NSs, where the TM atom substitutes Sn atom for Cr, Mn, Fe or Ni in 2D SnO2NSs. Our results show that the doped TM atom could induce magnetic moment, except Ni, and the magnetic moment mainly comes from the 3d orbital of the TM atom. The optical properties can get significantly improved in the visible region.

2. Computational details

The electronic properties and band structures were performed employing the vienna ab initio simulation package (VASP),32 and the optical properties were carried out by using the WIEN2k simulation package which implements a full potential linearized augmented plane wave (FLAPW) method.33 The exchange–correlation (XC) functional was approximated with a generalized gradient approximation (GGA)34 and GGA + U35 as proposed by the Perdew–Bueke–Emzerhof (PBE). And a 400 eV cutoff energy for the plane-wave basis set was used. Two-dimensional periodic boundary conditions were applied to the TM-doped SnO2NSs while a vacuum region of 12 Å is set along the direction perpendicular to the SnO2NSs surface to avoid mirror interaction. The k-point meshes for Brillouin zone were generated using a 9 × 9 × 1 Monkhorst–Pack grid.36 All atomic positions in all structures were fully relaxed until all atomic forces converged to be less than 0.02 eV Å−1.

3. Results and discussion

3.1 Formation energy

Firstly, we used pristine SnO2NSs, which is modeled with a 4 × 4 × 1 supercell containing sixteen SnO2 units, as shown in Fig. 1. We found that it forms a sandwiched structure as seen from the side view of SnO2NSs Fig. 1(b). These structures are similar to MoS2NSs37 and TiO2NSs38 in previous reports. In the top view of SnO2NSs Fig. 1(a), the tin atom is substituted by TM atom at the site of X. The optimized lattice parameters are 3.2 Å and the length of Sn–O bond is 2.11 Å, which is consistent with Luan’s30 research. In Fig. 1(c), we show the total density of states (TDOSs) for pristine SnO2NSs. It presents nonmagnetic behavior due to the electron wave function dose not exhibiting spin polarization in spin-up and spin-down channels. The band structure of pristine SnO2NSs is as shown in Fig. 1(d). The band gap is about 2.75 eV, which is close to the experimental value in Huang’s31 report. Thus, pristine SnO2NSs is also a direct gap semiconductor, which is in agreement with the bulk SnO2 calculated results.39,40
image file: c5ra00788g-f1.tif
Fig. 1 (a) Top view of pure SnO2NSs (the position of Sn is denoted by X) (b) side view of pure SnO2NSs (c) TDOS of pure SnO2NSs (d) the band structure of pure SnO2NSs.

Next, we investigated the effect of TM atom (Cr, Mn, Fe or Ni) substitution for Sn atom in SnO2NS. The doping atom is located at X shown in Fig. 1(a). In order to evaluate the stability of TM-doped SnO2NSs, we calculated the defect formation energy (Ef) defined as Ef = EX:SnO2ESnO2 + uSnuX,41 where EX:SnO2 and ESnO2 are the total energies of the TM-doped SnO2NSs and SnO2NSs supercell, and uSn, uX are the chemical potentials of Sn and TM atom (Cr, Fe, Mn or Ni), respectively. The values of Ef are shown in Table 1. Compared with pristine SnO2NSs, the Cr, Mn doped SnO2NSs release 0.56 eV, 0.37 eV in energy and Fe, Ni doped SnO2NSs absorb 0.35 eV, 1.75 eV in energy, respectively. This indicates that Cr, Mn as a dopant can obtain a more stable structure than pristine SnO2NSs. After structural optimization, we found that all the bond lengths of TM–O bond (dX–O) were shorter than the Sn–O bond (2.11 Å), see Table 1, because of the smaller ionic radius of the TM atom (Cr, Fe, Mn or Ni) which substitutes the Sn atom. And the bond length of TM–O atom gets shorter with an increase of atomic number.

Table 1 The bond length between the sit of X atom and its nearest O atoms (dX–O), the defect formation energy (Ef) whose plus and minus represent the absorbing and releasing energy, the total magnetic moment (M), the total magnetic moment with GGA + U method (Mu) and the local magnetic moment for TM atom (MX) whose plus and minus represent spin-up an spin-down, the charge transfer for TM atom (Q) whose minus represent losing electron
Configurations dX–O Ef M Mu MX Q
SnO2 2.11 0
Cr[thin space (1/6-em)]:[thin space (1/6-em)]SnO2 1.97 −0.56 2 2 1.97 −1.80
Mn[thin space (1/6-em)]:[thin space (1/6-em)]SnO2 1.96 −0.37 −3 −3 −2.85 −1.73
Fe[thin space (1/6-em)]:[thin space (1/6-em)]SnO2 1.95 +0.35 2 2 1.91 −1.57
Ni[thin space (1/6-em)]:[thin space (1/6-em)]SnO2 1.95 +1.75 0 0 0 −1.34


3.2 Electronic structures and magnetic properties

The TDOSs and partial density of states (PDOSs) for the TM atom (Cr, Fe, Mn or Ni) doped SnO2NSs are illustrated in Fig. 2(a)–(d), respectively. Taking Cr-doped SnO2NSs as a representative example from Fig. 2(a), the electronic states of the valence band mainly arise from the strong hybridization between the O 2p and Sn 5p states, while the Sn 5s states make mainly a contribution to the electronic states of the conduction band. We also observe that the Fermi level shifts upward to the conduction band, indicating an n-type conductivity character, which may display a low resistivity character. Moreover, there are some impurity energy levels near the fermi level which are mainly caused by Cr 3d, O 2p and Sn 5p states. The Fe-, Mn-, Ni-doped SnO2NSs can give the similar conclusion by analyzing the DOS. However, compared with three other TM atom (Cr, Fe, Mn) doped SnO2NSs, the Ni-doped SnO2NSs is different due to a symmetrical distribution of the wave functions of spin-up and spin-down channels on TDOSs. The result indicates that the other doped structures are magnetic except Ni-doped SnO2NSs. DOS analysis further shows that the magnetic moment is mainly from TM atom 3d orbital.
image file: c5ra00788g-f2.tif
Fig. 2 Total and partial DOS for (a) Cr, (b) Mn, (c) Fe, and (d) Ni atom and nearest O, Sn atoms. Fermi level is set to zero. DOS is broadened by Gaussian smearing with 0.2 eV.

In order to study the origin of the magnetism, we performed spin-polarized calculations and the spin density distributions are displayed in Fig. 3. We can find a strong localized magnetic moment except for Ni-doped SnO2NSs and the magnetism mainly results from the doped atom. The Cr, Fe atom supplies the spin-up magnetic moment and Mn atom provides the spin-down magnetic moment, while a Ni atom does not provide the magnetic moment. To further verify the localization of the magnetic moment, we calculated the net magnetization defined as m(r) = ρ(r) − ρ(r), where ρ(r)ρ(r) is the charge density in spin-up and spin-down channel. The corresponding values of magnetic moment are shown in Table 1. The total magnetic moment (M) of Cr, Mn, Fe and Ni doped SnO2NSs are found to be 2μB, −3μB, 2μB and 0μB, for which the Cr, Mn, Fe atom provides 1.97μB, −2.85μB, 1.91μB local magnetic moment (MX), respectively. The results are similar to the case of TM atom doped TiO2 carried out by Errico et al.42 Each doped TM atom contributes two valence electrons to the nearest O atom, and the unpaired electrons in Cr, Mn, Fe atoms induce the local magnetic moment. The mechanism can be understood as follow: a Cr atom loses two 3d-electrons and one electron of the 4s electrons transitions to the 3d orbit to form paired electrons. Two unpaired electrons provide 1.97μB spin-up local magnetic moment for Cr atom. However, the Mn and Fe atoms lose two 3d-electrons forming three unpaired spin-down electrons and two unpaired spin-up electrons, respectively. The magnetic moment can be widely applied to spintronic devices. The charge transfer (Q) for TM atom can be obtained from Bader analysis in Table 1. We found that the number electrons lost in Cr, Mn, Fe and Ni atoms is 1.80, 1.73 1.57 and 1.34, respectively, with the number of electron less than 2 due to the TM atom and O atom forming covalent bonds. We also found that the number of electrons lost tends to drop off with an increase of atomic number. The reason is that the electronegativity of atom is gradually strengthened.


image file: c5ra00788g-f3.tif
Fig. 3 The spin density distribution of TM-doped SnO2NSs. (a) Cr-doped SnO2NSs (b) Mn-doped SnO2NSs (c) Fe-doped SnO2NSs (d) Ni-doped SnO2NSs, yellow and blue represent spin-up and spin-down.

Further, the band structures of TM atom (Cr, Mn, Fe or Ni) doped SnO2NSs are illustrated in Fig. 4. In Fig. 4, the arrows ↑ and ↓ represent spin-up and spin-down, respectively. Around the Fermi level, some rather localized energy bands emerged, which mainly originate from 3d states of TM atom. The emergence of impurity energy levels may make electronic transition more active from the occupied bands to the unoccupied. This may be helpful to improve the optical properties. For the case of Cr, Mn, and Ni doping, the band structure still remains the character of direct gap semiconductor, while Fe doped SnO2NSs become an indirect gap semiconductor. We can obtain the same conclusion with DOS from band structures: the Cr, Mn, and Fe doped SnO2NSs can introduce magnetic moment while the Ni atom can not.


image file: c5ra00788g-f4.tif
Fig. 4 Electronic band structures for (a) Cr, (b) Mn, (c) Fe and (d) Ni adsorbed stanene. Fermi level is set to zero. The arrows ↑ and ↓ represent spin-up and spin-down, respectively.

Unfortunately, the GGA method usually underestimates the binding energy of d states and yields incorrect behavior for strongly correlated magnetic systems. In order to check our results, we include an on-site Coulomb correlation interaction as given by the GGA + U method, where U (U = 3.0 eV) are used to calculate the TM atom 3d orbit. This is essential for obtaining accurate values for band structures and band gaps.43 The Sn atom 4d orbit is far from the Fermi level and has little impact on the magnetic properties. In Fig. 5, we show the total DOS and partial DOS with GGA + U method. Compared with the GGA results, the Fermi energy level shifts down to the valence band, and the amount of impurity energy level is declined around the Fermi energy level. More interestingly, the band gap using the GGA + U method is larger than when using the GGA method, and shows that the GGA underestimates that the TM atom 3d orbitals hybridize quite strongly with oxygen 2p orbitals. The magnetic moments of GGA + U are listed in Table 1. We found that the GGA + U method obtains the same results for magnetism, which indicates that the results of the GGA method can qualitatively reflect the properties of TM atom doped SnO2NS. The band gap slightly increased and the magnetic moments remained unchanged, which demonstrates that the results of GGA method are reliable.


image file: c5ra00788g-f5.tif
Fig. 5 Total and TM atom 3d partial DOS with U = 3 (a) Cr, (b) Mn, (c) Fe, and (d) Ni. Fermi level is set to zero. DOS is broadened by Gaussian smearing with 0.2 eV.

3.3 Optical properties

The optical properties of TM atom doped SnO2NSs have also been calculated in our work. The optical properties of the medium can be derived from the complex dielectric function ε(ω) = ε1(ω) + iε2(ω).44 The real part of the dielectric function ε1(ω) can be evaluated from the imaginary part via the Kramers–Kronig transform, and the imaginary part of the dielectric function ε2(ω) momentum matrix elements between the occupied and unoccupied electronic states. The imaginary part of the dielectric function formulas is calculated as follows,45
image file: c5ra00788g-t1.tif
where subscripts C and V represent conduction band and valence band, respectively, BZ is the first Brillouin zone, k is reciprocal lattice vector, ω is angular frequency, EV(k) and EC(k) are intrinsic energy level of valence band and conduction band, respectively, |MCV(K)|2 is momentum matrix element.

The imaginary part spectrum of the dielectric function for pristine and TM atom doped SnO2NSs are shown in Fig. 6(a). The intensity of ε2(ω) for TM atom doping is stronger than for pure SnO2 nanosheets especially in the low energy region. The reason is that some impurity energy levels emerge around the Fermi level and the probability of electron transition becomes more common when the Sn atom is replaced by a TM atom. For Fe doped SnO2NSs, there are two obvious peaks located at 0.2 eV and 0.6 eV. These peaks are mainly ascribed to the transition between Fe 3d states and Sn 5s, 5p states. There is one obvious peak at 0.5 eV for Cr doped SnO2NSs. The reason is thought to be the band transition from Cr 4s orbital to Cr 3d orbital. There are some other peaks, and the tensity of dielectric function is larger in the visible light region for all TM doping models than for pure SnO2NSs, which indicates that the TM doped SnO2NSs are more suitable for visible photoelectric devices compared with pure SnO2NSs. In the ultraviolet region, the trend and peaks of TM atom doping are similar to pure SnO2NSs. As we all know, the absorption property of a semiconductor is related to its electronic band structure. The absorption coefficient can be defined as follow: image file: c5ra00788g-t2.tif In Fig. 6(b), we can see that the capacity of absorption is improved in the low energy region, which demonstrates that the TM atom doped SnO2NSs is more suitable making a photodetector than pure SnO2NSs. And some peaks emerge in the visible light region which are in agreement with the imaginary part spectrum of dielectric function. These peaks are ascribed to the interband transition between the TM 3d states and O 2p states. Interestingly, the absorption edges of TM doped SnO2NSs all show red shift phenomenon, which are caused by the impurity energy levels near the Fermi level. The absorption edges of Cr, Fe or Ni doped SnO2NSs are below 1.50 eV, which is lower than Mn doped SnO2NSs whose absorption edge is located at around 2.20 eV. These results indicate that Cr, Fe or Ni doped SnO2NSs have better optical properties compared with Mn doped SnO2NSs in the visible region. But the all doped models have a wider absorption region than pure SnO2NSs. These results show that the TM doped SnO2NSs can be widely used in infrared and visible photodetectors.


image file: c5ra00788g-f6.tif
Fig. 6 (a) The dielectric functions of pure SnO2NSs and TM atom doped SnO2NSs (b) optical absorption coefficient of pure SnO2NSs and TM atom doped SnO2NSs.

The reflectivity and refractivity of pure SnO2NSs and TM atom doped SnO2NSs are calculated and shown in Fig. 7. The reflectivity and refractivity enhance much in the low energy region, which also mainly originates from the impurity energy levels near the Fermi level. At the Fermi energy level, the values of refraction of pristine and Cr, Mn, Fe, Ni doped SnO2NSs are 1.181, 1.232, 1.192, 1.247, 1.203, respectively. The Fe doped SnO2NSs obtains the best effect while there is a minimum influence for Mn doped SnO2NSs. The same effect of TM doping is suitable for reflectivity. The location of reflectivity and refractivity peaks are also in agreement with the imaginary part spectrum of dielectric function. In conclusion, the optical properties of TM atom doped SnO2NSs can get significantly improved in the low energy region, while the effect is less obvious in the ultraviolet region.


image file: c5ra00788g-f7.tif
Fig. 7 (a) The reflectivity of pure SnO2NSs and TM atom doped SnO2NSs (b) refractivity of pure SnO2NSs and TM atom doped SnO2NSs.

4. Conclusion

In summary, we performed a first-principles study on the electronic structure, magnetic and optical properties of TM atom doped SnO2NSs. Computational results indicate that the pristine SnO2NSs is a direct gap semiconductor with nonmagnetic states. Doping with Cr, Mn, Fe can induce the magnetic moment with a value of 2μB, −3μB and 2μB, respectively, while the Ni atom doped SnO2NSs remains nonmagnetic. The magnetic moment results mainly from the TM atom 3d orbital. More interestingly, Cr, Mn and Ni atom doped SnO2NSs still maintains the character of a direct gap semiconductor except for Fe atom doping. For optical properties, the absorption edge shifts to the low energy region compared to pure SnO2NSs. More interestingly, the tensity of absorption, reflection and refraction coefficients are enhanced significantly in the visible light region. The results indicate that TM atom doped SnO2NSs may be very useful for the design of solar cells, photoelectronic devices and photocatalysts.

Acknowledgements

This work was supported by the National Natural Science Foundation of China (Grant no. 61172028, 11274143 and 11304121), the Natural Science Foundation of Shandong Province (Grant no. ZR2010EL017 and ZR2013AL004), the Research Fund for the Doctoral Program of University of Jinan (Grant no. XBS1433, XBS1402 and XBS1452).

References

  1. Y. W. Son, M. L. Cohen and S. G. Louie, Nature, 2006, 444, 347 CrossRef CAS PubMed.
  2. E. Kan, L. Yuan and J. L. Yang, J. Appl. Phys., 2007, 102, 033915 CrossRef PubMed.
  3. S. A. Wolf, D. D. Awschalom, R. A. Buhrman, J. M. Daughton, S. von Molnar, M. L. Roukes, A. Y. ChtchelkanVOa and D. M. Treger, Science, 2001, 294, 1488 CrossRef CAS PubMed.
  4. S. B. Ogale, R. J. Choudhary, J. P. Buban, S. E. Lofland, S. R. Shinde, S. N. Kale, V. N. Kulkarni, J. Higgins, C. Lanci, J. R. Simpson, N. D. Browning, S. Das Sarma, H. D. Drew, R. L. Greene and T. Venkatesan, Phys. Rev. Lett., 2003, 077205 CrossRef CAS.
  5. T. Fukumura, Z. Jin, A. Ohtomo, H. Koinuma and M. Kawasaki, Appl. Phys. Lett., 1999, 75, 3366 CrossRef CAS PubMed.
  6. Y. Liu, Y. Yang, J. Yang, Q. Guan, H. Liu, L. Yang, Y. Zhang, Y. Wang, M. Wei, X. Liu, L. Fei and X. Cheng, J. Solid State Chem., 2011, 184, 1273 CrossRef CAS PubMed.
  7. D. A. Schwartz, K. R. Kittilstved and D. R. Gamelin, Appl. Phys. Lett., 2004, 85, 1395 CrossRef CAS PubMed.
  8. J. H. Shim, T. Hwang, S. Lee, J. H. Park, S. J. Han and Y. H. Jeong, Appl. Phys. Lett., 2005, 86, 082503 CrossRef PubMed.
  9. N. H. Hong, J. Sakai, N. T. Huong and V. Brize, Appl. Phys. Lett., 2005, 87, 102505 CrossRef PubMed.
  10. N. H. Hong, J. Sakai and W. Prellier, Phys. Rev. B: Condens. Matter Mater. Phys., 2004, 70, 195204 CrossRef.
  11. Q. Y. Wen, H. W. Zhang, Y. Q. Song, Q. H. Yang, H. Zhu and J. Q. Xiao, J. Phys.: Condens. Matter, 2007, 19, 246205 CrossRef PubMed.
  12. A. Nuruddin and J. R. Abelson, Thin Solid Films, 2001, 394, 49–63 CrossRef CAS.
  13. Y. S. Zhang, K. Yu, G. D. Li, D. Y. Peng, Q. X. Zhang, H. M. Hu, F. Xu, W. Bai, S. X. Ouyang and Z. Q. Zhu, Appl. Surf. Sci., 2006, 253, 792–796 CrossRef CAS PubMed.
  14. S. Ferrere, A. Zaban and B. A. Gsegg, J. Phys. Chem. B, 1997, 101, 4490–4493 CrossRef CAS.
  15. M. Khan, J. N. Xu, N. Chen and W. B. Cao, J. Alloys Compd., 2012, 513, 539–545 CrossRef CAS PubMed.
  16. H. C. Wu, Y. C. Peng and C. C. Chen, Opt. Mater., 2013, 35, 509–515 CrossRef CAS PubMed.
  17. Z. Y. Zhang, C. L. Shao, X. H. Li, L. Zhang, H. M. Xue, C. H. Wang and Y. C. Liu, J. Phys. Chem. C, 2010, 114, 7920 CAS.
  18. J. Zhang, R. Skomski, L. P. Yue, Y. F. Lu and D. J. Sellmyer, J. Phys.: Condens. Matter, 2007, 19, 256204 CrossRef.
  19. C. B. Fitzgerald, M. Venkatesan, L. S. Dorneles, R. Gunning, P. Stamenov, J. M. D. Coey, P. A. Stampe, R. J. Kennedy, E. C. Moreira and U. S. Sias, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 74, 115307 CrossRef.
  20. N. H. Hong, J. Sakai, W. Prellier and A. Hassini, J. Phys.: Condens. Matter, 2005, 17, 1697 CrossRef CAS.
  21. C. B. Fitzgerald, M. Venkatesan, L. S. Dorneles, R. Gunning, P. Stamenov, J. M. D. Coey, P. A. Stampe, R. J. Kennedy, E. C. Moreira and U. S. Sias, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 74, 115307 CrossRef.
  22. J. M. D. Coey, A. P. Douvalis, C. B. Fitzgerald and M. Venkatesan, Appl. Phys. Lett., 2004, 84, 1332 CrossRef CAS PubMed.
  23. A. Punnoosea, K. M. Reddu, J. Hays, A. Thurber and M. H. Engelhard, Appl. Phys. Lett., 2006, 89, 112509 CrossRef PubMed.
  24. S. K. Misra, S. I. Andronenko, K. M. Reddy, J. Hays and A. Punnoose, J. Appl. Phys., 2006, 99, 08M106 CrossRef PubMed.
  25. N. H. Hong, J. Sakai, N. T. Huong, N. Poirot and A. Ruyter, Phys. Rev. B: Condens. Matter Mater. Phys., 2005, 72, 045336 CrossRef.
  26. S. J. Ding, D. Y. Luan, F. Y. C. Boey, J. S. Chen and X. W. Lou, Chem. Commun., 2011, 47, 7155 RSC.
  27. W. Zhou, R. Liu, Q. Wan, Q. Zhang, A. Pan, L. Guo and B. Zou, J. Phys. Chem. C, 2009, 113, 1719 CAS.
  28. C. Wang, Q. Wu, H. L. Ge, T. Shang and J. Z. Jiang, Nanotechnology, 2012, 23, 075704 CrossRef PubMed.
  29. G. Rahman, V. M. Garcıa-Suarez and J. M. Morbec, J. Magn. Magn. Mater., 2013, 328, 104–108 CrossRef CAS PubMed.
  30. H. X. Luan, C. W. Zhang, F. Li, P. Li, M. J. Ren, M. Yuan, W. X. Ji and P. J. Wang, RSC Adv., 2014, 4, 9602 RSC.
  31. B. J. Huang, F. Li, C. W. Zhang, P. Li and P. J. Wang, J. Phys. Soc. Jpn., 2014, 83, 064701 CrossRef.
  32. G. Kresse and J. Hafner, Phys. Rev. B: Condens. Matter Mater. Phys., 1993, 47, 558 CrossRef CAS.
  33. K. Schwarz, P. Blaha and G. Madsen, Comp. Phys. Commun., 2002, 147, 71–76 CrossRef.
  34. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS.
  35. V. I. Anisimov, F. Aryasetiawan and A. I. Lichtenstein, J. Phys.: Condens. Matter, 1997, 9, 767 CrossRef CAS.
  36. H. J. Monkhorst and J. D. Pack, Phys. Rev. B: Solid State, 1976, 13, 5188 CrossRef.
  37. K. F. Mak, C. Lee, J. Hone, J. Shan and T. F. Heinz, Phys. Rev. Lett., 2010, 105, 136805 CrossRef.
  38. T. He, M. W. Zhao, X. J. Zhang, H. Y. Zhang, Z. H. Wang, Z. X. Xi, X. D. Liu, S. S. Yan, Y. Y. Xia and L. M. Mei, J. Phys. Chem. C, 2009, 113, 13610 CAS.
  39. J. M. Themlin, R. Sporken, J. Darville, R. Caudano, J. M. Gilles and R. L. Johnson, Phys. Rev. B: Condens. Matter Mater. Phys., 1990, 42, 11914 CrossRef CAS.
  40. M. A. Maki-Jaskari and T. T. Rantala, Phys. Rev. B: Condens. Matter Mater. Phys., 2001, 64, 075407 CrossRef.
  41. A. F. Kohan, G. Ceder, D. Morgan and C. G. Van de Walle, Phys. Rev. B: Condens. Matter Mater. Phys., 2000, 61, 15019 CrossRef CAS.
  42. L. A. Errico, M. Renterıa and M. Weissmann, Phys. Rev. B: Condens. Matter Mater. Phys., 2005, 72, 184425 CrossRef.
  43. P. Guss, M. E. Foster, B. M. Wong, F. P. Doty, K. Shah, M. R. Squillante, U. Shirwadkar, R. Hawrami, J. Tower and D. Yuan, J. Appl. Phys., 2014, 115, 034908 CrossRef PubMed.
  44. F. Wooten, Optical properties of solids, Academic, New York, 1972, ch. 2, p. 152 Search PubMed.
  45. M. Naeem, S. K. Hasanain and A. Mumtaz, J. Phys.: Condens. Matter, 2008, 20, 025210 CrossRef.

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.