This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Crystal and electronic structure of a quasi-two-dimensional semiconductor Mg3Si2Te6

Chaoxin Huang The authors contributed equally to this work. Center for Neutron Science and Technology, Guangdong Provincial Key Laboratory of Magnetoelectric Physics and Devices, School of Physics, Sun Yat-Sen University, Guangzhou, 510275, China    Benyuan Cheng The authors contributed equally to this work. Shanghai Institute of Laser Plasma, Shanghai 201800, China Center for High Pressure Science and Technology Advanced Research, Shanghai 201203, China    Yunwei Zhang Center for Neutron Science and Technology, Guangdong Provincial Key Laboratory of Magnetoelectric Physics and Devices, School of Physics, Sun Yat-Sen University, Guangzhou, 510275, China    Long Jiang Instrumentation Analysis and Research Center, Sun Yat-Sen UniVersity, Guangzhou 510275, China    Lisi Li    Mengwu Huo    Hui Liu    Xing Huang    Feixiang Liang    Lan Chen    Hualei Sun    Meng Wang [email protected] Center for Neutron Science and Technology, Guangdong Provincial Key Laboratory of Magnetoelectric Physics and Devices, School of Physics, Sun Yat-Sen University, Guangzhou, 510275, China
Abstract

We report the synthesis and characterization of a Si-based ternary semiconductor Mg3Si2Te6, which exhibits a quasi-two-dimensional structure, where the trigonal Mg2Si2Te6 layers are separated by Mg ions. Ultraviolet-visible absorption spectroscopy and density functional theory calculations were performed to investigate the electronic structure. The experimentally determined direct band gap is 1.39 eV, consistent with the value of the density function theory calculations. Our results reveal that Mg3Si2Te6 is a direct gap semiconductor with a relatively narrow gap, which is a potential candidate for infrared optoelectronic devices.

I INTRODUCTION

Narrow-gap semiconductors (NGS) play an important role in optoelectronic and magnetic devices. Because of their unique advantages, the properties and applications of NGS have been widely studied for decades in a variety of fields, such as magnetic field sensors, photovoltaics, infrared photodetectors, lasers and so onElliott (1998); Maier and Hesse (1980); Stradling (1996); Xie et al. (2021); Zhang et al. (2017). Magnetic field sensors, including magnetoresistors and Hall sensors, are generally used in magnetic recording and magnetic measurement technologiesRipka and Janosek (2010); Heremans et al. (1993). Materials for magnetic field sensors are doped semiconductors with high electronic density of states and mobility, such as InSb and InAs which are extremely sensitive to magnetic fieldSolin et al. (2000); Berus et al. (2004); Mihajlović et al. (2005). Conventional photovoltaic materials mostly use wide-gap semiconductors that utilize photons in the ultraviolet and visible regions, limiting the conversion efficiency. Tuning the band gap is considered to be one of the effective solutionsCheng and Yang (2020). NGS-based photoelectrodes can absorb light at longer wavelengths and enhance the solar energy conversion efficiencyZheng et al. (2019). As a more widely used field, infrared photodetectors, such as the quantum dots-in-a-well photodetectors, have developed rapidly in recent yearsDowns and Vandervelde (2013); Martyniuk and Rogalski (2008); Chen et al. (2018). In addition, NGS lasers have attracted more and more attention due to their stability, tunability, simple structure, and many other advantagesTournie and Baranov (2012); Harman and Melngailis (1974).

One of the most important parameters for photoelectric materials is the value of the electronic band gap. Hg1-xCdxTe and Pb1-xSnxTe are among the hottest researched NGS systems in recent years, whose gaps can be tuned by compositions in a wide range, even close to zeroHarman and Melngailis (1974); Rogalski (2005); Lei et al. (2015); Piotrowski and Rogalski (2004). Alloying is also an effective method to modulate the band gap, which is beneficial for the application of the mid- and long-infrared photodetectorsYang et al. (2022). Although InSb and HgCdTe based detectors have been well available in commercial applications, their poor flexibility, instability of the material interface, complex manufacturing process, and low temperature working environment limit their further developmentPiotrowski and Rogalski (1998); Jiao et al. (2022); Yang et al. (2022). Therefore, it is crucial to search for more NGS that are suitable for manufacturing stable and efficient infrared detectors. Meanwhile, new NGS with excellent performance also have potential applications in laser, photovoltaics, and other magnetic devices.

In this work we report the successful synthesis of a Si-based nonmagnetic semiconductor Mg3Si2Te6, whose structure is found to exhibit a quasi-two-dimensional (quasi-2D) layered configuration. Fittings using a direct and indirect band gap models to the ultraviolet-visible (UV-vis) absorption spectra result in a direct gap of 1.39 eV or an indirect gap of 0.6 eV, respectively. Electronic structure calculations using density functional theory (DFT) yield a direct band gap of 1.2 eV, close to the direct gap fitted from the UV-vis absorption spectra, revealing that Mg3Si2Te6 is a direct gap semiconductor. The band gap is narrower than the energies of visible photons, allowing Mg3Si2Te6 a potential candidate for infrared photoelectric material.

Refer to caption
Figure 1: Crystal structure of Mg3Si2Te6 (space group: P3¯1cP\overline{3}1c) viewed (a) perpendicular to the cc-axis and (b) parallel to the cc-axis, respectively. The orange, blue, and grey balls represent Mg, Si, and Te.
Table 1: Parameters of Mg3Si2Te6 refined from single-crystal XRD at 253K.
Empirical formula Mg3Si2Te6
Formula weight 894.69
Temperature 252.99(10) K
Crystal system trigonal
Space group P3¯1cP\overline{3}1c
Unit-cell parameters aa = bb = 7.0642(9) Å
cc = 14.464(2) Å
α\alpha = β\beta = 90
γ\gamma = 120
Atomic parameters
Mg1 4f (1/3, 2/3, 0.4987(5))
Mg2 2d (2/3, 4/3, 1/4)
Si 4e (0, 1, 0.4192(4))
Te 12i (xx, yy, zz)
xx = 0.3402(1), yy = 0.9990(2),
zz = 0.3733(1)
Volume 625.11(18) Å3
Density 4.753 g/cm3
Absorp. coeff. 14.102 mm-1
FF(000) 725.0
Crystal size 0.09 ×\times 0.04 ×\times 0.02 mm3
Radiation Mo Kα (λ\lambda = 0.7107 Å )
2Θ\Theta range for data collection 5.632 to 53.96
Index ranges -9 \leq h \leq 7, -9 \leq k \leq 9,
-18 \leq l \leq 18
Reflections collected 3733
Independent reflections 439
Data/restraints/parameters 439/0/18
Goodness-of-fit on FF2 0.904
Final R indexes [II \geq 2σ\sigma(I)] RR1 = 0.0391, wRwR2 = 0.0790
Final R indexes [all data] RR1 = 0.0767, wRwR2 = 0.1014
Largest diff. peak/hole/e Å-3 1.38/ - 0.96

II EXPERIMENT AND CALCULATION

Single crystal samples of Mg3Si2Te6 were grown by a self-flux methodYin et al. (2020); Sun et al. (2021); Li et al. (2021). Mg (99.99%\%) robs, Si (99.99%\%) powders, and Te blocks (99.99%\%) were mixed in a molar ratio of 3:2:63:2:6 after rough grinding, then sealed in an evacuated quartz ampoule. The ampoule was firstly heated to 600 C in 10 h and held for 15 h, then heated to 1050 C in 20 h and dwelled for 10 h, followed by a slow cooling to 750 C in 150 h before the furnace was shut down. Shiny plate-like single crystals were obtained. The samples are air sensitive and were stored in an argon-filled glove box to minimize the exposure to air.

Refer to caption
Figure 2: (a) A zoomed-in view on the surface of the Mg3Si2Te6 single crystal taken from SEM. The three circles mark the measured spots. (b) An EDS spectrum of the spot 1 in (a). The elements represented by different characteristic peaks are indicated. Small amounts of carbon and oxygen may derive from the background of conductive adhesive tape. The inset is a photo of the single crystal under an optical microscope.

Single-crystal x-ray diffraction (XRD) measurements on a single-crystal x-ray diffractometer (SuperNova, Rigaku), a scanning electron microscopy (SEM) photography, and an energy dispersive x-ray spectroscopy (EDS) (EVO, Zeiss) were employed to determine the crystal structure, morphology and composition. Resistivity was measured using the standard four-probe method on single crystals with a typical size of 3 ×1×0.5\times 1\times 0.5 mm3 on a physical property measurement system (PPMS) (Quantum Design). UV-vis absorption spectroscopy measurements were conducted using an Ocean Optics DH-2000-BAL spectrometer. The sample was pre-peeled into thin flakes with right thickness and then loaded into a diamond anvil cell (DAC) to avoid oxidation. The outboard ring of the DAC is the T301 steel gasket, which was pre-indented and drilled a hole with a diameter of  150 μ\mum to serve as the sample chamber. Each spectrum was collected for 2 seconds with wavelengthes ranged from 180 to 860 nm. The signal from background was obtained by shining the beam on a spot inside the sample chamber but away from the sample.

Electronic band structure was calculated using the DFT within the Perdew-Burke-Ernzerhof (PBE) exchange-correlation as implemented in the Vienna AbinitioAb~{}initio Simulation Package (VASP) codePerdew et al. (1996); Vargas Hernández (2020). A plane wave energy cutoff of 600 eV and dense meshes of Monkhorst-Pack kk-points were used to ensure that all calculations are well converged to 1 meV/atom.

Refer to caption
Figure 3: (a) Resistivity of Mg3Si2Te6 in the temperature range of 328\sim400 K. (b) lnρ\ln\rho as a function of 1000/T1000/T. The red line is a fitting using the thermal activation-energy modol ρ\rho(TT) = ρ\rho0exp(EEa/kkBTT).

III RESULTS AND DISCUSSION

The structure of Mg3Si2Te6 is refined from XRD measurements on single crystals. The structural parameters are summarized in Table 1. Mg3Si2Te6 crystalizes in the trigonal space group P3¯1cP\overline{3}1c with aa = 7.0642(9) Å and cc = 14.464(2) Å at 253 K and exhibits a quasi-2D structure. The crystal structures viewed from different directions are illustrated in Fig. 1. The layer of the abab plane constituting the formula as Mg2Si2Te6 consists of edge shared MgTe6 octahedra and Si-Si dimers. The Mg2Si2Te6 layer is isomorphic to a widely studied 2D van der Waals ferromagnetic compound Cr2Si2Te6Carteaux et al. (1995); Cai et al. (2020). The layers are linked by Mg2 that is half of Mg1, constituting Mg3Si2Te6. Mg3Si2Te6 is isostructural to Mn3Si2Te6Vincent et al. (1986) and Mn3Si2Se6May et al. (2020) which are quasi-2D semiconductors with enriched properties such as ferrimagnetism and large colossal magnetoresistanceNi et al. (2021); Seo et al. (2021).

To determine the composition of the samples, we performed EDS measurements on Mg3Si2Te6 single crystals. Figure 2(a) is a zoomed-in view of the measured crystal in the abab plane taken from SEM. It is obvious that the crystal is layered with some Te flux on the surface. By normalizing the content of Mg to be 3, the composition determined from EDS is Mg3Si1.87(4)Te6.8(2). The ratio of Mg to Si ratio is close to expectation. The exceeding content of Te of 13% can be attributed to the excess Te flux on the surface.

Refer to caption
Figure 4: (a) UV-vis absorption spectra of Mg3Si2Te6 in a DAC at room temperature. (b) The Tauc plots of [FF(RR)hhν\nu]1/n = AA(hhν\nu - EgE_{g}) versus hνh\nu with nn = 1/2 for direct band gap and (c) nn = 2 for indirect band gap. The dotted red lines are fittings of the linear regions.

Figure 3 shows the resistivity measured between 328 and 400 K. The resistivity below 328 K exceeding 106 Ω\Omegacm is beyond the measured range of the instrument. As shown in Fig. 3 (a), the resistivity exhibits an insulating behavior, and no obvious anomaly can be observed up to 400 K. Figure 3 (b) is a plot of the resistivity in lnρ\ln\rho as a function of 1000/TT. A fitting using the thermal activation-energy model ρ\rho(TT) = ρ\rho0 exp(EEa/kkBTT) to the linear segment that corresponds to 350 to 400 K results in Ea=0.378E_{a}=0.378 eV. In the model, ρ\rho0 is a prefactor, kkB is the Boltzmann constant, and EaE_{a} is the thermal activation energy. The obtained activation energy of 0.378 eV indicates that the energy gap of Mg3Si2Te6 may be narrower than that of visible light of 1.643.191.64\sim 3.19 eV. We note the fitted thermal activation energy at 350400350\sim 400 K may be much smaller than the band gap.

Refer to caption
Figure 5: Electronic structure calculation using DFT. The direct gap of Mg3Si2Te6 is about 1.2 eV. PDOS is obtained by integrating density states of the energy bands, and the electrons close to the fermi surface are mainly provided by Te.

UV-vis absorption spectra were employed to study the electronic band gap of Mg3Si2Te6. Five pieces of single crystals were measured separately. The absorption spectra are shown in Fig. 4 (a), yielding a typical semiconducting state with a strong absorption in the range of 700860700\sim 860 nm, while the rest of the short wavelength data are removed because of relatively high noise. The band gap can be obtained by fitting the absorption spectra using the Tauc relation that is [FF(RR)hhν\nu]1/n = AA(hhν\nu - EgE_{g}), where FF(RR) is the Kubelka-Mubelka-Munk function, hh is the Planck constant, ν\nu is the frequency, AA is a prefactor, and EgE_{g} is the band gap energySarkar et al. (2017). For a direct band gap semiconductor, nn is 1/2; for an indirect band gap semiconductor, nn is 2. We plot the Tauc relations as a direct gap and indirect gap semiconductor in Figs. 4 (b) and 4 (c), respectively. The size of the band gap is extracted from the linear regression at the inflection point and the obtained hνh\nu-intercept value is taken as the band gap valueTauc et al. (1966). The measured absorption spectra could not distinguish the two models due to poor statistic. We fitted the linear parts of five sets of data by adopting the Tauc relation and n=1/2n=1/2, resulting in the direct band gaps of 1.41, 1.43, 1.43, 1.37, and 1.31 eV. Thus, the direct gap is estimated to be 1.39(5) eV by averaging the fitted values. Following the same procedure and adopting n=2n=2, an indirect band gap of 0.6(2) eV is obtained. The difference in the calculated energy gaps between different data sets can be attributed to the inconsistency of thicknesses of the samples, which causes calculation errors by affecting the transmittance and reflectivity of the samples to ultraviolet light.

In order to distinguish the direct gap and indirect gap from the fittings of the UV-vis absorption spectra, DFT calculations were performed to investigate the electronic band structure. Figure 5 displays the band structures along high symmetry directions in a Brillouin zone and partial density of states (PDOS) of Mg3Si2Te6. The calculated data reveal a direct gap of \sim1.2 eV at the Γ\Gamma point, close to the fitting results from the UV-vis absorption spectra using the direct gap model. Thus, Mg3Si2Te6 can be considered as a direct gap semiconductor with a gap size of 1.21.391.2\sim 1.39 eV. This value is larger than the thermal activation energy of 0.378 eV fitted from the resistivity between 350 and 400 K. The outcomes may attribute to the fact that the activation energy is fitted from resistance change at high temperature, where thermal fluctuations increase the electron hopping capability, and the gap seems to be narrowed. The direct gap size of Mg3Si2Te6 is smaller than the energies of visible light. Through doping, temperature, and pressure, the band gap of Mg3Si2Te6 may be further decreased, allowing Mg3Si2Te6 one of the potential candidates for infrared photoelectric materials.

IV SUMMARY

In conclusion, we synthesized single crystals of Mg3Si2Te6 by self-flux method, and characterized the structure, composition, resistivity, and electronic band gap. Mg3Si2Te6 exhibits a quasi-2D structure constituted by Mg2Si2Te6 layers (Mg1) that are linked by Mg2 atoms. The single crystals can be mechanically cleaved and are sensitive to air. By combining the UV-vis absorption spectra and DFT calculations, we demonstrate that Mg3Si2Te6 is a direct band gap semiconductor with a gap size of 1.21.391.2\sim 1.39 eV. This gap size is close to that of silicon, and may allow Mg3Si2Te6 become one of the candidates for infrared photoelectric or other functional materials by tuning the energy level by rare earth ions doping. Moreover, the direct gap characteristics may provide higher photon absorption efficiency compared with traditional indirect-gap Si-based semiconductors. The absorption capacity and conversion efficiency require further investigations.

V ACKNOWLEDGMENTS

Work was supported by the National Natural Science Foundation of China (Grants No. 12174454, No. 11904414, No. 11904416, and No. 12104427), the Guangdong Basic and Applied Basic Research Foundation (Grant No. 2021B1515120015), and National Key Research and Development Program of China (Grant No. 2019YFA0705702).

References

  • Elliott (1998) C. T. Elliott, in Infrared Technology and Applications XXIV, Vol. 3436 (SPIE, 1998) pp. 763–775.
  • Maier and Hesse (1980) H. Maier and J. Hesse, Organic Crystals, Germanates, Semiconductors , 145 (1980).
  • Stradling (1996) R. Stradling, Brazilian Journal of Physics 26, 7 (1996).
  • Xie et al. (2021) L. Xie, J. Wang, J. Li, C. Li, Y. Zhang, B. Zhu, Y. Guo, Z. Wang,  and K. Zhang, Advanced Electronic Materials 7, 2000962 (2021).
  • Zhang et al. (2017) H. Zhang, H. Liu, K. Wei, O. O. Kurakevych, Y. Le Godec, Z. Liu, J. Martin, M. Guerrette, G. S. Nolas,  and T. A. Strobel, Physical Review Letters 118, 146601 (2017).
  • Ripka and Janosek (2010) P. Ripka and M. Janosek, IEEE Sensors Journal 10, 1108 (2010).
  • Heremans et al. (1993) J. Heremans, D. Partin, C. Thrush,  and L. Green, Semiconductor Science and Technology 8, S424 (1993).
  • Solin et al. (2000) S. Solin, T. Thio, D. Hines,  and J. Heremans, Science 289, 1530 (2000).
  • Berus et al. (2004) T. Berus, M. Oszwaldowski,  and J. Grabowski, Sensors and Actuators A: Physical 116, 75 (2004).
  • Mihajlović et al. (2005) G. Mihajlović, P. Xiong, S. Von Molnár, K. Ohtani, H. Ohno, M. Field,  and G. J. Sullivan, Applied Physics Letters 87, 112502 (2005).
  • Cheng and Yang (2020) P. Cheng and Y. Yang, Accounts of Chemical Research 53, 1218 (2020).
  • Zheng et al. (2019) J. Zheng, H. Zhou, Y. Zou, R. Wang, Y. Lyu, S. Wang, et al., Energy & Environmental Science 12, 2345 (2019).
  • Downs and Vandervelde (2013) C. Downs and T. E. Vandervelde, Sensors 13, 5054 (2013).
  • Martyniuk and Rogalski (2008) P. Martyniuk and A. Rogalski, Progress in Quantum Electronics 32, 89 (2008).
  • Chen et al. (2018) W. Chen, Z. Deng, D. Guo, Y. Chen, Y. I. Mazur, Y. Maidaniuk, M. Benamara, G. J. Salamo, H. Liu, J. Wu, et al., Journal of Lightwave Technology 36, 2572 (2018).
  • Tournie and Baranov (2012) E. Tournie and A. N. Baranov, Semiconductors and Semimetals 86, 183 (2012).
  • Harman and Melngailis (1974) T. Harman and I. Melngailis, in Applied Solid State Science, Vol. 4 (Elsevier, 1974) pp. 1–94.
  • Rogalski (2005) A. Rogalski, Reports on Progress in Physics 68, 2267 (2005).
  • Lei et al. (2015) W. Lei, J. Antoszewski,  and L. Faraone, Applied Physics Reviews 2, 041303 (2015).
  • Piotrowski and Rogalski (2004) J. Piotrowski and A. Rogalski, Infrared Physics & Technology 46, 115 (2004).
  • Yang et al. (2022) S. Yang, J. Peng, H. Huang, Z. Li, H. Dong,  and F. Wu, Materials Science in Semiconductor Processing 144, 106552 (2022).
  • Piotrowski and Rogalski (1998) J. Piotrowski and A. Rogalski, Sensors and Actuators A: Physical 67, 146 (1998).
  • Jiao et al. (2022) H. Jiao, X. Wang, Y. Chen, S. Guo, S. Wu, C. Song, S. Huang, X. Huang, X. Tai, T. Lin, et al., Science Advances 8, eabn1811 (2022).
  • Yin et al. (2020) J. Yin, C. Wu, L. Li, J. Yu, H. Sun, B. Shen, B. A. Frandsen, D.-x. Yao,  and M. Wang, Physical Review Materials 4, 013405 (2020).
  • Sun et al. (2021) H. Sun, C. Chen, Y. Hou, W. Wang, Y. Gong,  and M. Huo, Science China: Physics, Mechanics and Astronomy 64, 118211 (2021).
  • Li et al. (2021) L. Li, X. Hu, Z. Liu, J. Yu, B. Cheng, S. Deng, L. He, K. Cao, D.-X. Yao,  and M. Wang, Science China: Physics, Mechanics and Astronomy 64, 287412 (2021).
  • Perdew et al. (1996) J. P. Perdew, K. Burke,  and M. Ernzerhof, Physical Review Letters 77, 3865 (1996).
  • Vargas Hernández (2020) R. A. Vargas Hernández, The Journal of Physical Chemistry A 124, 4053 (2020).
  • Carteaux et al. (1995) V. Carteaux, D. Brunet, G. Ouvrard,  and G. Andre, Journal of Physics: Condensed Matter 7, 69 (1995).
  • Cai et al. (2020) W. Cai, H. Sun, W. Xia, C. Wu, Y. Liu, H. Liu, Y. Gong, D.-X. Yao, Y. Guo,  and M. Wang, Physical Review B 102, 144525 (2020).
  • Vincent et al. (1986) H. Vincent, D. Leroux, D. Bijaoui, R. Rimet,  and C. Schlenker, Journal of Solid State Chemistry 63, 349 (1986).
  • May et al. (2020) A. F. May, H. Cao,  and S. Calder, Journal of Magnetism and Magnetic Materials 511, 166936 (2020).
  • Ni et al. (2021) Y. Ni, H. Zhao, Y. Zhang, B. Hu, I. Kimchi,  and G. Cao, Physical Review B 103, L161105 (2021).
  • Seo et al. (2021) J. Seo, C. De, H. Ha, J. E. Lee, S. Park, J. Park, Y. Skourski, E. S. Choi, B. Kim, G. Y. Cho, H. W. Yeom, S.-W. Cheong, J. H. Kim, B.-J. Yang, K. Kim,  and J. S. Kim, Nature 599, 576 (2021).
  • Sarkar et al. (2017) A. Sarkar, C. Loho, L. Velasco, T. Thomas, S. S. Bhattacharya, H. Hahn,  and R. Djenadic, Dalton Transactions 46, 12167 (2017).
  • Tauc et al. (1966) J. Tauc, R. Grigorovici,  and A. Vancu, Physica Status Solidi (b) 15, 627 (1966).