Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Growth, characterization and theoretical analysis of α-SrGeO3 as a candidate mid-IR stimulated Raman scattering crystal

Hailong Wang ab, Bin Li ab, Ying Zhou c, Guimei Zheng b, Xue Zhang b and Songming Wan *bd
aUniversity of Science and Technology of China, Hefei 230026, China
bAnhui Provincial Key Laboratory of Photonics Devices and Materials, Anhui Institute of Optics and Fine Mechanics, HFIPS, Chinese Academy of Sciences, Hefei 230031, China. E-mail: smwan@aiofm.ac.cn
cInstitutes of Physical Science and Information Technology, Anhui University, Hefei 230601, China
dAdvanced Laser Technology Laboratory of Anhui Province, Hefei 230037, China

Received 30th July 2024 , Accepted 14th September 2024

First published on 16th September 2024


Abstract

The α-BaGeO3 crystal is a potential stimulated Raman scattering (SRS) medium for use in generating lasers at wavelengths in the range of 2.1–2.4 μm. However, the growth of a large-sized and high-quality α-BaGeO3 crystal is still challenging. Here, we turn the attention to its analogue, α-SrGeO3. The α-SrGeO3 crystal was grown by the high-temperature solution method under conditions similar to those for growth of α-BaGeO3. The grown α-SrGeO3 single crystals have larger sizes (approximately 8 × 12 × 10 mm3) and higher quality than α-BaGeO3. A strong Raman peak at 813 cm−1 and a wide transparent window from 0.22 to 5.78 μm indicate that the α-SrGeO3 crystal is a promising SRS crystal working in the 2.1–2.4 μm wavelength range. From density functional theory (DFT) computations, the strongest Raman peak, the ultraviolet cut-off edge and the infrared cut-off edge of α-SrGeO3 are all related to the [Ge3O9]6− ring, the basic building unit in both α-SrGeO3 and α-BaGeO3, which suggests that the [Ge3O9]6− ring is a valuable structural group for developing new mid-IR SRS crystals.


Introduction

Lasers at wavelengths in the range of 2.1–2.4 μm have excellent atmospheric penetrability and eye safety, and also overlap with the characteristic absorption bands of many air pollutants,1–4 making them suitable for various applications, including laser ranging, remote sensing, environmental monitoring and free-space communication.3–7 A simple and effective method to generate the 2.1–2.4 μm lasers is by use of the stimulated Raman scattering (SRS) effect in crystal.8–11 Our previous work has revealed that the low-temperature phase BaGeO3 (α-BaGeO3) crystal is a promising SRS crystal for use in the range of 2.1–2.4 μm due to its strong Raman response and wide infrared transparent window.12 In order to circumvent the phase transition,13 the α-BaGeO3 crystal has to be grown by the high-temperature solution method.12 Although considerable effort has been devoted, the growth of large-sized, high-quality α-BaGeO3 crystals remains difficult. A problem frequently encountered is that small α-BaGeO3 crystal plates are apt to form and float on the solution surface, which influences the crystal growth and ultimately deteriorates the crystal quality.

Crystal growth in a high-temperature solution begins with the desolvation of crystal growth units (also the building units of the crystal in most cases14,15) achieved by slowly cooling the solution or evaporating the solvent. Then, the growth units are transported to the crystal–solution interface by diffusion, natural and forced convection (usually, the forced convection is the main contribution), and eventually integrated into the crystal by an interface kinetic process.16 Only when the integrated rate of the growth unit into the crystal exceeds the transport rate and the transport rate exceeds the desolvation rate is the crystal growth stable. In our experiments for growth of the α-BaGeO3 crystal, the desolvation rates are rather low (the cooling rates are less than 1.0 °C per day) and the forced convection is sufficient to transport the growth units (the maximal crystal rotational rates are 20 rotations per minute), therefore, we deem that the interface kinetic process is the key step to influence the α-BaGeO3 crystal growth.

The interface kinetic process of the α-BaGeO3 crystal growth is depicted in Fig. 1. The α-BaGeO3 crystal is a pseudo-wollastonite-type structure alternately stacked by the layer of the Ba2+ cation and the layer of the [Ge3O9]6− ring.12 During the crystal growth, the [Ge3O9]6− ring in the solution (the growth unit) is absorbed onto the crystal–solution interface by the Ba2+ cation at the interface, and vice versa. The absorbed strength determines the rate of the growth unit integrated into the crystal.17 Owing to the large size, the Ba2+ cation weakly bonds to the [Ge3O9]6− ring, which brings about a low crystal growth rate. As a result, numerous growth units that fail to precipitate on the interface organize them into new crystal nuclei and finally destroy the single crystal growth.


image file: d4tc03340j-f1.tif
Fig. 1 Illustration of the kinetic process occurring near the α-BaGeO3 (or α-SrGeO3) crystal–solution interface (M = Ba or Sr).

The low-temperature phase SrGeO3 (α-SrGeO3) crystal has a structure very similar to that of the α-BaGeO3 crystal. It is also the pseudo-wollastonite-type structure and crystallizes in the monoclinic C2/c space group with the unit cell parameters (a = 12.533(3) Å, b = 7.262(1) Å, c = 11.259(3) Å, β = 111.30(2)° and Z = 12) close to those of the α-BaGeO3 crystal (a = 13.178(10) Å, b = 7.626(6) Å, c = 11.670(9) Å, β = 111.638(8)° and Z = 12). The α-SrGeO3 crystal is built up of Sr2+ cations and [Ge3O9]6− rings.18 As compared with Ba2+, Sr2+ has a smaller size and thus a stronger interaction with the [Ge3O9]6− ring. On the base of the foregoing analysis regarding the interface kinetics of the α-BaGeO3 crystal growth, we believe that, under similar conditions, the integrated rate of the growth units into the α-SrGeO3 crystal is larger than that into the α-BaGeO3 crystal, and thus the spontaneous nucleation likely to occur in the α-SrGeO3 crystal growth can be suppressed. Due to the structural similarity, the α-SrGeO3 crystal is expected to exhibit similar spectral performances to those of the α-BaGeO3 crystal, that is, the α-SrGeO3 crystal also has good spectral characteristics as an SRS crystal for use in the range of 2.1–2.4 μm. Furthermore, like α-BaGeO3 which is stable below 1100 °C,19 the α-SrGeO3 crystal remains stable below 1020 °C.20

In this study, we use the high-temperature solution method to grow the α-SrGeO3 crystal under conditions similar to those for growth of the α-BaGeO3 crystal to verify that the α-SrGeO3 crystal has a growth performance superior to the α-BaGeO3 crystal. Raman and transmission spectra of the α-SrGeO3 crystal are studied by experimental and theoretical approaches to evaluate its potential as an SRS crystal.

Experimental and computational methods

Single crystal growth

Owing to the phase transition, the α-SrGeO3 crystal has to be grown by the high-temperature solution method.16 For comparison, the conditions for growth of the α-BaGeO3 crystal were applied to grow the α-SrGeO3 crystal. The crystal growth raw materials, polycrystalline α-SrGeO3 and NaBO2 (the flux), were synthesized through solid-state reactions (see the ESI for more details). The raw materials were mixed with a molar ratio of n(α-SrGeO3)[thin space (1/6-em)]:[thin space (1/6-em)]n(NaBO2) = 1[thin space (1/6-em)]:[thin space (1/6-em)]1.25, loaded into a platinum crucible, and then placed in a crystal growth furnace. After the raw materials were completely melted, an α-SrGeO3 seed crystal was immersed in the solution to determine the saturation temperature. Once the temperature was determined, another seed crystal with the direction perpendicular to the (001) plane was employed to grow the α-SrGeO3 crystal, with a cooling rate of 0.5–1.0 °C per day and a crystal rotational rate of 5–20 rotations per minute. One month later, the growth was ended. The grown α-SrGeO3 crystal was cut. One part was crushed into powder for X-ray diffraction (XRD) measurement; another part was polished into a crystal plate for collecting Raman and transmission spectra.

Powder X-ray diffraction

The XRD pattern of the powder was recorded at room temperature on a Rigaku TTRAX-III X-ray powder diffractometer equipped with the Cu Kα radiation (λ = 1.54056 Å). Diffraction data were collected in the 2θ range from 10 to 70° with a step width of 0.02°.

Raman spectroscopy

Raman spectra of the α-SrGeO3 and α-BaGeO3 crystals were recorded on a Renishaw inVia-Reflex Raman spectrometer with a backscattering configuration. A Q-switched SHG Nd:YAG laser, with a wavelength at 532 nm and a power of 10 mW, was utilized as the excitation resource. The spectral data were collected in the frequency range from 100 to 4000 cm−1 with a resolution less than 1.5 cm−1. Before each experiment, the frequency was calibrated by a single-crystal silicon wafer.

Transmission spectroscopy

A 3-mm-thick α-SrGeO3 crystal plate was employed to collect transmission spectra. The UV-Vis-NIR transmission spectrum was recorded on a PerkinElmer Lambda-900 UV-Vis-NIR spectrophotometer in the wavelength range from 0.19 to 2.5 μm, and the IR spectrum was recorded by using a Gangdong FTIR-650G spectrophotometer in the wavelength range from 2.5 to 25 μm.

Computational methods

The Raman spectrum and electronic structure of the α-SrGeO3 crystal were studied by the density functional theory (DFT) method,21 which were implemented in the Cambridge serial total energy package (CASTEP).22 The crystal structure of α-SrGeO3 reported by Nishi was adopted as the structural model.18 The local density approximation (LDA) in the Perdew and Zunger parameterization of the numerical results of Ceperley and Alder (CA-PZ) was selected to treat the exchange–correlation energy.23 The norm-conserving pseudo-potential was used to describe the interaction between ionic cores and valence electrons.24 The valence electron configurations of strontium, germanium and oxygen were 4s24p65s2, 4s24p2 and 2s22p4, respectively. The convergence criteria of the total energy change, displacement, force and stress for the structural optimization were 10−5 eV per atom, 0.001 Å, 0.03 eV Å−1 and 0.05 GPa, respectively. According to convergence tests, a k-point grid of 1 × 2 × 1 and an energy cutoff of 1000 eV were sufficient to ensure the total energy convergence to within 1 meV per atom (Table S1 and Fig. S3, ESI). The computational Raman peak intensities were corrected using the Bose–Einstein factors calculated with the wavelength of the excitation resource (532 nm) and the room temperature of 300 K.25 All of the Raman lines were broadened by the Lorentzian line-shape function with a fixed full width at half maximum (FWHM) of 10 cm−1.26 The Heyd–Scuseria–Ernzerhof 06 (HSE06) screened hybrid functional was employed to compute the electronic structure of the α-SrGeO3 crystal.27,28 The optimized structural model and parameters for computation of the crystal Raman spectrum were adopted in the electronic structure computation.

Results and discussion

Single crystal growth

A typical α-SrGeO3 crystal is present in Fig. 2a. The crystal is made up of two single crystals (twin crystals), with regular shapes and characterized by the hexagonal (001) end faces. The two single crystal sizes are about 8 × 12 × 10 mm3. Fig. 2c shows the powder XRD of the crystal. All the diffraction peaks are consistent with those in the standard pattern of α-SrGeO3 (JCPDS no. 87–469), indicating that the crystal is α-SrGeO3. For comparison, an α-BaGeO3 crystal grown under similar conditions [the same flux (NaBO2), flux concentration, seed crystal direction, cooling rate and crystal rotational rate] is shown in Fig. 2b. The crystal is made up of five single crystals, and thus exhibits a more complex external shape. Each of the single crystals is also characterized by the hexagonal (001) end faces; yet their sizes are smaller than that of the α-SrGeO3 single crystals. Suffering from the multiple twinning effect and the resulting crystal boundaries and inclusions, the α-BaGeO3 crystal quality is inferior to the α-SrGeO3 crystal. As we expected, the growth performance of the α-SrGeO3 crystals is superior to that of the α-BaGeO3 crystal.
image file: d4tc03340j-f2.tif
Fig. 2 (a) α-SrGeO3 single crystals grown by the high-temperature solution method, (b) α-BaGeO3 single crystals grown under similar conditions and (c) the powder XRD pattern of the α-SrGeO3 crystal, with the standard XRD pattern.

Raman spectral characteristics

The experimental Raman spectrum of the α-SrGeO3 crystal is shown in Fig. 3. At least 13 Raman peaks are observable, but only one significantly strong and narrow peak, located at 813 cm−1, is present in the spectrum. This spectral characteristic is beneficial for generating SRS lasers. For comparison, the Raman spectrum of the α-BaGeO3 crystal was recorded under the same conditions. The two spectra are very similar (Fig. 3). The intensity and FWHM of the strongest 813 cm−1 peak of α-SrGeO3 are comparable to those of the strongest 808 cm−1 peak of α-BaGeO3. Therefore, the α-SrGeO3 crystal also has a strong Raman response, as with the α-BaGeO3 crystal.12
image file: d4tc03340j-f3.tif
Fig. 3 Experimental Raman spectra of the α-SrGeO3 and the α-BaGeO3 crystal. The FWHMs of the α-SrGeO3 813 cm−1 peak and the α-BaGeO3 808 cm−1 peak are 4.8 cm−1 and 5.0 cm−1, respectively.

The DFT computation reveals that the α-SrGeO3 crystal has 90 vibrational modes (20Ag + 23Au + 22Bg + 25Bu), including three acoustic modes (Au + 2Bu), 42 Raman active modes (20Ag + 22Bg), and 45 IR active modes (22Au + 23Bu) (Table S2 in the ESI). The computational Raman spectrum is shown in Fig. 4a, which is in good agreement with the experimental one both in frequency and in intensity. On the base of the computational result, two important crystal characteristic peaks, located at 476 and 813 cm−1 (experimental values), are assigned. The 476 cm−1 peak originates from the breathing vibration of the three bridge oxygen atoms in the [Ge3O9]6− ring; the 813 cm−1 peak arises from the symmetrical stretching vibration of the six extra-ring Ge–O bonds (Fig. 4b). Benefiting from the synergistic vibrational effect, the 813 cm−1 peak has a strong Raman response. The other characteristic Raman peaks of the α-SrGeO3 crystal are also attributed to the vibrations of the [Ge3O9]6− ring (Fig. S4 in the ESI). These results are consistent with the computational results of α-BaGeO3,12 and manifest that the Raman spectral similarity between α-SrGeO3 and α-BaGeO3 is due to their structural similarity.


image file: d4tc03340j-f4.tif
Fig. 4 (a) Experimental and computational Raman spectra of the α-SrGeO3 crystal and (b) two characteristic vibrational modes of the crystal.

UV-Vis-NIR transmission spectral characteristics

The UV-Vis-NIR transmission spectrum of the α-SrGeO3 crystal is shown in Fig. 5a. The UV cut-off edge is located at 0.22 μm; the transmittance is greater than 70% from 1.2 to 2.5 μm. Therefore, the crystal is suitable for application in the range of 2.1–2.4 μm. Moreover, we employed the Tauc plot method to evaluate the crystal bandgap.29,30 The absorption plot of (αhν)1/2versus photon energy (), derived from the UV-Vis-NIR transmission spectrum, is displayed in Fig. 5b, which results in an experimental bandgap at 5.20 eV.
image file: d4tc03340j-f5.tif
Fig. 5 (a) A UV-Vis-IR transmission spectrum of the α-SrGeO3 crystal [the inset is the α-SrGeO3 crystal plate (16 × 9 × 3 mm3) for the spectrum measurement], (b) a plot of (αhv)1/2versus hν for the crystal, (c) the band structure of the crystal and (d) TDOS and PDOSs of the crystal.

The UV absorption is related to the electronic transition from the valence band (VB) top to the conduction band (CB) bottom.31 The computational band structure of the α-SrGeO3 crystal is presented in Fig. 5c. The VB top and CB bottom have different k-points, revealing its indirect bandgap nature. The computational bandgap (5.126 eV) is very close to the experimental value (5.20 eV). Fig. 5d shows the total and partial densities of states (TDOS and PDOSs) of the α-SrGeO3 crystal. The VB top and CB bottom are mainly composed of the O-2p state and Ge-4s state, respectively, indicating that the UV absorption of the α-SrGeO3 crystal is ascribed to the electronic transition from the O-2p to the Ge-4s state.

IR transmission spectral characteristics

As shown in Fig. 6a, the α-SrGeO3 crystal is highly transparent from 2.5 to 5 μm, which allows the crystal to extend its application to the mid-IR range. It is worth noting that a few weak absorption peaks are present from 2.7 to 4.5 μm. These peaks can be assigned to the vibrations of hydroxyl,32 which implies that the crystal is weakly hygroscopic. As a result, the surface was hydrolysed when the crystal was placed in air for a long time or cut/polished in a watery environment. However, the hydrolysis can be avoided if the crystal is stored in a dryer and processed under a water-free condition.
image file: d4tc03340j-f6.tif
Fig. 6 (a) A mid-IR transmission spectrum of the α-SrGeO3 crystal and (b) the highest-frequency Au and Bg modes.

The IR cut-off edge of the α-SrGeO3 crystal is at 5.78 μm (Fig. 6a), close to that of α-BaGeO3 (5.9 μm).12 In general, the IR cut-off edge corresponds to the highest-frequency second-order IR-active modes.33 The α-SrGeO3 crystal has the C2 point group symmetry. Its second-order IR-active mode is the coupled mode that combines an Au mode with a Bg mode (Au ⊗ Bg).34 According to the DFT computation, the highest frequencies of the Au and the Bg modes of α-SrGeO3 are 853 cm−1 and 884 cm−1 (Table S2 in the ESI), respectively. Therefore, the highest frequency of the Au ⊗ Bg mode is 1737 cm−1, which corresponds to an IR absorption wavelength of 5.76 μm. The computational wavelength is very close to the experimental IR cut-off edge, which manifests that the IR cut-off edge arises from the combination of the 853 cm−1 and 884 cm−1 modes (Fig. 6b).

Conclusions

The α-SrGeO3 crystal growth is dominated by an interface kinetic process. The stronger bonding between the Sr2+ cation and the [Ge3O9]6− ring can improve the single-crystal size and the crystal quality. Under the conditions that were adopted to grow the α-BaGeO3 crystal, α-SrGeO3 single crystals with sizes of about 8 × 12 × 10 mm3 were obtained. The multiple-twinning effect was suppressed. The α-SrGeO3 crystal has a strong Raman response, comparable to that of the α-BaGeO3 crystal, and a wide transparent window ranging from 0.22 to 5.78 μm. These experimental results confirm that the α-SrGeO3 crystal is a promising SRS medium for use in the mid-IR range, including the 2.1–2.4 μm range. According to the DFT computations, the primary spectral characteristics of the α-SrGeO3 crystal are determined by the [Ge3O9]6− ring. The strongest Raman peak originates from the symmetric stretching vibration of the six extra-ring Ge–O bonds; the UV cut-off edge is ascribed to the electronic transition from the O-2p to the Ge-4s state, within the ring; the IR cut-off edge corresponds to the highest-frequency Au ⊗ Bg mode, arising from the vibrations of the ring. The computational results reveal that the [Ge3O9]6− ring is a good structural group for exploring new IR SRS crystals.

Author contributions

H. W. and B. L. performed the crystal growth and the DFT computations. Y. Z. and S. W. conducted the spectral measurements. S. W. initiated and supervised the work. All authors participated in the discussion.

Data availability

The data supporting this article have been included as part of the ESI.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

We thank Prof. Guochun Zhang (Beijing Center for Crystal Research and Development, TIPC, Chinese Academy of Sciences) for his support in measuring the transmission spectra. This work was financially supported by the Open Project of Advanced Laser Technology Laboratory of Anhui Province (grant no. AHL2022ZR04), and the HFIPS Director's Fund (grant no. YZJJ202301-TS). The DFT computations were performed in part at the Centre for Computational Science, HFIPS, Chinese Academy of Sciences.

Notes and references

  1. H. Horvath, Atmos. Environ., 1993, 27, 293–317 CrossRef.
  2. S. W. Henderson, C. P. Hale, J. R. Magee, M. J. Kavaya and A. V. Huffaker, Opt. Lett., 1991, 16, 773–775 CrossRef CAS.
  3. Mid-infrared Coherent Sources and Applications, ed. M. Ebrahim-Zadeh and I. T. Sorokina, Springer, Netherlands, 2008 Search PubMed.
  4. R. L. Byer, Science, 1988, 239, 742–747 CrossRef CAS PubMed.
  5. S. D. Jackson, Nat. Photonics, 2012, 6, 423–431 CrossRef CAS.
  6. K. Scholle, S. Lamrini, P. Koopmann and P. Fuhrberg, in Frontiers in Guided Wave Optics and Optoelectronics, ed. B. Pal, INTECH, Croati, 2010, pp. 471–500 Search PubMed.
  7. J. G. Daly and C. A. Smith, 2.0-micron laser applications, Proc. SPIE, 1992, 1627, 26–31 CrossRef CAS.
  8. P. Černý, H. Jelínková, P. G. Zverev and T. T. Basiev, Prog. Quant. Electron., 2004, 28, 113–143 CrossRef.
  9. J. Q. Zhao, Y. Li, S. Zhang, L. Li and X. L. Zhang, Opt. Express, 2015, 23, 10075–10080 CrossRef CAS.
  10. U. Sheintop, D. Sebbag, P. Komm, S. Pearl, G. Marcus and S. Noach, Opt. Express, 2019, 27, 17112–17121 CrossRef CAS.
  11. E. Perez, U. Sheintop, R. Nahear, G. Marcus and S. Noach, Opt. Lett., 2020, 45, 5409–5412 CrossRef.
  12. S. M. Wan, Y. Zeng, Y. N. Yao, M. R. D. Mutailipu, J. Han, S. J. Jiang, S. J. Zhang and S. L. Pan, Inorg. Chem., 2020, 59, 3542–3545 CrossRef CAS.
  13. D. M. Többens, V. Kahlenberg, C. Gspan and G. Kothleitner, Acta Crystallogr., Sect. B: Struct. Sci., 2006, 62, 1002–1009 CrossRef.
  14. S. J. Zhang, S. M. Wan, Y. Zeng, S. J. Jiang, X. Y. Gong and J. L. You, Inorg. Chem., 2019, 58, 5025–5030 CrossRef CAS PubMed.
  15. X. S. Lv, Y. L. Sun, X. L. Tang, S. M. Wan, Q. L. Zhang, J. L. You and S. T. Yin, J. Cryst. Growth, 2013, 371, 107–111 CrossRef CAS.
  16. D. Elwell and H. J. Scheel, Crystal Growth from High-Temperature Solutions, Academic Press, London, UK, 1975 Search PubMed.
  17. C. T. Sun and D. F. Xue, CrystEngComm, 2016, 18, 1262–1272 RSC.
  18. F. Nishi, Acta Crystallogr., Sect. C: Cryst. Struct. Commun., 1997, 53, 399–401 CrossRef.
  19. J. P. Guha, D. Kolar and A. Porenta, J. Thermal Anal., 1976, 9, 37–41 CrossRef CAS.
  20. O. Yamaguchi, H. Sasaki, K. Sugiura and K. Shimizu, Ceram. Int., 1983, 9, 75–79 CrossRef CAS.
  21. K. Refson, P. R. Tulip and S. J. Clark, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 73, 155114 CrossRef.
  22. V. Milman, K. Refson, S. J. Clark, C. J. Pickard, J. R. Yates, S. P. Gao, P. J. Hasnip, M. I. J. Probert, A. Perlov and M. D. Segall, J. Mol. Struct.: THEOCHEM, 2010, 954, 22–35 CrossRef CAS.
  23. J. P. Perdew and A. Zunger, Phys. Rev. B: Condens. Matter Mater. Phys., 1981, 23, 5048–5079 CrossRef CAS.
  24. D. R. Hamann, M. Schlüter and C. Chiang, Phys. Rev. Lett., 1979, 43, 1494–1497 CrossRef CAS.
  25. D. Porezag and M. R. Pederson, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 7830–7836 CrossRef CAS.
  26. C. M. Foster, M. Grimsditch, Z. Li and V. G. Karpov, Phys. Rev. Lett., 1993, 71, 1258–1260 CrossRef CAS PubMed.
  27. A. V. Krukau, O. A. Vydrov, A. F. Izmaylov and G. E. Scuseria, J. Chem. Phys., 2006, 125, 224106 CrossRef PubMed.
  28. T. M. Henderson, J. Paier and G. E. Scuseria, Phys. Status Solidi B, 2011, 248, 767–774 CrossRef CAS.
  29. J. Tauc, R. Grigorovici and A. Vancu, Phys. Status Solidi B, 1966, 15, 627–637 CrossRef CAS.
  30. P. Makuła, M. Pacia and W. Macyk, J. Phys. Chem. Lett., 2018, 9, 6814–6817 CrossRef PubMed.
  31. C. Kittel, Introduction to Solid State Physics, John Wiley & Sons, New Jersey, USA, 2005 Search PubMed.
  32. H. N. Li, T. Baikie, S. S. Pramana, J. Felix Shin, P. J. Keenan, P. R. Slater, F. Brink, J. Hester, T. An and T. J. White, Inorg. Chem., 2014, 53, 4803–4812 CrossRef CAS.
  33. S. J. Jiang, S. M. Wan, W. Luo, B. Li and J. Y. Yao, J. Mater. Chem. C, 2022, 10, 649–654 RSC.
  34. E. Kroumova, M. I. Aroyo, J. M. Perez-Mato, A. Kirov, C. Capillas, S. Ivantchev and H. Wondratschek, Phase Transitions, 2003, 76, 155–170 CrossRef CAS.

Footnotes

Electronic supplementary information (ESI) available: Synthesis of polycrystalline raw materials for growth of the α-SrGeO3 crystal, convergence tests for the total energy of α-SrGeO3, and the crystal vibrational modes and their corresponding frequencies. See DOI: https://doi.org/10.1039/d4tc03340j
H. Wang and B. Li contributed equally in this work.

This journal is © The Royal Society of Chemistry 2024
Click here to see how this site uses Cookies. View our privacy policy here.