Next Article in Journal
Comparison of the Flexural Behavior of High-Volume Fly AshBased Concrete Slab Reinforced with GFRP Bars and Steel Bars
Next Article in Special Issue
Zirconium Containing Periodic Mesoporous Organosilica: The Effect of Zr on CO2 Sorption at Ambient Conditions
Previous Article in Journal
In Vitro Electrochemical Corrosion Assessment of Magnesium Nanocomposites Reinforced with Samarium(III) Oxide and Silicon Dioxide Nanoparticles
Previous Article in Special Issue
Molecular Dynamics Study of Melting Behavior of Planar Stacked Ti–Al Core–Shell Nanoparticles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Negative Thermal Expansion Properties of Sm0.85Sr0.15MnO3-δ

Department of Avionics Engineering, Aviation Maintenance NCO Academy, Air Force Engineering University, Xinyang 464000, China
*
Author to whom correspondence should be addressed.
J. Compos. Sci. 2022, 6(6), 156; https://doi.org/10.3390/jcs6060156
Submission received: 1 May 2022 / Revised: 19 May 2022 / Accepted: 24 May 2022 / Published: 25 May 2022
(This article belongs to the Special Issue Metal Composites)

Abstract

:
A novel negative thermal expansion (NTE) material composed of Sm0.85Sr0.15MnO3-δ was synthesized using the solid-state method. By allowing Sr2+ to partially replace Sm3+ in SmMnO3, the ceramic material Sm0.85Sr0.15MnO3-δ exhibits NTE properties between 360K and 873K, and its average negative thermal expansion coefficient was −10.08 × 10−6/K. The structure of Sm0.85Sr0.15MnO3-δ is orthogonal, the space group is pbnm, the morphology is regular, and the grain size is uniform. The results of X-ray diffraction and XPS (X-ray photoelectron spectroscopy) suggest that the NTE phenomenon is related to the electron transfer of Mn ions. With the increase in temperature, Mn4+ is rapidly transformed into Mn3+, accompanied by Mn4+O6 octahedron distortion and oxygen defects. It was found that the sample volume continually decreased at the same time.

1. Introduction

We know that most instruments are composed of various materials, but with increases in temperature, different thermal expansion coefficients of various constituent materials may lead to thermal mismatches, and small cracks in the equipment can lead to performance failures and even instrument damage. NTE materials have attracted considerable research attention in the production of composites with accurately controllable positive, negative, or zero coefficients of thermal expansion [1,2,3,4,5,6,7,8,9,10,11].
A great number of NTE materials have been found, such as oxides (Cu1.5Mg0.5V2O7, Cu2V2O7, and HfMnMo3O12, etc.) [12,13,14,15,16], antiperovskite Mn3XN, and perovskite (BiNiO3, Gd1-xSrxMnO3-σ, and Er0.7Sr0.3NiO3-δ, etc.) [17,18,19,20,21]. However, each material has limitations because of some defects. ZrW2O8 is a metastable phase at room temperature (RT), which is difficult to prepare due to it readily decomposing [1]. ZrV2O7 exists as a phase transformation at 375K [2]. Y2Mo3O12 has a water-absorbing quality at RT. Although antiperovskite (Mn3Cu(Ge)N, Mn3NiN, and Mn3ZnN, etc.) possesses the properties of superconductivity, giant magnetoresistance, magnetocaloric effects, and constant electrical resistivity [8], the NTE temperature range is usually under RT, and its preparation conditions are very strict. Mn3Cu(Ge)N needs to be grown on a silicon surface with high pressure and argon gas protection. The NTE perovskite ABO3 (A = Gd, Er, and Bi, etc.; B = Mn, Er, Sr, Ni and Sr, etc.) not only shows NTE properties in a large temperature range above RT but also has simple preparation conditions.
Kurimamachiya-chouses conducted research on Sr2+ partly substituting Gd3+ in GdMnO3. They pointed out that Gd1-xSrxMnO3-δ had excellent NTE properties [18]. L. J. Fu reported the NTE material of Er0.7Sr0.3NiO3-δ with Sr2+ partly substituting Er3+ in ErNiO3 [19]. These studies suggest that the substituting method is an effective way to prepare new kinds of NTE materials with excellent properties [16,18,19,20]. In the present study, we conducted research on Sr2+ partly substituting Sm3+ in SmMnO3. The thermal properties are discussed.

2. Experimental Procedures

The sample was prepared according to the conventional solid-state method. Analytic-grade Sm2O3 (purity 99.5%), SrO (purity 99.5%), and MnO2 powder were used as raw materials. Using MnO2 as the raw material, Mn2O3 powder was prepared by burning in a 923 K furnace for 10 h.
Sm2O3, SrO, and Mn2O3 powders were mixed according to the mole ratio of Sm:Sr:Mn = 0.85:0.15:1. The mixtures were ground using an agate mortar for 1 h and then ground with ethanol for 2 h. The obtained mixtures were then dried for 1 h at 353 K in a baking oven. Afterward, the mixtures were pressed into cylindrical-shape compacts (Ø10 × 5 mm) using a powder pellet machine (769YP-15A, 200 MPa). The compacts were initially sintered in a pipe furnace (AY-BF-555-180) at 1273 K for 10 h in air and subsequently sintered at 1623 K for 10 h. The sample was allowed to cool in the furnace naturally.
The linear thermal expansion coefficient was measured using a dilatometer Linseis L76 (heating and cooling rates of 5 K/min). The XRD measurement was carried out using Bruker D8 Advance with CuKα radiation. The XRD pattern of the sample was analyzed using X’Pert HighScore Plus software. The lattice constants a, b, and c and the unit cell volume of the sample were calculated using powderX software and the least square method. The surface morphology of the sample was observed using the FEI Quanta 250 scanning electron microscopy (SEM), and the EDS energy spectrum was obtained using an Appllo XP. The TGA and DSC were tested using a LabsysTM thermal analyzer. The XPS (X-ray photoelectron spectroscopy) was performed using a Thermo Scientific K-Alpha instrument for the valence analysis of the Mn element. The BET tests were performed to determine the size and volume of the holes using an ASAP2460 device.

3. Results and Discussion

3.1. Phase Analysis

Figure 1a is the XRD pattern of the sample at RT. Comparing the XRD pattern with the JCPDS cards for SmMnO3 (00-025-0747), Eu0.9Sr0.1MnO3 (No. 00-051-0252), and Eu0.8Sr0.2MnO3 (00-051-0251), we found that the diffraction peaks were similar to those of the JCPDS cards, except for some shifts, which suggests that the as-prepared sample had similar structure to that of SmMnO3, Eu0.9Sr0.1MnO3, and Eu0.8Sr0.2MnO3. It can be confirmed that the ceramic Sm0.85Sr0.15MnO3-δ crystallizes in an orthorhombic structure. As the ionic radius of Sr2+ (ionic radius 1.18 Å) is bigger than that of Sm3+ (ionic radius 0.958 Å), the difference in the ionic radius may cause lattice distortion. As Sr2+ partly substitutes for Sm3+, the diffraction peaks also shift.
Figure 2a shows the SEM image of the sample. We found that the ceramic sample was composed of homogenous spherical or elliptic spherical particles with some obvious agglomerations. There were pores and microcracks in the sintered body. The size of the particles was uniform, with an average grain size of about 1~2 μm. The EDS analysis of the sample revealed the primary elements of Sm, Sr, Mn, and O, and their atomic ratio (Sm:Sr:Mn:O) was about 0.85:0.15:1:3 (seeing Table 1). Combined with the XRD analysis, we identified the composition of the samples as being Sm0.85Sr0.15MnO3.

3.2. Thermal Expansion Property

Figure 3a–c show the relative length (dL/L) with the temperature increases of SmMnO3, SrMnO3, and Sm0.85Sr0.15MnO3-δ, respectively. SmMnO3 (Figure 3a) and SrMnO3 (Figure 3b) showed positive thermal expansion. Calculating according to the curve, the expansion coefficients were 5.24 × 10−6/K and 12.7 × 10−6/K, respectively. When the temperature was below 360 K, the ceramic Sm0.85Sr0.15MnO3-δ showed a positive thermal expansion of 0.46875 × 10−6/K. As the temperature increased, the ceramic Sm0.85Sr0.15MnO3-δ showed an NTE property in the range of 360 to 873 K. The average linear expansion coefficient was −10.08 × 10−6/K.
Figure 4 shows the high-temperature XRD patterns of ceramic Sm0.85Sr0.15MnO3-δ from RT to 873 K. As the temperature increased, the diffraction peaks of Sm0.85Sr0.15MnO3-δ moved slightly to small angles, except three diffraction peaks (31.54°, 33.79°, and 52.65°) that moved to a large angle.
Figure 5 shows the variation in the Sm0.85Sr0.15MnO3-δ lattice parameters and volume with temperature increases, which was calculated using the powderX software. In a, c in Figure 5, the increase occurred gradually, while in b in Figure 5, it decreased as the temperature increased gradually. We believe that the thermal expansion of Sm0.85Sr0.15MnO3-δ was due to anisotropy. We can see that from RT to 360 K, Sm0.85Sr0.15MnO3-δ showed a positive expansion property. As the temperature increased to 360~873 K, Sm0.85Sr0.15MnO3-δ showed an NTE property with the average linear expansion coefficient of −3.33 × 10−6/K. However, the original calculation of the negative thermal expansion coefficient of Sm0.85Sr0.15MnO3-δ in this temperature range was −10.08 × 10−6/K, according to Figure 3. As can be seen from Figure 2 above, there were pores and microcracks in the crystal. Therefore, we believe that when the temperature rises, the crystal squeezes the open space, namely, these pores and microcracks, which is another reason for the negative thermal expansion.
Table 2 shows the pore size, pore volume, and BET surface area of the sample. The specific surface of the material itself is large, and the general level of adsorption is good. When the pore structure of carbon materials is more complex, it is easy to have a flexible hole, and the pore size becomes larger after gas adsorption. With the doping of Sr2+, oxygen defects are caused, and the gas is adsorbed in the pores. With the further doping of Sr2+, the adsorption oxygen saturation does not change. With the increase in temperature, the gas is sintered out, the b-axis shrinks at the same time, and the pore size becomes smaller, resulting in the negative expansion property.
Figure 6a is the XPS spectrum of Sm0.85Sr0.15MnO3-δ; the characteristic peaks of Sm, Sr, Mn, and O are shown in the figure, respectively. The surface of the sample was free from any pollutants, and element C was used for the calibration of the XPS atlas. Figure 6b,c show the XPS spectra of Mn. In the XPS spectrum, the sample had a bimodal structure, which indicates that the Mn elements on the sample surface existed in two forms: Mn3+ and Mn4+, which led to the oxygen vacancy. The presence of the oxygen vacancy facilitated the movement of electrons between Mn4+ and Mn3+. The oxygen vacancy also led to the shortening of the bond length of the Mn-O bond, which led to lattice distortion and generated internal stress; this reduced the bond angle of Mn-O-Mn and increased the double-exchange effect.

3.3. Discussion

SmMnO3 is a typical manganite perovskite structure. The structure of SmMnO3 is shown in Figure 7. As for the MnO6 octahedron in SmMnO3, the distortion was caused by a change in the length of the Mn-O bond.
There are three kinds of common modes for this change [22,23,24,25,26] as follows. (1) The surface tension contract model Q1, as shown in Figure 8a. Six oxygen atoms of the unit cell move close to or far away from the manganese atom at the same time, making the Mn-O bond length decrease or increase significantly. This model can increase the energy of the system, which is not conducive to the system energy being able to decrease and makes the system extremely unstable in turn. (2) The plane distortion model Q2, as shown in Figure 8b. In a unit, two oxygen atoms in the horizontal plane leave a manganese atom, while the other two oxygen atoms become close to the manganese atom. The location of the two oxygen atoms in the vertical plane remains unchanged. (3) The expansion mode, or inspiratory mode Q3, which is shown in Figure 8c. In a MnO6 octahedron, the two oxygen atoms in the vertical plane leave manganese atoms, while the four oxygen atoms in the horizontal plane become close to the manganese atom simultaneously. For a MnO6 octahedron, the Q1 and Q2 models normally exist. Since the Q1 model is unstable, the distortion of the MnO6 octahedron is mainly the Q2 model, also called the plane distortion model.
We used MnO2, Sm2O3, and SrO as the raw materials to prepare Sm0.85Sr0.15MnO3-δ. In the reaction process, there was a reciprocal transformation between Mn3+ and Mn4+.
When Sr2+ substitutes the Sm3+ in SmMnO3, Sr2+ will occupy the position of Sm3+. To maintain the valence balance, electron transfer occurs in the Mn3+ converting into Mn4+ in Sm0.85Sr0.15MnO3-δ. Additionally, the p electron of O2− will migrate to the orbit of the nearby Mn4+, and the d electron of Mn3+ will migrate to the orbit of the nearby Mn3+. Thus, this mechanism results in the electronic conduction and position exchanges of Mn4+ and Mn3+ ions. The system energy remains unchanged throughout. This process is known as the double exchange [27]. The structure of Mn3+-O2−-Mn4+ forms in the process. However, according to the theory of Zener [28], the route of electron transfer between two Mn3+ changes between Mn3+ and Mn4+. In order to keep the electron transfer between two Mn3+, the magnetic moment between Mn3+ and Mn4+ ions should be parallel to each other. In this situation, it is favorable for there to be more electron transfer between Mn3+ and Mn4+ ions.
According to the analysis of the variable-temperature XRD data, we considered that the thermal property of Sm0.85Sr0.15MnO3-δ might be related to the interaction of the lattice vibration and electron transfer between Mn3+ and Mn4+. As the temperature rose, the lattice vibrated dramatically and Mn4+ converted into Mn3+. Moreover, the electron transfer rate increased between the Mn3+ and Mn4+ ions. The number of Mn3+ ions that can cause the Jahn–Teller [29] effect increased. The oxygen ions in the Mn3+O6 octahedron became slant, or even produced oxygen defects, making the unit cell volumes decrease. From RT to 360 K, the unit cell volume increased. The reason is that the contribution of the lattice vibration to the thermal expansion exceeded that of the MnO6 octahedral distortion and oxygen defects. As the temperature increased, Sm0.85Sr0.15MnO3-δ showed a low positive thermal expansion property, and above 360 K, the unit cell volume decreased. With more Mn4+ ions converting into Mn3+, the Mn3+O6 octahedral distortion was enhanced and oxygen defects occurred. These contributed more to the thermal expansion than the lattice vibration. Therefore, Sm0.85Sr0.15MnO3-δ shows a negative thermal expansion property between 360 K and 873 K.
The DSC and TGA results of ceramic Sm0.85Sr0.15MnO3-δ also support the above statements. Figure 9a presents the DSC curve of ceramic Sm0.85Sr0.15MnO3-δ. In the curve, Sm0.85Sr0.15MnO3-δ has an endothermic peak at about 360 K. This shows that more Mn4+ ions were converted to Mn3+ with the increase in temperature. Thus, Mn3+O6 octahedral distortion was enhanced and oxygen defects occurred. The unit cell volume began to decrease, which is consistent with the results calculated by the high-temperature XRD (seeing b in Figure 5). As electron transfer occurred between the Mn3+ and Mn4+ ions, the amount of Mn4+ decreased, and oxygen ions in the Mn3+O6 octahedron became slant or even produced oxygen defects. The TGA results of Sm0.85Sr0.15MnO3-δ confirm the existence of oxygen defects. In Figure 8b, the weight of the Sm0.85Sr0.15MnO3-δ sample decreased when the temperature increased from RT to 873 K. In addition, the variable-temperature XRD (seeing Figure 5) showed that there was no phase transition with the increase in temperature. As electron transfer occurred between the Mn3+ and Mn4+ ions, Mn4+O6 converted into Mn3+O6 and oxygen defects appeared. Therefore, we consider that the loss of the weight can be ascribed to the oxygen defects.
Moreover, the non-stoichiometric ratio of Sm0.85Sr0.15MnO3-δ caused the mole ratio mismatch of the Sm, Mn, and O atoms. Some lattice vacancies and interstitials existed in the crystal lattice, making the lattice distortion continuous. In the structure analysis, the crystal distortion was found to have a direct impact on the bond length and angle of the MnO6 octahedron. As for ABO3, when we conducted the substitution in the A position with a different ionic radius, especially in the non-stoichiometric ratio manganese perovskite, the size mismatch effects of the A position ion together with lattice space and interstitial caused a difference in the crystal structure. These eventually led to a great change in the lattice parameters and unit cell size [30,31,32].

4. Conclusions

(1)
A novel negative thermal expansion material composed of Sm0.85Sr0.15MnO3-δ was synthesized using the solid-state method with an NTE coefficient of −10.08 × 10−6/K from 360 to 873 K.
(2)
The particles were homogenous spherical or elliptic–spherical particles with a uniform particle size of about 1~2 μm.
(3)
The ceramic Sm0.85Sr0.15MnO3-δ crystallized in an orthorhombic structure with the space group Pbnm. When Sr2+ substituted the Sm3+ in SmMnO3, Sr2+ occupied the position of Sm3+. To maintain the valence balance, electronic transfer occurred in the Mn3+, converting into Mn4+ in Sm0.85Sr0.15MnO3-δ. The Mn3+-O2−-Mn4+ structure formed in the process.
(4)
The thermal property of Sm0.85Sr0.15MnO3-δ is considered to be related to the interaction of the lattice vibration and electron transfer between Mn ions. As the temperature rise, the lattice vibrated dramatically and more Mn3+ converted into Mn4+. Additionally, the electron transfer rate increased between the Mn3+ and Mn4+ ions as the temperatures increased. The number of Mn3+ ions that can cause the Jahn–Teller effect increasesd. The oxygen ions in the Mn3+O6 octahedron became slant or even produced oxygen defects. The contributions of the lattice vibrations and electron transfer between Mn3+ and Mn4+ to the thermal expansion changed with the increasing temperature.
(5)
The pore energy in the sintered body partially absorbed the expansion of the a-axis a and the c-axis; the negative expansion phenomenon can be explained from the perspective of the contraction of the b-axis. The abnormal thermal expansion behavior of the Sm0.85Sr0.15MnO3-δ perovskite system is caused by the presence of pores in the sintered body combined with the negative expansion of the b-axis in the perovskite system.

Author Contributions

Conceptualization, Y.L. (Yucheng Li); Methodology, Y.L. (Yucheng Li); Data curation, Y.L. (Yucheng Li); software, Y.L. (Yucheng Li); validation, Y.L. (Yongtian Li); formal analysis, Y.L. (Yucheng Li); investigation, Y.Z.; resources, Y.Z.; writing—original draft preparation, Y.Z.; writing—review and editing, Y.W.; visualization, Y.W.; supervision, Y.L. (Yongtian Li); project administration, Y.Z. 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

Informed consent was obtained from all subjects involved in the study.

Data Availability Statement

The study did not report any data.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Evans, J.S.O.; Mary, T.A.; Vogt, T.; Subramanian, M.A. Sleight, Negative Thermal Expansion in ZrW2O8 and HfW2O8. Mater. Chem. 1996, 8, 2809–2823. [Google Scholar] [CrossRef]
  2. Sahoo, P.P.; Sumithra, S.; Madras, G. Row, Synthesis, Structure, Negative Thermal Expansion, and Photocatalytic Property of Mo Doped ZrV2O7. Inorg. Chem. 2011, 50, 8774–8781. [Google Scholar] [CrossRef] [PubMed]
  3. Shi, N.; Sanson, A.; Sun, Q.; Fan, L.; Venier, A.; de Souza, D.O.; Xing, X.; Chen, J. Strong Negative Thermal Expansion of Cu2PVO7 in a Wide Temperature Range. Chem. Mater. 2021, 33, 1321–1329. [Google Scholar] [CrossRef]
  4. Shi, N.; Sanso, A.; Venier, A.; Fan, L.; Sun, C.; Xing, X.; Chen, J. Negative and zero thermal expansion in α-(Cu2−xZnx)V2O7 solid solutions. Chem Commun. 2020, 56, 10666–10669. [Google Scholar] [CrossRef] [PubMed]
  5. Liu, X.S.; Li, F.; Song, W.B.; Yuan, B.H.; Cheng, Y.G.; Liang, E.J.; Chao, M.J. Control of reaction processes for rapid synthesis of low-thermal-expansion Ca1−xSrxZr4P6O24 ceramics. Ceram. Int. 2014, 40, 6013–6020. [Google Scholar] [CrossRef]
  6. Oba, Y.; Tadano, T.; Akashi, R.; Tsuneyuki, S. First-principles study of phonon anharmonicity and negative thermal expansion in ScF3. Phys. Rev. Mater. 2019, 3, 033601. [Google Scholar] [CrossRef] [Green Version]
  7. Goodwin, A.L.; Calleja, M.; Conterio, M.J.; Dove, M.T.; Evans, J.S.O.; Keen, D.A.; Peters, L.; Tucker, M.G. Colossal positive and negative thermal expansion in the framework material Ag3(Co(CN)6). Science 2008, 319, 794–797. [Google Scholar] [CrossRef] [Green Version]
  8. Huang, R.J.; Xu, W.; Xu, X.D.; Li, L.F.; Pan, X.Q.; Evans, D. Negative thermal expansion and electrical properties of Mn3Cu0.6NbxGe0.4−xN (x = 0.05–0.25) compounds. Mater. Lett. 2008, 62, 2381–2384. [Google Scholar] [CrossRef]
  9. Li, Y.; Liu, C.; Chao, M.; Zhang, Y.; Li, Y.; Wu, Y. Negative thermal expansion property of Eu0.8Sr0·2MnO3−δ. Results Mater. 2020, 8, 100154. [Google Scholar] [CrossRef]
  10. Liu, Y.; Wang, Z.; Chang, D.; Sun, Q.; Chao, M.; Jia, Y. Charge transfer induced negative thermal expansion in perovskite BiNiO3. Comput. Mater. Sci. 2016, 113, 198–202. [Google Scholar] [CrossRef]
  11. Long, Y.W.; Hayashi, N.; Saito, T.; Azuma, M.; Muranaka, S.; Shimakawa, Y. Temperature-induced A–B intersite charge transfer in an A-site-ordered LaCu3Fe4O12 perovskite. Nature 2009, 458, 60–63. [Google Scholar] [CrossRef] [PubMed]
  12. Liu, Y.Y.; Yuan, B.H.; Cheng, Y.G.; Liang, E.J.; Ge, X.H.; Yuan, H.L.; Zhang, Y.; Guo, J.; Chao, M.J. Phase transition and negative thermal expansion of HfMnMo3O12. Mater. Res. Bull. 2018, 99, 255–259. [Google Scholar] [CrossRef]
  13. Wang, H.; Yang, M.; Chao, M.; Guo, J.; Tang, X.; Jiao, Y.; Liang, E. Negative thermal expansion properties of Cu1.5Mg0.5V2O7. Ceram. Int. 2019, 45, 9814–9819. [Google Scholar] [CrossRef]
  14. Wang, H.; Yang, M.; Chao, M.; Guo, J.; Gao, Q.; Jiao, Y.; Tang, X.; Liang, E. Negative thermal expansion property of β-Cu2V2O7. Solid State Ion. 2019, 343, 115086. [Google Scholar] [CrossRef]
  15. Warne-Lang, V.; Sato, M.; Ozeki, M.; Kadowaki, Y.; Yokoyama, Y.; Katayama, N.; Okamoto, Y.; Takenaka, K. Annealing effects on negative thermal expansion properties of ball-milled β-Cu1.8Zn0.2V2O7 fine particles. Ceram. Int. 2020, 46, 27655–27659. [Google Scholar] [CrossRef]
  16. Rotermel, M.V.; Krasnenko, T.I.; Titova, S.G.; Praynichnikov, S.V. Negative volume thermal expansion of monoclinic Cu2−2xZn2xV2O7 in the temperature range from 93 to 673 K. J. Solid State Chem. 2020, 285, 121221. [Google Scholar] [CrossRef]
  17. Azuma, M.; Chen, W.T.; Seki, H.; Czapski, M.; Olga, S.; Oka, K. Colossal negative thermal expansion in BiNiO3 induced by intermetallic charge transfer. Nat. Commun. 2011, 2, 347. [Google Scholar] [CrossRef] [Green Version]
  18. Hirano, A.; Hirano, F.; Matsumura, T.; Imanishi, N.; Takeda, Y. An anomalous thermal expansion in the perovskite system Gd1−xSrxMnO3 (0 ≤ x ≤ 0.3). Solid State Ion. 2007, 177, 749–755. [Google Scholar] [CrossRef]
  19. Fu, L.J.; Chao, M.J.; Chen, H.; Liu, X.S.; Liu, Y.M.; Yu, J.M.; Liang, E.J.; Li, Y.C.; Xiao, X. Negative thermal expansion property of Er0.7Sr0.3NiO3−δ. Phys. Lett. A 2014, 378, 1909–1912. [Google Scholar] [CrossRef]
  20. Li, Y.; Zhang, Y.; Zhang, N.; Li, Y.; Wu, Y. Negative thermal expansion property of Sm1−xCuxMnO3−δ. J. Mater. Res. Technol. 2021, 12, 2267–2272. [Google Scholar] [CrossRef]
  21. Yamamoto, H.; Imai, T.; Sakai, Y.; Azuma, M. Colossal Negative Thermal Expansion in Electron-Doped PbVO3 Perovskites. Angew. Chem. Int. Ed. 2018, 57, 8170–8173. [Google Scholar] [CrossRef] [PubMed]
  22. Ishizaki, H.; Sakai, Y.; Nishikubo, T.; Pan, Z.; Oka, K.; Yamamoto, H.; Azuma, M. Negative Thermal Expansion in Lead-Free La-Substituted Bi0.5Na0.5VO3. Chem. Mater. 2020, 32, 4832–4837. [Google Scholar] [CrossRef]
  23. Wang, J.; Xu, P.; Yuan, H.; Gao, Q.; Sun, Q.; Liang, E. Negative thermal expansion driven by acoustic phonon modes in rhombohedral Zn2GeO4. Results Phys. 2020, 19, 103531. [Google Scholar] [CrossRef]
  24. Wei, S.; Kong, X.; Wang, H.; Mao, Y.; Chao, M.; Guo, J.; Liang, E. Negative thermal expansion property of CuMoO4. Optik 2018, 160, 61–67. [Google Scholar] [CrossRef]
  25. Mohamed, Z.; Tka, E.; Dhahri, J.; Hlil, E.K. Critical behavior near the paramagnetic to ferromagnetic phase transition temperature in La0.67Sr0.1Ca0.23MnO3 compound. J. Alloys Compd. 2016, 688, 1260–1267. [Google Scholar] [CrossRef]
  26. Khiem, N.V.; Bau, L.V.; Son, L.H.; Phuc, N.X.; Nam, D.N.H. Influence of A-site cation size on the magnetic and transport properties of (Nd1−yYy)0.7Sr0.3MnO3 (0 < y < 0.42). J. Magn. Magn. Mater. 2003, 262, 490–495. [Google Scholar]
  27. Abdulvagidov, B.; Djabrailov, Z.; Abdulvagidov, B.; Kurbakov, A.I. New Universality Class Associated with Jahn-Teller Distortion and Double Exchange. J. Exp. Theor. Phys. 2020, 130, 528–542. [Google Scholar] [CrossRef]
  28. Wu, F.; Ping, Y. Combining Landau Zener Theory and Kinetic Monte Carlo Sampling for Small Polaron Mobility of Doped BiVO4. arXiv 2018, arXiv:1808.02507. [Google Scholar]
  29. Mohanty, S.; Mukherjee, S. Effect of Jahn-Teller distortion on microstructural and dielectric properties of La based double perovskites. J. Alloys Compd. 2022, 892, 162204. [Google Scholar] [CrossRef]
  30. Zhang, Z.; Wang, Y.; Sun, W.; Zhang, X.; Liu, H.; Chen, X.; Zeng, X. Phase transition temperature and negative thermal expansion of Sc-substituted In2(MoO4)3 ceramics. J. Mater. Sci. 2020, 55, 5730–5740. [Google Scholar] [CrossRef]
  31. Wei, X.S.; Zhou, Q.; Yan, X.; Li, X.; Zhu, B. Preparation and properties of Zr2WP2O12 with negative thermal expansion without sintering additives. Process. Appl. Ceram. 2020, 14, 173–180. [Google Scholar]
  32. Sakthipandi, K.; Rajendran, V. Phase transitions of bulk and nanocrystalline La1−xPbxMnO3 (x = 0.3, 0.4, and 0.5) perovskite manganite materials using ultrasonic measurements. Mater. Chem. Phys. 2013, 138, 581–592. [Google Scholar] [CrossRef]
Figure 1. The XRD patterns: (a) Sm0.85Sr0.15MnO3-δ and (b) SmMnO3.
Figure 1. The XRD patterns: (a) Sm0.85Sr0.15MnO3-δ and (b) SmMnO3.
Jcs 06 00156 g001
Figure 2. (a) SEM image of the ceramic Sm0.85Sr0.15MnO3-δ; (b) EDS spectrum corresponding to the SEM image.
Figure 2. (a) SEM image of the ceramic Sm0.85Sr0.15MnO3-δ; (b) EDS spectrum corresponding to the SEM image.
Jcs 06 00156 g002
Figure 3. Relative length change (dL/L) with the temperature of the samples: (a) SmMnO3, (b) SrMnO3, and (c) Sm0.85Sr0.15MnO3-δ.
Figure 3. Relative length change (dL/L) with the temperature of the samples: (a) SmMnO3, (b) SrMnO3, and (c) Sm0.85Sr0.15MnO3-δ.
Jcs 06 00156 g003
Figure 4. XRD patterns of Sm0.85Sr0.15MnO3-δ ceramics at high temperatures.
Figure 4. XRD patterns of Sm0.85Sr0.15MnO3-δ ceramics at high temperatures.
Jcs 06 00156 g004
Figure 5. The variation in the Sm0.85Sr0.15MnO3-δ lattice parameters and volume with temperature increases.
Figure 5. The variation in the Sm0.85Sr0.15MnO3-δ lattice parameters and volume with temperature increases.
Jcs 06 00156 g005
Figure 6. The XPS spectra of Sm0.85Sr0.15MnO3-δ: (a) elemental analysis, (b) Mn2p, (c) Mn3s, and (d) C1s.
Figure 6. The XPS spectra of Sm0.85Sr0.15MnO3-δ: (a) elemental analysis, (b) Mn2p, (c) Mn3s, and (d) C1s.
Jcs 06 00156 g006aJcs 06 00156 g006b
Figure 7. Structure diagram of SmMnO3.
Figure 7. Structure diagram of SmMnO3.
Jcs 06 00156 g007
Figure 8. Three kinds of distortion models for the Mn-O bond (a) The surface tension contract model Q1; (b) The plane distortion model Q2; (c) The expansion mode, or inspiratory mode Q3.
Figure 8. Three kinds of distortion models for the Mn-O bond (a) The surface tension contract model Q1; (b) The plane distortion model Q2; (c) The expansion mode, or inspiratory mode Q3.
Jcs 06 00156 g008
Figure 9. DSC (a) and TGA (b) curve of Sm0.85Sr0.15MnO3-δ.
Figure 9. DSC (a) and TGA (b) curve of Sm0.85Sr0.15MnO3-δ.
Jcs 06 00156 g009
Table 1. Atomic ratio of Sm, Sr, Mn, and O in Sm0.85Sr0.15MnO3-δ by EDS.
Table 1. Atomic ratio of Sm, Sr, Mn, and O in Sm0.85Sr0.15MnO3-δ by EDS.
ElementSmSrMnO
(at.%)14.462.3916.2066.95
Table 2. Pore size, pore volume, and BET surface area of the sample.
Table 2. Pore size, pore volume, and BET surface area of the sample.
SamplePore Size (nm)Pore Volume (cm3/g)BET Surface Area (m2/g)
Sm0.85Sr0.15MnO3-δ15.78420.0025630.6351
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, Y.; Zhang, Y.; Li, Y.; Wu, Y. Negative Thermal Expansion Properties of Sm0.85Sr0.15MnO3-δ. J. Compos. Sci. 2022, 6, 156. https://doi.org/10.3390/jcs6060156

AMA Style

Li Y, Zhang Y, Li Y, Wu Y. Negative Thermal Expansion Properties of Sm0.85Sr0.15MnO3-δ. Journal of Composites Science. 2022; 6(6):156. https://doi.org/10.3390/jcs6060156

Chicago/Turabian Style

Li, Yucheng, Yang Zhang, Yongtian Li, and Yifeng Wu. 2022. "Negative Thermal Expansion Properties of Sm0.85Sr0.15MnO3-δ" Journal of Composites Science 6, no. 6: 156. https://doi.org/10.3390/jcs6060156

Article Metrics

Back to TopTop