A doubling of the superconducting transition temperature occurs in strained SrTiO3 films, which are known to become ferroelectric.
Abstract
The nature of superconductivity in SrTiO3, the first oxide superconductor to be discovered, remains a subject of intense debate several decades after its discovery. SrTiO3 is also an incipient ferroelectric, and several recent theoretical studies have suggested that the two properties may be linked. To investigate whether such a connection exists, we grew strained, epitaxial SrTiO3 films, which are known to undergo a ferroelectric transition. We show that, for a range of carrier densities, the superconducting transition temperature is enhanced by up to a factor of two compared to unstrained films grown under the same conditions. Moreover, for these films, superconductivity emerges from a resistive state. We discuss the localization behavior in the context of proximity to ferroelectricity. The results point to new opportunities to enhance superconducting transition temperatures in oxide materials.
INTRODUCTION
Although SrTiO3 was the first oxide superconductor to be discovered (1), the nature of its superconducting state has been a longstanding subject of debate in the literature (2–10), reflecting in many ways the elusiveness of other families of superconductors, such as the cuprates. A striking feature is that superconductivity already appears at very low carrier densities (2, 11), when the Fermi temperature is lower than the Debye temperature, which is at odds with the Bardeen-Cooper-Schrieffer (BCS) theory. Bulk, undoped SrTiO3 is an incipient ferroelectric for which quantum fluctuations suppress a low-temperature transition to a ferroelectric ground state at the lowest temperatures (12, 13). Several recent theoretical proposals have suggested that a connection between the ferroelectric and superconducting properties exists (14–17), providing strong motivation for developing experimental approaches that search for such a link.
The complex relationship between metallicity and ferroelectricity is one of the main challenges in experimental tests of the role of (incipient) ferroelectricity in the superconducting pairing mechanism of SrTiO3. In particular, free carriers, needed for superconductivity, and ferroelectricity do not easily coexist. For example, sufficiently large concentrations of mobile carriers can screen the splitting of the transverse and longitudinal optical phonon modes (18, 19), which is essential for ferroelectricity in materials such as SrTiO3 (20). Furthermore, conducting samples cannot sustain sufficiently large electric fields that are needed to switch a ferroelectric polarization. Thus, although there are no fundamental reasons why the two properties cannot coexist (21), practically, they are difficult to obtain within a homogeneous material.
Despite these challenges, several recent experiments have reported on the superconductivity of SrTiO3 crystals that were tuned toward ferroelectricity by using approaches that were previously known (22–24) to stabilize ferroelectricity in insulating SrTiO3. In particular, oxygen isotope doping (25) and alloying with Ca (26) were found to result in modest changes in TC (superconducting transition temperatures) of doped single crystals.
Epitaxial coherency strains are well known to stabilize ferroelectricity in SrTiO3 films (27–30). Films under different epitaxial strains thus offer an attractive platform to compare the superconducting properties of films poised to undergo a ferroelectric transition with those that remain paraelectric. Here, we show that doped, compressively strained SrTiO3 films exhibit TC values that are enhanced by a factor of two compared to unstrained films. Moreover, films with enhanced TC exhibit a pronounced upturn in the normal state resistivity with decreasing temperature, which is highly unusual. In contrast, films with higher carrier densities remain metallic and do not exhibit enhanced TC, although they are strained by the same amount. We discuss the implications of the results with regard to the connection between superconductivity and (incipient) ferroelectricity in SrTiO3.
RESULTS
Electrical measurements were carried out on epitaxial SrTiO3 films, which were grown on (001) LSAT [(LaAlO3)0.3(Sr2AlTaO6)0.7] and SrTiO3 substrates using a hybrid molecular beam epitaxy (MBE) technique (31, 32). The films were doped with different amounts of Sm+3 to obtain a range of carrier densities.
Figure 1A shows the resistivities (ρ) of strained Sm:SrTiO3 films grown on LSAT with different carrier densities, measured between 300 and 2 K. Here, the resistivities were determined from the sheet resistance and film thickness (200 nm) measured by cross-sectional transmission electron microscopy. The highest doped film (n3D = 2.8 × 1020 cm−3 at 300 K, where n3D is the carrier density determined from Hall measurements) shows metallic behavior, , down to 1.8 K. In contrast, lower doped films (n3D = 2 × 1019, 6 × 1019, and 1.4 × 1020 cm−3) exhibit a crossover to upon lowering the temperature. The transition temperatures, defined as, for films with n3D = 6 × 1019 and 1.4 × 1020 cm−3 are ~110 and ~60 K, respectively (see arrows). The lowest doped film with n3D = 2 × 1019 cm−3 not only demonstrates an upturn in resistivity at low temperatures but also shows a more complicated temperature dependence. All films exhibit an abrupt decrease in the Hall carrier density upon lowering the temperature (Fig. 1B). Roughly the same mobile carrier density (4 × 1019 to 7 × 1019 cm−3) is lost in all films.
Figure 2 shows ρ between 1 K and 10 mK. All films become superconducting at low temperatures. Taking TC as the temperature for which ρ corresponds to, where ρn is the normal state value and e is Euler’s number, the values for TC are 0.37, 0.60, 0.67, and 0.16 K for films with n3D = 2 × 1019, 6 × 1019, 1.4 × 1020, and 2.8 × 1020 cm−3, respectively.
Figure 3 compares the TC values for the films shown in Fig. 2 with those of other SrTiO3 samples with similar carrier densities reported in the literature, including SrTiO3 crystals that were tuned toward ferroelectricity using the approaches mentioned in the Introduction (11, 25, 26, 33, 34). The TC of the Sm:SrTiO3 film with the highest doping density, which remains metallic, is comparable to previous reports. In contrast, films with lower doping densities have a substantially increased TC, which reaches almost a factor of two near the peak of the superconducting dome (“optimal doping”). Thus, the enhancement of TC is seen for underdoped and optimally doped films but not on the overdoped side of the superconducting dome of SrTiO3. For direct comparison with unstrained films grown by the same MBE method and TC defined in the same way, Fig. 3 also shows TC of unstrained Sm- and La-doped SrTiO3 thin films grown on SrTiO3 substrates, which show TC similar to La-doped SrTiO3 crystals reported in the literature (34). All samples shown in Fig. 3, except for the three lower doped films on LSAT, which have enhanced TC, exhibit metallic behavior.
Figure 4 shows the superconducting upper critical magnetic field, HC2, as a function of temperature. Data follow the relation, where t = T/TC, which is shown as dashed lines, similar to the behavior observed in La-doped films (35). The values of HC2 (i.e., 1.75 T at 50 mK for the film with n3D = 6 × 1019 cm−3) are comparable with those reported for La-doped SrTiO3 films (35). Table 1 lists quantities derived from the data, including the values for the BCS superconducting gap, Δ = 1.75 kBTC, where kB is Boltzmann’s constant, and the superconducting coherence length, , where ϕ0 is the magnetic quantum flux. Here, films with n3D = 2 × 1019 and 1.4 × 1020 cm−3, despite having different critical temperatures, show relatively similar HC2 at low temperatures. Thus, TC peaks near n3D = 1.4 × 1020 cm−3, while HC2 peaks near n3D = 6 × 1019 cm−3. The behavior of HC2 might be related to the multiband nature of this superconductor (36) and requires further investigations beyond the scope of the present study.
Table 1. Values for TC, HC2, Δ, and ξ for Sm-doped SrTiO3 films on LSAT with different carrier densities.
n3D at 300 K (cm−3) | TC (K) | HC2 at 50 mK (T) | Δ (μeV) | ξ (nm) |
2 × 1019 | 0.37 | 0.83 | 56 | 18 |
6 × 1019 | 0.6 | 1.75 | 90 | 14 |
1.4 × 1020 | 0.67 | 0.91 | 100 | 19 |
2.8 × 1020 | 0.16 | 0.07 | 20 | 67 |
DISCUSSION
To briefly summarize the results, our main findings are as follows: (i) There is an enhancement in TC of compressively strained SrTiO3 films up to a factor of two, and (ii) the increase in TC depends on the carrier density and (iii) it is connected to the presence of a crossover in the resistivity in the normal state to at low temperatures.
The first important conclusion from these results is that the increase in TC is caused not simply by the epitaxial strain, such as strain-induced modification of phonon modes. If this were the case, then all films, independent of their carrier density, should exhibit enhanced TC, because they are all under the same epitaxial film strain. Instead, the carrier density dependence of the enhancement in TC and the resistivity upturn point to a direct connection between a (proximal) ferroelectric state and the modified superconducting properties.
As mentioned above, SrTiO3 films grown on LSAT substrates are known to transition to a ferroelectric state as a result of the epitaxial coherency strain (27–30). Even SrTiO3 films containing substantial amounts of carriers (on the order of 1019 cm−3) were found to undergo a ferroelectric transition around 140 K (29). Thus, the films in the present study are, at minimum, in proximity to a ferroelectric transition.
Beyond this, the observed crossover in the temperature dependence of the resistivity also hints at the emerging ferroelectric or polar nature of the films. The behavior is in marked contrast to unstrained SrTiO3 films grown on SrTiO3 [Fig. 3; see also (37) for more data] or bulk crystals of SrTiO3 (38), which remain metallic even with orders of magnitude lower carrier densities. The reason is the high dielectric constant of SrTiO3. The Mott criterion (39) for metallic behavior, , where aB is the Bohr radius, is thus easily exceeded, even if we assume a substantially reduced dielectric constant (~1000), which is more typical for ferroelectric SrTiO3 (29). While carrier localization can be caused by disorder or traps, unstrained films with similar dopant densities and disorder remain metallic. Screening of a polar charge is a reasonable explanation for the observed localization of a fixed amount of the mobile charge density (Fig. 1B). Remotely doped ferroelectric BaTiO3 films undergo a transition that is very similar to the one observed here (40). In contrast, TC is not increased at higher carrier densities on the overdoped side of the superconducting dome, which remains metallic. At high carrier densities, we expect the long-range interactions needed for ferroelectricity to be screened (20). While there is currently no agreement in the literature as to the pair-breaking mechanism that causes TC to decrease on the overdoped side, it is expected to play a role here as well.
In summary, our experimental results, especially the carrier density dependence of the observed TC enhancement, should be of interest for testing the different theoretical models that have been proposed in the literature that relate superconductivity in SrTiO3 to a (proximal) ferroelectric state (14–17). Independent of the precise mechanism, the results point to opportunities to enhance TC by searching for superconducting oxides that are in proximity to ferroelectricity. It would also be interesting to explore whether similar enhancement could be obtained in superconducting/ferroelectric composite structures.
Last, we would like to note that, during the preparation of this article, we became aware of a recent report of increased TC in SrTiO3 crystals under uniaxial tensile strain (41). The increase in TC was attributed to a strain-induced modification of phonon modes, suggesting that other promising approaches to enhance TC exist for SrTiO3.
MATERIALS AND METHODS
Epitaxial SrTiO3 films, which were doped with Sm+3, were grown by a hybrid MBE technique described elsewhere (31, 32). The LSAT substrate temperature was 900°C (thermocouple reading), and the growth rate was ~130 nm/hour. The carrier density scaled with the Sm flux during growth, consistent with all Sm dopants acting as donors (+3 formal valence state). The lattice mismatch results in compressive in-plane film strains (~1%), and lightly doped films are known to show a ferroelectric transition at ~140 K on this substrate (29). The film thickness was ~200 nm. This thickness is below the critical thickness of SrTiO3 on LSAT (42) while being sufficiently thick to avoid substantial carrier depletion from the well-known surface depletion of SrTiO3 (43). A combination of high-resolution x-ray diffraction (see fig. S1) and reflection high-energy electron diffraction oscillations was used to calibrate the film thickness. Laue thickness fringes were visible in x-ray diffraction and confirmed the film thickness. Reciprocal space mapping (see fig. S2) was carried out around the 113 reflection of SrTiO3 and used to determine the in- and out-of-plane lattice parameters of the film and to confirm that the films remained coherently strained to the LSAT. Cross-sectional high-angle annular dark-field (HAADF) imaging in scanning transmission electron microscopy (STEM) was also used to determine the film thickness and structural quality (fig. S3). We also investigated unstrained Sm-doped SrTiO3 and La-doped SrTiO3 films on (001) SrTiO3 substrates grown under the same conditions.
Temperature-dependent measurements of the longitudinal and Hall resistances were carried out using a Quantum Design Physical Property Measurement System. The Hall carrier densities extracted from the Hall measurements, n3D = −1/(teRH), where t is the thickness of the SmxSr1−xTiO3 thin film, e is the electron charge, and RH is the Hall coefficient RH = dRxy/dB, were extracted from linear fits to the transverse resistance RH(B) with the magnetic field (B). Magnetotransport measurements below 1 K were carried out in a dilution refrigerator (Triton, Oxford Instruments Group). Transport measurements were carried out in van der Pauw geometry with square-shaped samples (5 mm × 5 mm). Ohmic contacts (40-nm Ti/400-nm Au) were deposited on the sample corners (<0.5 mm × 0.5 mm) through a shadow mask using an electron beam evaporation. The superconductivity measurements were carried out using a lock-in amplifier (SR830, Stanford Research Systems) in AC mode with an excitation current of 1 μA and a frequency of 33.33 Hz. The critical field was obtained from measuring the longitudinal resistance while sweeping an out-of-plane magnetic field at different temperatures (see fig. S4).
Supplementary Material
Acknowledgments
Funding: This work was supported by the NSF (grant nos. 1740213 and 1729489) and a Vannevar Bush Faculty Fellowship of the Department of Defense (grant no. N00014-16-1-2814). The dilution fridge used in the measurements was funded through the Major Research Instrumentation program of the U.S. NSF (award no. DMR 1531389). This work made use of the MRL Shared Experimental Facilities, which are supported by the MRSEC Program of the U.S. NSF under award no. DMR 1720256. Author contributions: K.A. grew the samples and performed electrical measurements and data analysis. Y.L. and L.G. assisted with electrical measurements and analysis. S.S.-R. performed transmission electron microscopy measurements, and W.W. assisted with x-ray diffraction studies. K.A. and S.S. wrote the manuscript. Competing interests: The authors declare that they have no competing interests. Data and materials availability: All data needed to evaluate the conclusions in the paper are present in the paper and/or the Supplementary Materials. Additional data related to this paper may be requested from the authors.
SUPPLEMENTARY MATERIALS
REFERENCES AND NOTES
- 1.Schooley J. F., Hosler W. R., Cohen M. L., Superconductivity in semiconducting SrTiO3. Phys. Rev. Lett. 12, 474–475 (1964). [Google Scholar]
- 2.Appel J., Soft-mode superconductivity in SrTiO3−x. Phys. Rev. 180, 508–516 (1969). [Google Scholar]
- 3.Binnig G., Baratoff A., Hoenig H. E., Bednorz J. G., Two-band superconductivity in Nb-doped SrTiO3. Phys. Rev. Lett. 45, 1352–1355 (1980). [Google Scholar]
- 4.Takada Y., Theory of superconductivity in polar semiconductors and its application to N-type semiconducting SrTiO3. J. Phys. Soc. Jpn. 49, 1267–1275 (1980). [Google Scholar]
- 5.Baratoff A., Binnig G., Mechanism of superconductivity in SrTiO3. Phys. B & C 108, 1335–1336 (1981). [Google Scholar]
- 6.van der Marel D., van Mechelen J. L. M., Mazin I. I., Common Fermi-liquid origin of T2 resistivity and superconductivity in n-type SrTiO3. Phys. Rev. B 84, 205111 (2011). [Google Scholar]
- 7.Gor’kov L. P., Phonon mechanism in the most dilute superconductor n-type SrTiO3. Proc. Natl. Acad. Sci. U.S.A. 113, 4646–4651 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Ruhman J., Lee P. A., Superconductivity at very low density: The case of strontium titanate. Phys. Rev. B 94, 224515 (2016). [Google Scholar]
- 9.Swartz A. G., Inoue H., Merz T. A., Hikita Y., Raghu S., Devereaux T. P., Johnston S., Hwang H. Y., Polaronic behavior in a weak-coupling superconductor. Proc. Natl. Acad. Sci. U.S.A. 115, 1475–1480 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Thiemann M., Beutel M. H., Dressel M., Lee-Hone N. R., Broun D. M., Fillis-Tsirakis E., Boschker H., Mannhart J., Scheffler M., Single-gap superconductivity and dome of superfluid density in Nb-doped SrTiO3. Phys. Rev. Lett. 120, 237002 (2018). [DOI] [PubMed] [Google Scholar]
- 11.Lin X., Zhu Z., Fauqué B., Behnia K., Fermi surface of the most dilute superconductor. Phys. Rev. X 3, 021002 (2013). [Google Scholar]
- 12.Müller K. A., Burkard H., SrTiO3: An intrinsic quantum paraelectric below 4 K. Phys. Rev. B 19, 3593–3602 (1979). [Google Scholar]
- 13.Zhong W., Vanderbilt D., Effect of quantum fluctuations on structural phase transitions in SrTiO3 and BaTiO3. Phys. Rev. B 53, 5047–5050 (1996). [DOI] [PubMed] [Google Scholar]
- 14.Edge J. M., Kedem Y., Aschauer U., Spaldin N. A., Balatsky A. V., Quantum critical origin of the superconducting dome in SrTiO3. Phys. Rev. Lett. 115, 247002 (2015). [DOI] [PubMed] [Google Scholar]
- 15.Rowley S. E., Spalek L. J., Smith R. P., Dean M. P. M., Itoh M., Scott J. F., Lonzarich G. G., Saxena S. S., Ferroelectric quantum criticality. Nat. Phys. 10, 367–372 (2014). [Google Scholar]
- 16.S. E. Rowley, C. Enderlein, J. F. de Oliveira, D. A. Tompsett, E. B. Saitovitch, S. S. Saxena, G. G. Lonzarich, Superconductivity in the vicinity of a ferroelectric quantum phase transition. arXiv:1801.08121 [cond-mat.supr-con] (24 January 2018).
- 17.Dunnett K., Narayan A., Spaldin N. A., Balatsky A. V., Strain and ferroelectric soft-mode induced superconductivity in strontium titanate. Phys. Rev. B 97, 144506 (2018). [Google Scholar]
- 18.Mooradia A., Wright G. B., Observation of the interaction of plasmons with longitudinal optical phonons in GaAs. Phys. Rev. Lett. 16, 999–1001 (1966). [Google Scholar]
- 19.Steigmeier E. F., Harbeke G., Soft phonon mode and ferroelectricity in GeTe. Solid State Commun. 8, 1275–1279 (1970). [Google Scholar]
- 20.Zhong W., King-Smith R. D., Vanderbilt D., Giant LO-TO splittings in perovskite ferroelectrics. Phys. Rev. Lett. 72, 3618–3621 (1994). [DOI] [PubMed] [Google Scholar]
- 21.Filippetti A., Fiorentini V., Ricci F., Delugas P., Iñiguez J., Prediction of a native ferroelectric metal. Nat. Commun. 7, 11211 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Itoh M., Wang R., Inaguma Y., Yamaguchi T., Shan Y.-J., Nakamura T., Ferroelectricity induced by oxygen isotope exchange in strontium titanate perovskite. Phys. Rev. Lett. 82, 3540–3543 (1999). [Google Scholar]
- 23.Uwe H., Sakudo T., Stress-induced ferroelectricity and soft phonon modes in SrTiO3. Phys. Rev. B 13, 271–286 (1976). [Google Scholar]
- 24.Bednorz J. G., Müller K. A., Sr1−xCax TiO3: An XY quantum ferroelectric with transition to randomness. Phys. Rev. Lett. 52, 2289–2292 (1984). [Google Scholar]
- 25.Stucky A., Scheerer G. W., Ren Z., Jaccard D., Poumirol J.-M., Barreteau C., Giannini E., van der Marel D., Isotope effect in superconducting n-doped SrTiO3. Sci. Rep. 6, 37582 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Rischau C. W., Lin X., Grams C. P., Finck D., Harms S., Engelmayer J., Lorenz T., Gallais Y., Fauqué B., Hemberger J., Behnia K., A ferroelectric quantum phase transition inside the superconducting dome of Sr1−xCaxTiO3−δ. Nat. Phys. 13, 643–648 (2017). [Google Scholar]
- 27.Haeni J. H., Irvin P., Chang W., Uecker R., Reiche P., Li Y. L., Choudhury S., Tian W., Hawley M. E., Craigo B., Tagantsev A. K., Pan X. Q., Streiffer S. K., Chen L. Q., Kirchoefer S. W., Levy J., Schlom D. G., Room-temperature ferroelectricity in strained SrTiO3. Nature 430, 758–761 (2004). [DOI] [PubMed] [Google Scholar]
- 28.Pertsev N. A., Tagantsev A. K., Setter N., Phase transitions and strain-induced ferroelectricity in SrTiO3 epitaxial thin films. Phys. Rev. B 61, R825–R829 (2000). [Google Scholar]
- 29.Verma A., Raghavan S., Stemmer S., Jena D., Ferroelectric transition in compressively strained SrTiO3 thin films. Appl. Phys. Lett. 107, 192908 (2015). [Google Scholar]
- 30.Haislmaier R. C., Engel-Herbert R., Gopalan V., Stoichiometry as key to ferroelectricity in compressively strained SrTiO3 films. Appl. Phys. Lett. 109, 032901 (2016). [Google Scholar]
- 31.Jalan B., Engel-Herbert R., Wright N. J., Stemmer S., Growth of high-quality SrTiO3 films using a hybrid molecular beam epitaxy approach. J. Vac. Sci. Technol. A 27, 461–464 (2009). [Google Scholar]
- 32.Son J., Moetakef P., Jalan B., Bierwagen O., Wright N. J., Engel-Herbert R., Stemmer S., Epitaxial SrTiO3 films with electron mobilities exceeding 30,000 cm2 V−1 s−1. Nat. Mater. 9, 482–484 (2010). [DOI] [PubMed] [Google Scholar]
- 33.Koonce C. S., Cohen M. L., Schooley J. F., Hosler W. R., Pfeiffer E. R., Superconducting transition temperatures of semiconducting SrTiO3. Phys. Rev. 163, 380–390 (1967). [Google Scholar]
- 34.Suzuki H., Bando H., Ootuka Y., Inoue I. H., Yamamoto T., Takahashi K., Nishihara Y., Superconductivity in single-crystalline Sr1-xLaxTiO3. J. Phys. Soc. Jpn. 65, 1529–1532 (1996). [Google Scholar]
- 35.Olaya D., Pan F., Rogers C. T., Price J. C., Superconductivity in La-doped strontium titanate thin films. Appl. Phys. Lett. 84, 4020–4022 (2004). [Google Scholar]
- 36.Lin X., Bridoux G., Gourgout A., Seyfarth G., Kramer S., Nardone M., Fauqué B., Behnia K., Critical doping for the onset of a two-band superconducting ground state in SrTiO3−δ. Phys. Rev. Lett. 112, 207002 (2014). [Google Scholar]
- 37.Cain T. A., Kajdos A. P., Stemmer S., La-doped SrTiO3 films with large cryogenic thermoelectric power factors. Appl. Phys. Lett. 102, 182101 (2013). [Google Scholar]
- 38.Spinelli A., Torija M. A., Liu C., Jan C., Leighton C., Electronic transport in doped SrTiO3: Conduction mechanisms and potential applications. Phys. Rev. B 81, 155110 (2010). [Google Scholar]
- 39.N. F. Mott, Metal-Insulator Transitions (Taylor & Francis, ed. 2, 1990). [Google Scholar]
- 40.Raghavan S., Zhang J. Y., Shoron O. F., Stemmer S., Probing the metal-insulator transition in BaTiO3 by electrostatic doping. Phys. Rev. Lett. 117, 037602 (2016). [DOI] [PubMed] [Google Scholar]
- 41.C. Herrera, J. Cerbin, K. Dunnett, A. V. Balatsky, I. Sochnikov, arXiv:1808.03739 [cond-mat.supr-con] (11 August 2018).
- 42.Wang T., Ganguly K., Marshall P., Xu P., Jalan B., Critical thickness and strain relaxation in molecular beam epitaxy-grown SrTiO3 films. Appl. Phys. Lett. 103, 212904 (2013). [Google Scholar]
- 43.Ohtomo A., Hwang H. Y., Surface depletion in doped SrTiO3 thin films. Appl. Phys. Lett. 84, 1716–1718 (2004). [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.