We report the epitaxial integration of defect-induced room temperature ferromagnetic insulators, Cr2O3 and MgO, with topological insulators Bi2Se3 on c-sapphire substrate by pulsed laser deposition. The structural, magnetic, and magnetotransport properties of ∼15 nm Bi2Se3 thin films are investigated on each template. The lattice misfits of Cr2O3/Bi2Se3 and MgO/Bi2Se3 are ∼16% and ∼39%, respectively, where the critical thickness for pseudomorphic growth is less than one monolayer. The insulating behavior is more pronounced due to the additional scattering of the surface states of the Bi2Se3 layer by interfacing with MgO and Cr2O3. The weak antilocalization effect from the surface states is clearly suppressed, accounting for the presence of magnetic bottom layers. This work demonstrates an effective way to study the emergence of a ferromagnetic phase in topological insulators by the magnetic proximity effect in Bi2Se3, a step toward unveiling their exotic properties.

Topological insulators (TIs) have received considerable attention due to their novel properties arising from strong spin-orbit coupling and massless Dirac-cone-like surface states protected by time reversal symmetry (TRS).1,2 The special kind of surface states is expected to exhibit various unique quantum phenomenon, which may bring revolutionary developments in low power electronics and topological quantum computation. While such unique systems offer nontrival surface states that can be utilized to perform dissipationless spin transport, it is equally important to break TRS of TIs to create a variety of exotic topological effects including the half-integer quantum Hall effect,3 the topological magnetoelectric effect,4 and the magnetic monopole.5 One major drawback for the implementation of TIs into real electronics devices, such as field effect transistors, is their linear energy spectra, which normally allows incident electrons to pass through a potential barrier via Klein tunneling without reflection, leading to an appreciable off-current and thus a poor on/off current ratio.6,7 Therefore, opening a surface energy gap and generating of massive surface carriers by breaking TRS is the key for both fundamental physics research and new materials displaying exotic phenomena aimed at technological applications.

Two routes for breaking TRS have been developed by introducing ferromagnetic ordering within TIs. The first involves doping the TI host with specific elements, for example, by which ferromagnetism has been observed in Mn-doped Bi2Te3 single crystals8,9 and Mn- and Cr-doped Bi2Se3 thin films.10,11 However, it is difficult to separate the surface and the bulk phases using this approach. Furthermore, doping of magnetic elements inevitably introduces crystal defects, magnetic scattering centers, and impurity states in the insulating gap, which are detrimental to mobility and the transport of surface states in TIs.12–14 In contrast, the second route is based on introducing ferromagnetism to the TI surface by proximity to a ferromagnetic insulator (FMI). Suitable FMIs have the potential to achieve strong and uniform exchange coupling via contact with TIs without significant spin-dependent random scattering of helical carriers on magnetic atoms.15–17 To date, the prototype three dimensional (3D) TI material, Bi2Se3, has drawn much attention for the study of proximity effect with FMIs. Several FMI materials have been investigated to interface with Bi2Se3, such as EuS, GdN, and Y3Fe5O12 (YIG). The reference for low ferromagnetic transition temperature (Tc) of these materials, however, minimizes device applications, for example, the Tc for Bi2Se3/EuS, Bi2Se3/GdN, and Bi2Se3/YIG are ∼15.7 K, ∼13 K, and 130 K, respectively.18–21 

Here, we employ MgO and Cr2O3 controllable defect-induced room temperature FMI materials, as substrates for Bi2Se3 growth and demonstrate proximity-induced ferromagnetism in Bi2Se3. The MgO and Cr2O3 are selected as the FMI templates not only based on processing perspectives related to epitaxial thin film growth but also for device applications. Using domain matching epitaxy,22 the misfit strain is relieved within a couple of monolayers as a result of the large lattice misfit; the lattice misfit for MgO/Bi2Se3 is ∼16% and for Cr2O3/Bi2Se3 is ∼39%. As a result, the critical thickness of pseudomorphic growth is less than one monolayer and the misfit strain can be engineered and confined near the interface, with the rest of the film grown free of lattice misfit strain. From the practical point of view, both MgO and Cr2O3 are considered potential candidates for tunnel barriers in magnetic tunnel junctions, where the contact resistance is spin dependent and becomes comparable to the spin independent resistance of the normal metal.23–25 In addition to being crucial components for nonvolatile magnetoresistive random-access memory and low-noise magnetic sensors based on the tunneling magnetoresistance (MR) effect,26 MgO and Cr2O3 show room-temperature ferromagnetism (RTFM), associated with a strain-dependent defect-mediated mechanism. Li et al.27 showed that undoped MgO thin films grown by pulsed laser deposition (PLD) have the RTFM signature and that the ferromagnetism exhibits strong correlation between magnesium vacancies (VMg) and the crystallinity. They found that reduced crystallinity increases the ferromagnetic spin-order of MgO thin films due to increased VMg. As for Cr2O3, though bulk Cr2O3 is an antiferromagnetic with TN = 307 K, He et al.28 showed the existence of a roughness-insensitive ferromagnetic state at the Cr2O3 (0001) surface, and confirmed its spin-polarized property using experimental and theoretical approaches. Later, Punugupati et al.29 demonstrated that epitaxial Cr2O3 thin films exhibit ferromagnetic-like hysteresis loops with high saturation and finite coercivity up to 400 K due to oxygen related defects whose concentration is controlled by the strain present in the films.

In this research, we present structural, magnetic, and magnetotransport characterization of TI/FMI heterostructures. Two hybrid structures are studied based on proximity-induced ferromagnetism in a TI: Bi2Se3/Cr2O3/c-sapphire and Bi2Se3/MgO/c-sapphire. The key advantage here is that the TRS breaking occurs mostly at the TI/FMI interface, rather than affecting the majority of bulk states. Furthermore, the approach avoids the creation of secondary phases, clusters or defect agglomerations, which result from the doping 3d transition metal impurities into TIs.

All samples presented here were grown using PLD (λ = 248 nm and τ = 25 ns). Three heterostructures were studied: Bi2Se3/Cr2O3/c-sapphire, Bi2Se3/MgO/c-sapphire, and Bi2Se3/c-sapphire, which are referred hereafter as Cr2O3-bilayer, MgO-bilayer, and control sample, respectively. This enables us to compare the structural, magnetic, and transport properties of Bi2Se3 thin films grown on a non-magnetic substrate to those grown on FMI substrates. The Cr2O3 film was deposited from a Cr target that was held at 650 °C under an oxygen partial pressure of 5 × 10−2 Torr with laser energy density ∼3.2 J/cm2. The MgO thin film was deposited at the same temperature under an oxygen partial pressure of 5 × 10−5 Torr along with∼2.8 J/cm2 laser energy density. The detailed Bi2Se3 growth conditions have been reported previously,30 using alternating laser pulses; here, a 1:1 ratio of Bi2Se8: Se pulses was used at 150 °C under 3 × 10−1 Torr Ar pressure, to reduce the formation of Se vacancies. All Bi2Se3 films have the same thickness, ∼15 nm, well above the 6 nm threshold, below which the wavefunctions of the top and bottom surfaces overlap substantially, resulting in a hybridization gap near the Dirac point where the transport characteristics are similar to those brought by strong magnetic interactions.31–33 The thicknesses of the Cr2O3 and MgO films were around 100 nm.

To characterize the crystal structure, x-ray diffraction (XRD) was performed using a Rigaku X-ray diffractometer with CuKa1 radiation (λ = 1.54056 A °). Phi-scans were performed by Philips X-Pert Pro X-ray diffractometer equipped with a high resolution goniometer to investigate the crystallographic relationship between substrate and grown layer. Microstructural and growth characteristics of the heterostructures were investigated using a JEOL-2010 field emission transmission electron microscope (TEM), operating at 200 kV with a point-to-point resolution of 1.8 A, and, with high-angle annular dark field scanning transmission electron microscopy (HAADF-STEM) using a FEI probe-corrected Titan G2 60–300 operated at 200 kV. Magnetotransport characterization was performed on 10 mm × 10 mm thin films using a standard four-probe method in the van der Pauw geometry in magnetic fields up to 5 T in an Ever Cool Quantum Design Physical Property Measurement System (PPMS) system with a base temperature as low as 2 K. A custom-built gold-coated pogo pin setup was used to make contacts and ensures the same distance between the contacts on the films. Angular-dependence measurements were performed with a horizontal rotator in the PPMS. The magnetic properties were measured by applying magnetic field parallel and perpendicular to the Bi2Se3 (003 n) surface using a Super Conducting Quantum Interference Device.

Fig. 1(a) displays a XRD θ-2θ pattern of a Cr2O3-bilayer sample. Both the Cr2O3 and Bi2Se3 layers exhibit c-axis preferred orientation crystallographic peaks, suggesting films are highly textured. To confirm the epitaxial growth of the films and establish the epitaxial relationships, XRD in-plane Φ-scans were used. Fig. 1(b) shows the Φ-scans obtained by exciting (104) lattice planes of Cr2O3 and sapphire and (015) planes of Bi2Se3. The three fold symmetry illustrates that the Cr2O3 is rhombohedral and grows epitaxially on c-sapphire without any in-plane rotation. The presence of Bi2Se3 six peaks indicates that the film is epitaxial as a result of the existence of two domains. A schematic showing the alignment of two different domain variants of Bi2Se3 on a Cr2O3 template is seen in Fig. 1(c). In one case, the Bi2Se3 basal plane is aligned with that of Cr2O3 and in the other there is a 60°/180° rotation. The epitaxial relationships are written as (0001)Bi2Se3||(0001)Cr2O3||(0001)Al2O3 and [2-1-10]Bi2Se3||[2-1-10]Cr2O3|| [2-1-10]Al2O3 (or) [2-1-10]Bi2Se3|| [11-20]Cr2O3|| [11-20]Al2O3.

FIG. 1.

Typical XRD patterns of Bi2Se3/Cr2O3/c-sapphire heterostructures: (a) out-of-plane θ-2θ patterns, (b) in-plane Φ-scan patterns excited from the Bi2Se3 {015}, Cr2O3, and c-sapphire {104} planes, and (c) schematic showing the alignment of two different domain variants of Bi2Se3 on the Cr2O3 template.

FIG. 1.

Typical XRD patterns of Bi2Se3/Cr2O3/c-sapphire heterostructures: (a) out-of-plane θ-2θ patterns, (b) in-plane Φ-scan patterns excited from the Bi2Se3 {015}, Cr2O3, and c-sapphire {104} planes, and (c) schematic showing the alignment of two different domain variants of Bi2Se3 on the Cr2O3 template.

Close modal

The out-of-plane XRD θ-2θ pattern of the MgO-bilayer sample is shown in Fig. 2(a). The MgO is completely orientated in the (111) plane, and Bi2Se3 exists only the (003 n) reflections, suggesting a highly textured single-phase heterostructure. The in-plane orientation relationship between Bi2Se3, MgO, and c-sapphire substrate was established by Φ-scan; the scan results are shown in Fig. 2(b). It can be seen that the Φ-scan for (200) planes of MgO results in six sharp peaks, confirming the epitaxial growth. The presence of these reflections is attributed to the threefold symmetry of MgO film growth which is epitaxial with two domain types that have 60° in-plane rotation with respect to each other about the [111] growth direction. The occurrence of two in-plane orientations is explained by stacking sequences in the [111] direction, which make up a face centered cubic (fcc) structure. Any fcc on a hexagonal (0001) or rhombohedral (001) systems, when grown in the [111] direction, has two possible orientations. An fcc structure is described as having an ABCABC… type stacking in the closed-packed direction. During the nucleation on the closed packed plane (cpp) of Al2O3, the first monolayer plane sets in one orientation, which is the same everywhere if the substrate surface has no steps. Fixing this first monolayer cpp as a reference and denoting it as A, the next cpp monolayer grows as a B or a C layer. Upon further growth in the [111] direction, the fcc can stack as either ABCABC… or ACBACB…. These two possibilities are random and manifest themselves in two different fcc with a 60° in-plane rotation with respect to each other. Here, all constituent layers within the heterostructure were found to be epitaxial with the relationships: (0001)Bi2Se3||(111)MgO||(0001)Al2O3 and [10-10]Bi2Se3||[-110]MgO||[10-10]Al2O3; a schematic showing the relative orientations between each layer is presented in Fig. 2(c).

FIG. 2.

Typical XRD patterns of Bi2Se3/MgO/c-sapphire heterostructures: (a) out-of-plane θ-2θ patterns, (b) in-plane Φ-scan patterns excited from the Bi2Se3 {015}, MgO {200}, and c-sapphire {104} planes, and (c) schematic showing the relative orientations between each layer.

FIG. 2.

Typical XRD patterns of Bi2Se3/MgO/c-sapphire heterostructures: (a) out-of-plane θ-2θ patterns, (b) in-plane Φ-scan patterns excited from the Bi2Se3 {015}, MgO {200}, and c-sapphire {104} planes, and (c) schematic showing the relative orientations between each layer.

Close modal

Detailed microstructural analysis was performed by HRTEM and HAADF-STEM on both Cr2O3 and MgO–bilayer samples. All the images present here were taken in the c-sapphire [11–20] zone axis. Figs. 3(a) and 3(b) show the interfaces within Cr2O3/c-sapphire and Bi2Se3/Cr2O3, indicating that both are clean, sharp, and reaction free. The fast Fourier transform (FFT) diffraction patterns in the insets confirm the highly epitaxial growth of the Cr2O3 layer. The atomically ordered structure of Bi2Se3 is observed in Fig. 3(b) where the horizontal arrows and the kinks in the zig-zag lines identify twin boundaries which are in agreement with the Φ-scan results. The Bi2Se3 layered structure, formed by Bi and Se, is stacked along the c-direction in five layer packets Se1-Bi-Se2-Bi-Se1 that connect to each other by weak van der Waals bonds that are also shown in Fig. 3(b), indicating the high quality of Bi2Se3.

FIG. 3.

(a) HRTEM image of Cr2O3/c-sapphire interface and (b) cross-sectional HAADF-STEM image of the Bi2Se3/Cr2O3. The horizontal arrows identify twin boundaries; (c) HRTEM images of the MgO/c-sapphire interface; (d) enlargement of the image in (c), showing the spinel structure at the MgO/c-sapphire interface and (e) cross-sectional HAADF-STEM image of the Bi2Se3/MgO interface in the c-sapphire [11‐20] zone axis.

FIG. 3.

(a) HRTEM image of Cr2O3/c-sapphire interface and (b) cross-sectional HAADF-STEM image of the Bi2Se3/Cr2O3. The horizontal arrows identify twin boundaries; (c) HRTEM images of the MgO/c-sapphire interface; (d) enlargement of the image in (c), showing the spinel structure at the MgO/c-sapphire interface and (e) cross-sectional HAADF-STEM image of the Bi2Se3/MgO interface in the c-sapphire [11‐20] zone axis.

Close modal

Fig. 3(c) shows the interface between MgO and c-sapphire in an MgO-bilayer. Fig. 3(d) shows more clearly the existence of a ∼2 nm spinel structure that formed at the interface when MgO was grown at 650 °C This interfacial layer, identified as MgAl2O4, was not observed when the deposition was carried out at 550 °C. It should be noted that the presence of this ∼2 nm interfacial layer at the MgO/c-sapphire interface does not affect the subsequent MgO growth as a rocksalt phase with the [111] axis aligned parallel to the substrate normal. The inset FFT diffraction patterns confirm the epitaxial growth of MgO. Since electrical current will only flow through the Bi2Se3 surface, i.e., charge transport is restricted solely to the upperlaying Bi2Se3 due to the highly insulating nature of MgO, the existence of ∼2 nm MgAl2O4 should not affect the transport properties. Fig. 3(e) shows an HRTEM image at the Bi2Se3 and MgO interface, showing a clean and sharp interface with no evidence of any interfacial reaction, which is crucial for a strong proximity effect in hybrid structures. Lamellar twins are found at the boundaries of the different stacking sequences parallel to the growth plane in Bi2Se3; these are seen in the marked region in Fig. 3(e). The microstructure results are consistent with the XRD Φ-scan with six peaks seen presented in Bi2Se3 bilayer.

Fig. 4 compares the resistivity versus temperature for all three sample types, directly revealing the influence of the magnetic layer on transport in Bi2Se3. At high temperatures, all films show metallic behavior. Although they have approximately the same temperature coefficients in resistivity, the bilayer samples have larger resistivity (or suppressed conductivity), compared to the control sample, with the highest resistivity found in the MgO-bilayer for all temperatures. The Hall effect measurements at 300 K indicate that charge carriers of Bi2Se3 in all samples are of n-type with a net carrier density ∼1.9 × 1019/cm3, 1.75 × 1019/cm3, and 1.84 × 1019/cm3 for the pristine, Cr2O3-bilayer, and MgO-bilayer, respectively. The carrier concentration does not vary significantly among the three samples. However, the mobility values of both bilayer samples are consistently smaller in magnitude than the pristine samples, having ∼426 cm2/V s for pristine and ∼334 cm2/V s and ∼108 cm2/V s cm3 for Cr2O3-bilayer and MgO-bilayer, respectively. One probable cause for this suppression in mobility is due to the presence of insulating magnetic layers. In addition, as the temperature decreases, the resistivity shows a minimum and then increases as the temperature decreases further, suggesting stronger electron-electron interactions in two-dimensional (2D) systems. A comparison of the low-temperature insulating behaviors of all samples is shown in the inset of Fig. 4. The insulating behavior in the bilayer samples is more pronounced than in the control sample. The onset temperature of the insulating behavior is also higher in the bilayer samples. These results suggest that the magnetic layers may be responsible for the stronger insulating behavior in the bilayer samples. The onset temperature in the MgO-bilayer sample (∼42 K) is higher than that in the Cr2O3-bilayer sample (∼28 K), which is explained on the basis of their microstructural differences. Fig. 5(a) shows the low magnification bright-field TEM image of a cross-section of the MgO-bilayer sample where Bi2Se3 contains a small misorientation between two grains, labelled A and B, respectively. The selected area electron diffraction (SAED) pattern in Fig. 5(b) further confirms that the Bi2Se3 spot is split into two components with coherence; the angle between these two spots is less than 2°. The observation of extra diffraction patterns with the same radius in Bi2Se3 is believed due to the small angle grain boundaries, referred to as sub-grain boundaries, when the Bi2Se3 was grown on MgO. It should be noted that the sub-grain boundaries are not distinctly sharp owing to the fact that these boundaries can run in any direction leading to an overlap of neighboring domains in the projected TEM zone direction. Nonetheless, such small-angle tilt grain boundaries are not observed in the Cr2O3-bilayer samples. The SAED pattern of Cr2O3-bilayer sample is shown in Fig. 5(c), and no evidence of the sub-grain diffraction patterns is observed. Therefore, the higher resistivity and onset temperature in the MgO-bilayer sample could be associated with grain boundary scattering, which could increase the resistivity.

FIG. 4.

Resistivity as a function of temperature for the Cr2O3-bilayer, the MgO-bilayer, and the control sample. The inset shows the normalized resistivity versus T.

FIG. 4.

Resistivity as a function of temperature for the Cr2O3-bilayer, the MgO-bilayer, and the control sample. The inset shows the normalized resistivity versus T.

Close modal
FIG. 5.

(a) Low magnification TEM image of an MgO-bilayer sample, showing two small misorientated Bi2Se3 grains (labelled A and B), (b) selected area diffraction pattern of an MgO-bilayer sample, showing the small misorientation (depicted by the small arcs) in Bi2Se3, and (c) selected area diffraction pattern of the Cr2O3-bilayer.

FIG. 5.

(a) Low magnification TEM image of an MgO-bilayer sample, showing two small misorientated Bi2Se3 grains (labelled A and B), (b) selected area diffraction pattern of an MgO-bilayer sample, showing the small misorientation (depicted by the small arcs) in Bi2Se3, and (c) selected area diffraction pattern of the Cr2O3-bilayer.

Close modal

It is important to note that in TI/FMI hybrid structures, even though strong exchange coupling exists across the interface, no gap in the topological surface states is expected with in-plane magnetization. When the magnetic easy axis of the FM is out-of-plane, a gap opens in the surface states of the vicinal TI.33,38 The magnetic properties were measured with the applied magnetic field parallel (Fig. 6) and perpendicular (Fig. 7) to the Bi2Se3 (003 n) surface for all bilayer sample types. The control samples are diamagnetic in both directions, consistent with previous reports.39 The in-plane magnetization at 300 K and 5 K for Cr2O3 and MgO-bilayer samples provide the evidence of ferromagnetic signatures with clear saturation and hysteresis. Interesting, the Cr2O3–bilayer sample is paramagnetic with observable coercivity in the out-of-plane direction, whereas the MgO-bilayer is diamagnetic in this orientation. More detailed analysis is required for investigating the paramagnetic mechanism along the c-axis of the Cr2O3–bilayer sample.

FIG. 6.

Magnetic hysteresis loops for in-plane magnetic field for (a) a control sample, (b) a Cr2O3-bilayer sample, and (c) an MgO-bilayer sample.

FIG. 6.

Magnetic hysteresis loops for in-plane magnetic field for (a) a control sample, (b) a Cr2O3-bilayer sample, and (c) an MgO-bilayer sample.

Close modal
FIG. 7.

Magnetic hysteresis loops for out-of-plane magnetic field for (a) a control sample, (b) a Cr2O3-bilayer sample, and (c) an MgO-bilayer sample.

FIG. 7.

Magnetic hysteresis loops for out-of-plane magnetic field for (a) a control sample, (b) a Cr2O3-bilayer sample, and (c) an MgO-bilayer sample.

Close modal

The normalized MRs as a function of temperature (2 K–25 K) and magnetic field perpendicular to the films are compared in Figs. 8(a) and 8(b) for Cr2O3-bilayer versus control sample and MgO-bilayer versus control sample, respectively. The shape of the MR in bilayer structures are drastically different from the control samples, though they have overall positive MR, and at low magnetic field, the MR exhibits the weak antilocalization (WAL) cusp feature. This WAL could reflect the nontrival topology of the surface states and is suppressed when the temperature increases as a result of the decreasing coherence length. At higher magnetic fields (>4 T), the MR of the control sample does not saturate and follows a linear-like dependence, showing a positive, weak temperature-dependence up to B = 5 T. The linearity of the MR can be interpreted through the quantum MR model proposed for a zero gap band structure with Dirac linear dispersion.34 In contrast, the bilayer samples quickly saturate at low magnetic field (∼2.5 T).

FIG. 8.

Normalized magnetoresistance, R(H)/R(0), versus temperature and perpendicular magnetic field for (a) Cr2O3-bilayer sample versus control sample and (b) MgO-bilayer sample versus control sample. (c) R(H)/R(0) comparison of three samples at T = 2 K for −2.5 T to +2.5 T (enlarged part is for −0.3 T to + 0.3 T).

FIG. 8.

Normalized magnetoresistance, R(H)/R(0), versus temperature and perpendicular magnetic field for (a) Cr2O3-bilayer sample versus control sample and (b) MgO-bilayer sample versus control sample. (c) R(H)/R(0) comparison of three samples at T = 2 K for −2.5 T to +2.5 T (enlarged part is for −0.3 T to + 0.3 T).

Close modal

Fig. 8(c) compares the MR at T = 2 K for all samples as a function of perpendicular magnetic field. As a result of proximity with Cr2O3 and MgO, the localization behavior of the Bi2Se3 is significant altered. The enlarged part in Fig. 8(c) clearly shows that the introduction of magnetic components reduces the sharpness of bilayer samples, the cusp features at low magnetic field, indicating a weakened WAL effect, which may be associated with an increase of backscattering of Dirac fermions.

To understand the origin of the weakened WAL effect in the bilayer samples, the low-field magnetoconductance data are analyzed quantitatively using the Hikami-Larkin-Nagaoka equation, and the phase coherence length (lΦ) as a function of temperatures is extracted for all samples:

ΔGWAL(B)=G(B)G(0)αe22π2[ Ψ(12+4eBl2Φ)ln(4eBl2Φ) ],
(1)

where lΦ is the phase coherence length, Ψ is the digamma function, and α is a coefficient determined by the type of localization. Theoretically, the coherence length is proportional to T−1/2 and T−3/4 for 2D and 3D systems, respectively.35 The monotonic decrease of the coherence length with increasing temperature is observed in both bilayer and the control samples, similar to other TI systems.36,37 The fitting gives lΦ = T−0.4289 for the control sample, lΦ = T−0.4271 for the Cr2O3-bilyer sample and lΦ = T−0.4144 for the MgO-bilayer sample, suggesting the dominant dephasing mechanism is e-e interactions in 2D surface states.

Yet, the Cr2O3-bilayer sample has the smallest lΦ among all samples whereas the MgO-bilayer and control samples have similar lΦ; lΦ(T) is plotted in Fig. 9. The noticeable difference in lΦ between Cr2O3-bilayer and MgO-bilayer samples is explained by the anisotropic magnetization in Cr2O3 and MgO layers. The Cr2O3–bilayer sample is paramagnetic with observable coercivity in the out-of-plane direction, whereas the MgO-bilayer is diamagnetic in this orientation. The considerable reduction of lΦ in the Cr2O3-bilayer is likely caused by additional inelastic scattering, such as electron-magnon scattering, to the topological surface states proximate to the magnetic layer with a perpendicular anisotropy component.32,40

FIG. 9.

Temperature dependence of the phase coherence length comparison among three sample types: a control sample, a Cr2O3-bilayer sample, and an MgO-bilayer sample.

FIG. 9.

Temperature dependence of the phase coherence length comparison among three sample types: a control sample, a Cr2O3-bilayer sample, and an MgO-bilayer sample.

Close modal

The angular dependence of the resistance was measured by rotating the samples 360° in a 7.5 T magnetic field at T = 2 K; the result are presented in Fig. 10. It is known that the WAL induced by 2D surface states is characterized by a sole dependence on the perpendicular component of the applied magnetic field (0° and 180°). A strong dependence of the resistance on angle is observed for the control sample; the data are accurately described by a cosine function. For the bilayer samples, the WAL contribution from the bottom surface states expected to be limited or may coexist with a 3D bulk Fermi surface owing to the magnetic layers. Therefore, both bilayer samples seem to suppress the surface transport channel, as well as the WAL, in the topological surface states. We attribute these transport phenomena to the increased magnetic scattering at the TI/FMI interface.

FIG. 10.

The angular dependence of resistance at 7.5 T, 2 K for (a) a control sample, (b) a Cr2O3-bilayer sample, and (c) an MgO-bilayer sample.

FIG. 10.

The angular dependence of resistance at 7.5 T, 2 K for (a) a control sample, (b) a Cr2O3-bilayer sample, and (c) an MgO-bilayer sample.

Close modal

We have demonstrated the growth of new platforms for topological insulator/ferromagnetic insulator heterostructures by interfacing Bi2Se3 with Cr2O3 and MgO thin films, showing proximity-induced interfacial magnetization. The lattice misfit of Cr2O3/Bi2Se3 and MgO/Bi2Se3 are ∼16% and ∼39%, respectively, where the critical thickness of pseudomorphic growth is less than one monolayer. The Bi2Se3 bilayers are characterized by atomically sharp and reaction-free interfaces. The insulating behavior in the bilayer samples are more pronounced and the onset temperatures are higher as compared to the control sample. The MgO-bilayer sample has the largest resistivity and the highest onset temperature (∼42 K) due to the existence of small misorientations between grains in the Bi2Se3 thin films. This could give rise to an increase in resistivity owing to grain boundary scattering. The shortest phase coherent length (lΦ) in the Cr2O3–bilayer samples is attributed to the additional inelastic scattering between the Cr2O3 and Bi2Se3 interface with a perpendicular anisotropy component. Our hybrid structures demonstrate the suppression of the 2D surface channel as well as a weakened WAL in the topological surface state. However, stronger perpendicular magnetocrystalline anisotropy is required to observe the time-reversal symmetry breaking and further investigation is needed to establish the exact mechanism for ferromagnetism in these heterostructures.

The authors acknowledge the use of the Analytical Instrumentation Facility (AIF) at North Carolina State University, which is supported by the State of North Carolina and the National Science Foundation. Y. F. Lee partly supported by ECCS-1306400 would like to Dr. Xiahan Sang and Dr. Yang Liu for technical support on the HAADF-STEM, and Sandhyarani Punugupati for the stimulating scientific discussion.

1.
X. L.
Qi
and
S. C.
Zhang
,
Rev. Mod. Phys.
83
(
4
),
1057
1110
(
2011
).
2.
M. Z.
Hasan
and
C. L.
Kane
,
Rev. Mod. Phys.
82
(
4
),
3045
3067
(
2010
).
3.
S. C.
Zang
and
X. L.
Qi
,
Phys. Today
63
(
1
),
33
(
2010
).
4.
A. M.
Essin
,
J. E.
Moore
, and
D.
Vanderbilt
,
Phys. Rev. Lett.
102
(
14
),
146805
(
2009
).
5.
H.
Peng
,
K.
Lai
,
D.
Kong
,
S.
Meister
,
Y.
Chen
,
X. L.
Qi
,
S. C.
Zhang
,
Z. X.
Shen
, and
Y.
Cui
,
Nat. Mater.
9
(
3
),
225
229
(
2010
).
6.
C.
Bai
and
X.
Zhang
,
Phys. Rev. B
76
(
7
),
075430
(
2007
).
7.
B. L.
Gelmont
,
M.
Shur
, and
M.
Stroscio
,
J. Appl. Phys.
77
(
2
),
657
660
(
1995
).
8.
Y. S.
Hor
,
P.
Roushan
,
H.
Beidenkopf
,
J.
Seo
,
D.
Qu
,
J. G.
Checkelsky
,
L. A.
Wray
,
D.
Hsieh
,
Y.
Xia
,
S. Y.
Xu
,
D.
Qian
,
M. Z.
Hasan
,
N. P.
Ong
,
A.
Yazdani
, and
R. J.
Cava
,
Phys. Rev. B
81
(
19
),
195203
(
2010
).
9.
V. A.
Kulbachinskii
,
A. Y.
Kaminskii
,
K.
Kindo
,
Y.
Narumi
,
K.
Suga
,
P.
Lostak
, and
P.
Svanda
,
Phys. B: Condens. Matter
311
(
3–4),
292
297
(
2002
).
10.
P. P. J.
Haazen
,
J. B.
Laloë
,
T. J.
Nummy
,
H. J. M.
Swagten
,
P.
Jarillo-Herrero
,
D.
Heiman
, and
J. S.
Moodera
,
Appl. Phys. Lett.
100
(
8
),
082404
(
2012
).
11.
X.
Kou
,
M.
Lang
,
Y.
Fan
,
Y.
Jiang
,
T.
Nie
,
J.
Zhang
,
W.
Jiang
,
Y.
Wang
,
Y.
Yao
,
L.
He
, and
K. L.
Wang
,
ACS Nano
7
(
10
),
9205
9212
(
2013
).
12.
Y. L.
Chen
,
J. H.
Chu
,
J. G.
Analytis
,
Z. K.
Liu
,
K.
Igarashi
,
H. H.
Kuo
,
X. L.
Qi
,
S. K.
Mo
,
R. G.
Moore
,
D. H.
Lu
,
M.
Hashimoto
,
T.
Sasagawa
,
S. C.
Zhang
,
I. R.
Fisher
,
Z.
Hussain
, and
Z. X.
Shen
,
Science
329
(
5992
),
659
662
(
2010
).
13.
L. A.
Wray
,
S. Y.
Xu
,
Y.
Xia
,
D.
Hsieh
,
A. V.
Fedorov
,
Y. S.
Hor
,
R. J.
Cava
,
A.
Bansil
,
H.
Lin
, and
M. Z.
Hasan
,
Nat. Phys.
7
(
1
),
32
37
(
2011
).
14.
S. Y.
Xu
,
M.
Neupane
,
C.
Liu
,
D.
Zhang
,
A.
Richardella
,
L.
Andrew Wray
,
N.
Alidoust
,
M.
Leandersson
,
T.
Balasubramanian
,
J.
Sánchez Barriga
,
O.
Rader
,
G.
Landolt
,
B.
Slomski
,
J.
Hugo Dil
,
J.
Osterwalder
,
T. R.
Chang
,
H.-T.
Jeng
,
H.
Lin
,
A.
Bansil
,
N.
Samarth
, and
M.
Zahid Hasan
,
Nat. Phys.
8
(
8
),
616
622
(
2012
).
15.
W.
Luo
and
X. L.
Qi
,
Phys. Rev. B
87
(
8
),
085431
(
2013
).
16.
S. V.
Eremeev
,
V. N.
Men'shov
,
V. V.
Tugushev
,
P. M.
Echenique
, and
E. V.
Chulkov
,
Phys. Rev. B
88
(
14
),
144430
(
2013
).
17.
V. N.
Men'shov
,
V. V.
Tugushev
,
S. V.
Eremeev
,
P. M.
Echenique
, and
E. V.
Chulkov
,
Phys. Rev. B
88
(
22
),
224401
(
2013
).
18.
Q. I.
Yang
,
M.
Dolev
,
L.
Zhang
,
J.
Zhao
,
A. D.
Fried
,
E.
Schemm
,
M.
Liu
,
A.
Palevski
,
A. F.
Marshall
,
S. H.
Risbud
, and
A.
Kapitulnik
,
Phys. Rev. B
88
(
8
),
081407
(
2013
).
19.
J. M.
Zhang
,
W.
Ming
,
Z.
Huang
,
G. B.
Liu
,
X.
Kou
,
Y.
Fan
,
K. L.
Wang
, and
Y.
Yao
,
Phys. Rev. B
88
(
23
),
235131
(
2013
).
20.
A.
Kandala
,
A.
Richardella
,
D. W.
Rench
,
D. M.
Zhang
,
T. C.
Flanagan
, and
N.
Samarth
,
Appl. Phys. Lett.
103
(
20
),
202409
(
2013
).
21.
P.
Wei
,
F.
Katmis
,
B. A.
Assaf
,
H.
Steinberg
,
P.
Jarillo Herrero
,
D.
Heiman
, and
J. S.
Moodera
,
Phys. Rev. Lett.
110
(
18
),
186807
(
2013
).
22.
J.
Narayan
and
B. C.
Larson
,
J. Appl. Phys.
93
(
1
),
278
285
(
2003
).
23.
W. H.
Butler
,
X. G.
Zhang
,
T. C.
Schulthess
, and
J. M.
MacLaren
,
Phys. Rev. B
63
(
5
),
054416
(
2001
).
24.
J.
Mathon
and
A.
Umerski
,
Phys. Rev. B
63
(
22
),
220403
(
2001
).
25.
R.
Cheng
,
C. N.
Borca
,
N.
Pilet
,
B.
Xu
,
L.
Yuan
,
B.
Doudin
,
S. H.
Liou
, and
P. A.
Dowben
,
Appl. Phys. Lett.
81
(
11
),
2109
2111
(
2002
).
26.
J. S.
Moodera
,
L. R.
Kinder
,
T. M.
Wong
, and
R.
Meservey
,
Phys. Rev. Lett.
74
(
16
),
3273
3276
(
1995
).
27.
J.
Li
,
Y.
Jiang
,
Y.
Li
,
D.
Yang
,
Y.
Xu
, and
M.
Yan
,
Appl. Phys. Lett.
102
(
7
),
072406
(
2013
).
28.
X.
He
,
Y.
Wang
,
N.
Wu
,
A. N.
Caruso
,
E.
Vescovo
,
K. D.
Belashchenko
,
P. A.
Dowben
, and
C.
Binek
,
Nat. Mater.
9
(
7
),
579
585
(
2010
).
29.
S.
Punugupati
,
J.
Narayan
, and
F.
Hunte
,
Appl. Phys. Lett.
105
(
13
),
132401
(
2014
).
30.
Y. F.
Lee
,
R.
Kumar
,
F.
Hunte
,
J.
Narayan
, and
J.
Schwartz
,
Acta Mater.
95
,
57
64
(
2015
).
31.
K.
He
,
Y.
Zhang
,
C. Z.
Chang
,
C. L.
Song
,
L. L.
Wang
,
X.
Chen
,
J. F.
Jia
,
Z.
Fang
,
X.
Dai
,
W. Y.
Shan
,
S. Q.
Shen
,
Q.
Niu
,
X. L.
Qi
,
S. C.
Zhang
,
X. C.
Ma
, and
Q. K.
Xue
,
Nat. Phys.
6
(
8
),
584
588
(
2010
).
32.
H. Z.
Lu
,
J.
Shi
, and
S. Q.
Shen
,
Phys. Rev. Lett.
107
(
7
),
076801
(
2011
).
33.
H. Z.
Lu
and
S. Q.
Shen
,
Phys. Rev. B
84
(
12
),
125138
(
2011
).
34.
H.
He
,
B.
Li
,
H.
Liu
,
X.
Guo
,
Z.
Wang
,
M.
Xie
, and
J.
Wang
,
Appl. Phys. Lett.
100
(
3
),
032105
(
2012
).
35.
B. L.
Altshuler
,
A. G.
Aronov
, and
D. E.
Khmelnitsky
,
J. Phys. C: Solid State Phys.
15
(
36
),
7367
(
1982
).
36.
S.
Matsuo
,
T.
Koyama
,
K.
Shimamura
,
T.
Arakawa
,
Y.
Nishihara
,
D.
Chiba
,
K.
Kobayashi
,
T.
Ono
,
C.-Z.
Chang
,
K.
He
,
X.-C.
Ma
, and
Q.-K.
Xue
,
Phys. Rev. B
85
(
7
),
075440
(
2012
).
37.
L.
Bao
,
L.
He
,
N.
Meyer
,
X.
Kou
,
P.
Zhang
,
Z.-G.
Chen
,
A. V.
Fedorov
,
J.
Zou
,
T. M.
Riedemann
,
T. A.
Lograsso
,
K. L.
Wang
,
G.
Tuttle
, and
F.
Xiu
,
Sci. Rep.
2
,
726
(
2012
).
38.
X. L.
Qi
,
T. L.
Hughes
, and
S. C.
Zhang
,
Phys. Rev. B
78
(
19
),
195424
(
2008
).
39.
H.
Li
,
Y. R.
Song
,
M.-Y.
Yao
,
F.
Zhu
,
C.
Liu
,
C. L.
Gao
,
J.-F.
Jia
,
D.
Qian
,
X.
Yao
,
Y. J.
Shi
, and
D.
Wu
,
J. Appl. Phys.
113
(
4
),
043926
(
2013
).
40.
H. T.
He
,
G.
Wang
,
T.
Zhang
,
I.-K.
Sou
,
G. K. L.
Wong
,
J. N.
Wang
,
H. Z.
Lu
,
S. Q.
Shen
, and
F. C.
Zhang
,
Phys. Rev. Lett.
106
(
16
),
166805
(
2011
).