Open Access Article
Thomas J. Rehaaga,
Ethan D. N. Dommettab,
Zhuo Yua,
Hong Yi Cheng Lina,
David Ironmongera,
Yisong Hana and
Gavin R. Bell
*a
aDepartment of Physics, University of Warwick, Coventry CV4 7AL, UK. E-mail: gavin.bell@warwick.ac.uk
bCY Cergy Paris University, Boulevard du Port 95011, Cergy-Pontoise, France
First published on 17th October 2025
Epitaxial growth of the topological semimetal SrMnSb2 is demonstrated using InAs and GaSb substrates. The growth is studied by reflection high energy electron diffraction, X-ray photoemission spectroscopy and electron microscopy. The epitaxial orientation is SrMnSb2 (001)||III–V (001), with the epilayer <010> axes parallel to substrate <110>. The relaxed in-plane lattice parameter of the epilayer depends on the vicinal miscut of the substrate and results in an overall volume expansion of the unit cell compared to bulk crystals, along with a possible lifting of the bulk orthorhombic distortion. Segregation of substrate atoms through the growing film is observed, along with Mn-driven endotaxial growth of secondary phases. Interface control methods to suppress endotaxial growth are discussed.
Epitaxial thin film material is preferable to bulk single crystals for device applications: not only does it allow efficient device processing and improved compatibility with existing electronic materials, but additional control is possible via quantum size effects, strain engineering and magnetic proximity effects. Crucially, the effects of the coherent phonon could be studied by ultra-fast pump-probe spectroscopy of thin-film SrMnSb2.11 Although the related materials SrMnBi2 (ref. 12 and 13) and MnSb (ref. 14) have been grown using molecular beam epitaxy (MBE), thin-film SrMnSb2 has not yet been grown and characterised. The MBE growth of transition metal pnictides such as MnSb,15–18 CrSb (ref. 19) and NiSb (ref. 20) is quite well established.21 In many respects the surface chemistry of the epitaxy appears to follow that of the III–V semiconductors, at least in respect of the pnictogen. At sufficiently high substrate temperature, incorporation of the group V element is self-limiting, leading to stoichiometry between the transition metal and pnictogen components. At too low a substrate temperature, excess pnictogen may be incorporated. Excessively high substrate temperature leads to the group V element re-evaporating before incorporation.18 However, the growth window between these two extremes is wide (at least 200 °C for MnSb) and so there is scope for MBE optimisation on different substrates. Endotaxial growth of secondary phases has been noted for MnSb grown on both GaSb (ref. 22) and InP, (ref. 23) and for MnP on GaP,24 i.e. these phases form as nano-crystals below the initial substrate surface. This is potentially problematic for SrMnSb2, whose antiferromagnetism may be masked by ferromagnetic secondary phases such as MnSb or MnAs.
In this paper we describe the MBE growth of SrMnSb2(001) on InAs and GaSb(001) substrates. The primary interfacial effects of segregation and endotaxy are explored using structural and chemical analysis.
Fig. 1 shows a possible epitaxial arrangement on InAs(001). The (a, c) axes of the SrMnSb2 align with the [110] and [110] primitive surface mesh axes of the InAs with average lattice mismatch 3.2%. This is indeed the observed epitaxial relationship, although the precise interfacial atomic structure is not known. The figure illustrates the lower Sb(1) layer of the SrMnSb2 adopting the approximate atomic positions of the surface As atoms in bulk-like As-terminated InAs(001). The vertical black bar represents the a/2 ≈ 3 Å step height on InAs(001), which is far smaller than the c lattice parameter of SrMnSb2. Short grey lines joining In to Sb atoms represent interfacial bonds with an exaggerated in-plane epitaxial mismatch.
Epitaxial mismatches (% change between film and substrate lattice parameter, normalised to the latter) are shown in Table 1 for several common semiconductor materials. Negative values indicate that the SrMnSb2 must be compressively strained to match the substrate, and the positive values (for InSb only) would require tensile strain in the SrMnSb2. The (a, b) column uses the averaged in-plane lattice paramter of SrMnSb2 and this quanitity will be discussed subsequently. A good lattice match to the average (a, b) spacing of bulk SrMnSb2 could be achieved by growing on the ternary materials InAs0.5Sb0.5 or Ga0.6In0.4Sb.
| Substrate | % (a) | % (b) | % (a, b) | dc (nm) |
|---|---|---|---|---|
| InAs | −2.6 | −3.7 | −3.2 | 4.8–11.4 |
| GaSb | −2.0 | −3.1 | −2.5 | 9.9–22.5 |
| InSb | 4.1 | 3.0 | 3.5 | 3.6–8.7 |
| GaAs | −9.9 | −11.1 | −10.5 | |
| Si | −14.5 | −15.7 | −15.1 | |
| SrTiO3 | −12.9 | −14.2 | −13.6 |
The critical thickness dc for plastic relaxation via misfit dislocations is also tabulated. This was estimated using the People-Bean model29 evaluated in closed form using the Lambert W-function.30 Elastic constants are taken from first-principles calculations reported in Materials Project.31 Because the Burgers vectors of misfit dislocations in SrMnSb2 are not known, a range from a√ 2 to a is used. The three high-mismatch substrates are out of the range of validity of the continuum elastic model. The calculation suggests that pseudomorphic SrMnSb2 should be observable on GaSb(001), but that relaxation may occur on InAs(001) before even 2 ML film thickness (here defining 1 monolayer, ML, of SrMnSb2 as one bulk unit cell, 2.3 nm height).
:
isopropanol mix) immediately before loading to vacuum under dry nitrogen, argon ion sputtering (500 eV ion energy, 1 μA beam current) and annealing. The types of substrate and buffer layer are summarised in Table 2. For all substrates, SrMnSb2 growth was performed either directly on the cleaned substrate or on a GaSb buffer layer.
| Material | Miscut (º) | Buffer (nm) | Section |
|---|---|---|---|
| InAs | 0 ± 0.1, 0.5, 1, 2 | None | 4.1, 4.2 |
| InAs | 0 ± 0.1, 0.5, 1, 2 | GaSb 10–20 nm | 4.1, 4.4 |
| GaSb | 0 ± 0.1 | None | 4.1, 4.3 |
| GaSb | 0 ± 0.1 | GaSb 10–20 nm | SI |
The MBE growth was performed using shuttered effusion cells for Ga, Sr, Mn and Sb4. These were calibrated using beam equivalent pressure (BEP) from a retractable ion gauge, and via scanning transmission electron microscopy (STEM) measurements of total film thicknesses. Growth was monitored using reflection high energy electron diffraction (RHEED) with a 12.5 keV beam. The fluxes and substrate temperature were initially based on our experience with MBE growth of MnSb,17,32 treating the summed Sr and Mn fluxes as equivalent to the Mn flux for MnSb. Optimising the quality of the RHEED patterns from this starting point led to the following standard growth conditions: substrate temperature Tsub = 340 °C, BEP ratios Sr
:
Mn
:
Sb approximately 1.0
:
1.2
:
6.5, total growth rate 0.6 nm min−1, film thickness between 12 and 120 nm (5 to 50 ML). Any variations on these conditions are discussed in the individual results sections. The GaSb(001) buffer layers were grown at Tsub = 360 °C and 0.7 nm min−1.
The STEM experiments were performed in a double aberration-corrected JEOL ARM200f TEM, operated at 200 kV. This TEM is equipped with an Oxford Instruments SDD energy dispersive X-ray analysis (EDX) detector, and elemental maps were obtained for the main elements in the heterostructures. Additional standard TEM images were obtained using a JEOL 2100 microscope, also operating at 200 kv. Samples were transferred either through air or via vacuum suitcase to separate chambers for study by X-ray photoelectron spectroscopy (XPS). X-ray diffraction (XRD) was also performed on selected films (Panalytical X'Pert Pro MRD, Cu Kα radiation). Samples transferred by air were protected from oxidation by arsenic caps deposited in a side chamber on the MBE vacuum system. Some samples were also capped with Sb by cooling in the MBE chamber under Sb4 flux. Caps were not completely removed in XPS since the maximum annealing temperature used was set at the MBE substrate temperature.
Fig. 3 shows XRD data for a thick SrMnSb2 epilayer grown directly on InAs(001). Apart from the substrate peaks, a family of sharp peaks is visible which can be assigned to SrMnSb2(0004n) reflections. Together with the RHEED data, this confirms the epitaxial orientation of SrMnSb2. The peak series can be fitted with a lattice parameter of 25.074(4) Å, slightly larger than the c parameter of bulk crystals.10 Scherrer analysis (SI) gives a crystallite size of 10.7 nm, significantly smaller than the film thickness. This indicates that the epilayer contains domains or other extended defects, consistent with the poorer RHEED patterns for the thickest films. The XRD peaks labelled with asterisks correspond to secondary phases. The blue asterisk peaks (5.37 Å) can be tentatively identified with strained orthorhombic MnAs.21,32 The remaining peaks, black asterisks, can be tentatively assigned to cubic MnSb (ref. 32) and InSb. Note that if the unit cell of the SrMnSb2 epilayer is identified as tetragonal rather than orthorhombic, then the main peaks can be assigned to (0002n) reflections with c = 12.537(4) Å.
![]() | ||
| Fig. 3 XRD data for 80 nm SrMnSb2 grown directly on InAs(001). Substrate and SrMnSb2 peaks are labelled. The asterisks denote secondary phases. | ||
Fig. 4 shows example STEM and EDX data. Panel (a) gives an annular dark field (ADF) STEM image of a well ordered region of a SrMnSb2 film, which was grown on an InAs(001) substrate miscut by 1° towards (111)B. The orthorhombic bulk unit cell is outlined by an orange box (4.6 Å × 24.1 Å). Panels (b, c, d, e) show a STEM image and the corresponding EDX elemental maps, obtained from a SrMnSb2 sample grown directly on InAs(001). The SrMnSb2 epilayer is less uniform and its interface to InAs, appearing as a dark horizontal band, is rougher. Furthermore, a triangular or dome-like feature is visible below the substrate surface. As shown by the elemental maps, Sr and Mn appear in the epilayer, as expected. However, the Mn signal follows the shape of the dome-like sub-surface feature. The As signal of the InAs substrate is quite uniform, allowing us to identify the subsurface nano-cluster as endotaxial MnAs. A small amount of Sr also appears in the endotaxial nano-cluster. Finally, a significant As signal [panel (c)] is observed in the SrMnSb2 epilayer above the endotaxial cluster, indicating local segregation of As from the substrate into the growing film.
Fig. 5 gives XPS data for films grown directly on InAs and GaSb substrates. Panel (a) shows an XPS survey scan after vacuum suitcase transfer of a 35 nm thick film. No O 1s or C 1s peaks can be observed, indicating that the suitcase transfer did not introduce any surface contamination. The films have complete coverage and are sufficiently thick to prevent photoelectrons escaping from the substrate, and so the presence of clear In 3d and 3p peaks implies that In has segregated through the SrMnSb2 film.
The XPS was quantified using standard atomic sensitivity factors, with results for [Sr]/[Mn] ratios and [Sb]/([Sr] + [Mn]) ratios shown in Fig. 5(b) and (c) respectively. The [Sr]/[Mn] ratio inferred from XPS follows stoichiometric incorporation of the fluxes as measured by BEP, with a slight excess of Sr possible at low BEP ratio. The Sb incorporation shows a similar trend with BEP ratio. Although the Sb incorporation does not seem to show clear self-limiting behaviour, as is the case for MnSb,21 only two data points lie above the line [Sb]/([Sr] + [Mn]) = 1. The BEP standard ratios used aim to provide stoichiometric SrMnSb2 and the average Sr
:
Mn
:
Sb atomic ratios inferred from XPS for samples grown close to the standard conditions were 1
:
0.89
:
1.95.
To summarise these general results, the RHEED and XRD confirm that the epitaxial orientation of the SrMnSb2 has the long [001] axis out of plane and the [100] and [010] axes aligned with substrate [110] and [110] bulk axes (i.e. the primitive surface mesh [10] and [01] axes). Diffraction results are compatible with relaxation of the orthorhombic distortion of bulk material towards a smaller tetragonal unit cell. The standard growth conditions produce SrMnSb2 films with close to stoichiometric composition. For growth on InAs substrates, indium surface-segregates through the films. Furthermore, endotaxial growth of MnAs nano-clusters within the substrate occurs, accompanied by some local diffusion of As into the epilayer and Sb-related secondary phases.
The in-plane lattice parameter measured by RHEED rapidly increases from the substrate value as growth proceeds, as shown in Fig. 7 for two different substrate miscuts. At the earliest stages of growth, up to 1 ML SrMnSb2 coverage, the lattice parameter remains at the substrate value (SI). Full relaxation occurs by around 5 to 10 ML coverage (10 to 20 nm). These values are broadly in line with the estimated critical thicknesses (Table 1). However, the lattice parameter clearly increases past the nominal bulk values. The “over-relaxation” is unusual, especially when considered alongside the out-of-plane lattice expansion relative to bulk material inferred from XRD (Fig. 3). Combining the two expansions, the unit cell volume of the epitaxial material (4.6 × 4.6 × 25.074 Å3) is increased by 16% compared to the bulk crystal (4.41 × 4.46 × 23.1 Å3).10 This is not compatible with a biaxial distortion from the substrate mismatch. The different strain relaxation for growth on two different InAs(001) miscuts is suggestive of strain relaxation via vicinal surface epitaxy, as previously observed for III-nitride growth on SiC,33,34 Ga2O3 on sapphire35 and chalco-genide films on Si.36
When the RHEED patterns become spotty, it is possible to measure the out-of-plane lattice parameter of the film (SI). These measurements give values close to c/2. This agrees with the X-ray diffractogram (Fig. 3), which shows the (0 0 4n) family of peaks (n = 1, 2, …) rather than the (0 0 2n). This can be interpreted as a lifting of the distortions which distinguish the two halves of the full c = 23.10 Å unit cell and split the (a, b) lattice parameters, and is consistent with observation by RHEED of equal (a, b) values. This smaller tetragonal unit cell is very close in energy (1 meV per atom (SI)) to the orthorhombic ground state31 and could readily be stabilised by epitaxial strain or the presence of segregated substrate atoms in the structure.
The crystallite size was inferred from XRD as being smaller than film thickness, and this is confirmed by TEM. Fig. 8 shows example images from SrMnSb2 grown directly on InAs(001). Image A shows a region of crystallographic disruption at the interface and grain-like structures above in the film. There is also clear endotaxial disruption below the interface. Atomic scale imaging is shown in B, with unit cells highlighted. The TEM data supports the epitaxial relationship and lattice parameters inferred from RHEED and XRD. However, a clear grain boundary is visible, most likely associated with a stacking fault. There is also a significant loss of atomic scale contrast close to the interface. The effective gain size is inevitably less than the film thickness for such samples. Smaller and more crystallographically diverse grains are observed in films with poorer RHEED patterns and a higher degree of endotaxial disruption (SI).
Three methods to suppress the endotaxial growth were investigated: (1) low temperature seed layer, (2) Sb soak, and (3) Sr–Sb growth initiation (delayed Mn flux). For (1), the substrate temperature was reduced to the range 30 °C to 200 °C for initial SrMnSb2 growth up to a few ML thickness. However, this resulted in rapid and unrecoverable degradation of the RHEED pattern and was abandoned. For the Sb soak (2), the InAs(001) surface was exposed to Sb flux for several minutes at the standard substrate temperature prior to initiation of SrMnSb2 growth. The intention was to allow any excess In atoms at the surface to react with Sb and hence suppress surface segregation of In. This procedure caused the fractional order RHEED streak intensity of the initial (4 × 2) to be reduced but did not degrade the RHEED pattern of the subsequent SrMnSb2 epilayers. However, segregated In was still observed in XPS, suggesting that the out-diffusion of In is not just a consequence of In nano-clusters or adatoms left from the surface cleaning process.37 For procedure (3), the Sr shutter was opened 30 s before the Mn shutter (since the predominant endotaxial species was MnAs, it was hoped to discourage direct Mn interaction with the substrate). The SrMnSb2 RHEED patterns were marginally better, and TEM data suggest that there is an overall reduction in secondary phase formation (SI). In addition, the delayed Mn flux reduced the concentration of surface-segregated indium by around 20%, according to XPS, and so this simple procedure was retained.
Clearly, the endotaxial growth of MnAs and related phases, along with concomitant segregation of In and As into the epilayer, has a deleterious effect on the uniformity of the SrMnSb2 and the quality of the interface. Furthermore, MnAs is ferromagnetic21 and could easily mask the desired antiferromagnetic response of the SrMnSb2. We therefore consider GaSb substrates.
However, some segregation and endotaxial growth were observed. Fig. 9 shows a STEM image and selected (SI) EDX elemental maps for a SrMnSb2/GaSb(001) interface. On this wider view, the interface is still quite abrupt but some evidence for endotaxial growth and segregation. Two dome-like protrusions of Mn some tens of nm into the substrate can be seen in the centre of the Mn EDX map. The corresponding Sb map is quite uniform, suggesting that a MnSb phase has formed. There is a noticeable Ga deficiency close to the interface above the right hand side MnSb endotaxial nano-cluster, and above that a sharp “streak” of Ga can be seen within the SrMnSb2 film. This shows that local Ga segregation has occurred. Sb is deficient on this streak, and there is a slight excess of Mn. The surface of the SrMnSb2 film is defined by the Mn and Sb EDX maps. It is not planar but rather comprises mesa-like structures with flat tops. This morphology is consistent with the spotty RHEED patterns observed at higher film thicknesses.
![]() | ||
| Fig. 9 STEM image and three EDX maps (Ga, Sb, Mn) for SrMnSb2 grown directly on GaSb(001). The interface is highlighted by white dashes on the right in each panel. | ||
The over-relaxation of the SrMnSb2 epilayer also occurs on this substrate. Estimating unit cell dimensions from STEM (Fig. 4) gives a volume of (4.6 × 4.6 × 24.1 Å3), a volume expansion of 12% compared to bulk material.
Two TEM images of a sample grown on 15 nm GaSb–InAs(001) buffer layer are shown in Fig. 10. In panel (a), the InAs-to-GaSb interface appears as a lighter band, with the GaSb buffer layer darker than the substrate (colour highlights on the right guide the eye). The interface between the GaSb buffer and the SrMnSb2 is not atomically abrupt. However, there is no evidence for significant endotaxial phase formation. Available measurements over several samples suggest that endotaxial growth was significantly less prevalent for samples growth on buffer layers compared to direct growth on GaSb. Note that this sample was capped with a thin GaSb layer, which shows a granular structure (top layer highlighted in red). Panel (b) shows a higher magnification image of the interfaces. Yellow dashes highlight the abrupt GaSb-on-InAs interface, with the GaSb showing a continuous pseudomorphic crystal structure. The red dashes show the SrMnSb2-on-GaSb interface, which is less abrupt and fluctuates by several atomic layers. The SrMnSb2 film shows three domains. The TEM images and RHEED (Fig. 2) confirm the improved SrMnSb2 film quality when grown on GaSb buffer layers, even when they are thin (as low as 10 nm).
Growth is hampered by the formation of endotaxial secondary phases alongside segregation of atoms from the substrate. This is particularly severe for direct growth on InAs substrates, where MnAs nanoclusters and other phases are formed. Both In and As segregate into the SrMnSb2 films, the former surface-segregating through the films to give a prominent In signal in XPS. Arsenic segregation appears to associated more with local defect phases, as does Ga segregation for growth on GaSb. Controlled interface formation by Sb soak and delayed Mn exposure helped to limit endotaxial phase formation. Overall, in order of prevalence of endotaxial growth, direct growth on InAs substrates was worst, followed by direct growth on GaSb, with growth on GaSb buffer layers the most favourable.
A promising method for reliably suppressing endotaxial growth is remote epitaxy via the inclusion of an ultra-thin amorphous carbon layer.41 This acts to decouple the growing film from the substrate while maintaining the epitaxial relationship. A key advantage of amorphous carbon is that it can be grown in situ on any III–V substrate and does not require wet or dry transfer, as would be the case for a 2D material such as graphene. To our knowledge, this application of remote epitaxy has not yet been studied. A longer Sb soak for interface optimisation of GaSb buffer layers39 may also improve the surface quality for SrMnSb2 epitaxy. Atomic exchange models with multiple species,38 extended to include Mn, may be useful in understanding endotaxial growth.
Supplementary information: this includes additional analysis of RHEED, XRD, TEM and XPS. See DOI: https://doi.org/10.1039/d5lf00284b.
| This journal is © The Royal Society of Chemistry 2025 |