Ferrimagnetic alloy thin films that exhibit perpendicular (out-of-plane) magnetic anisotropy (PMA) with low saturation magnetization, such as GdCo and Mn4N, were predicted to be favorable for hosting small Néel skyrmions for room temperature applications. Due to the exponential decay of interfacial Dzyaloshinskii–Moriya interaction and the limited range of spin–orbit torques, which can be used to drive skyrmion motion, the thickness of the ferrimagnetic layer has to be small, preferably under 20 nm. While there are examples of sub-20 nm, rare earth-transition metal (RE-TM), ferrimagnetic thin films fabricated by sputter deposition, to date, rare-earth-free sub-20 nm Mn4N films with PMA have only been reported to be achieved by molecular beam epitaxy, which is not suitable for massive production. Here, we report the epitaxial growth of sub-20 nm Mn4N films with PMA at 400 °C–450 °C substrate temperatures on MgO substrates by reactive sputtering. The Mn4N films were achieved by reducing the surface roughness of MgO substrate through a high-temperature vacuum annealing process. The optimal films showed low saturation magnetization (Ms = 43 emu/cc), low magnetic anisotropy energy (0.7 Merg/cc), and a remanent magnetization to saturation magnetization ratio (Mr/Ms) near 1 at room temperature. Preliminary ab initio density functional theory calculations have confirmed the ferrimagnetic ground state of Mn4N grown on MgO. The magnetic properties, along with the high thermal stability of Mn4N thin films in comparison with RE-TM thin films, provide the platform for future studies of practical skyrmion-based spintronic materials.

As industry rapidly transitions to using big data in their operations and decision making, there is an urgent need to develop technologies that accommodate the increasing requirements of high-density data storage.1 One promising candidate that has received increasing attention is using magnetic skyrmions as information carriers and manipulating them with current by spin–orbit torques (SOTs) for logic or memory operations2,3 (e.g., racetrack memories). Magnetic skyrmions are swirling spin configurations of neighbor atoms in a magnetic material with topologic protection. The size of the skyrmion could be as small as a few nanometers,4 and the manipulation of skyrmions could be energy efficient5,6 in high-density data storage applications.7,8 To date, most RT skyrmions have been discovered in ferromagnetic (FM) multilayer stacks.7–10 The use of a ferromagnets as skyrmions has certain shortcomings. For instance, ferromagnets suffer from large stray fields and large saturation magnetization (Ms), which makes the region of parameter space for small skyrmion less accessible, especially at room temperature (RT).11 

Unlike ferromagnets, the ferrimagnetic counterparts have small stray fields and low Ms, which makes them suitable for hosting small RT skyrmions.11 This makes ferrimagnet a promising candidate material for high-density data storage applications. One prototypical example is the amorphous GdCo ferrimagnetic thin film. Small (10 nm–30 nm) room temperature skyrmions have been reported in Pt/GdCo (6 nm)/TaOx.12 While the amorphous GdCo ferrimagnet shows promising characteristics relative to the ferromagnets, it suffers from poor thermal stability. It has been shown that in amorphous GdCo, PMA is lost after annealing at 300 °C–400 °C,13 the temperature range applied in complementary metal–oxide–semiconductor (CMOS) fabrication.14 Since Néel skyrmions can only exist in a heterostructure with PMA, the poor thermal stability of RE-TM films will have a deleterious effect in both the fabrication and performance of amorphous RE-TM in devices that leverage skyrmions for storage technology. One of the potential solutions to this problem is to explore crystalline ferrimagnets with PMA synthesized by deposition at high-temperature near 400 °C, thus ensuring compatibility with conventional CMOS processing.

In addition to the intrinsic effects that we have discussed thus far, the film thickness is an important extrinsic effect that impacts the overall device performance. The importance of film thickness applies to both ferromagnets and ferrimagnets. It is now well-established that the interfacial Dzyaloshinskii–Moriya interaction (DMI), which stabilizes the magnetic skyrmions in multilayers and heterostructures, decays exponentially with increasing film thickness.15–17 Further, the SOT scales inversely with the thickness.18 This necessitates the growth of thin film magnets with sub-20 nm thickness for the realization of small skyrmions in practical applications.12,19,20

Anti-perovskite Mn4N has been known as a crystalline, rare-earth-free ferrimagnetic material. In Mn4N, the Mn-atoms have a face-centered cubic structure with one N-atom at the body center. A schematic of the unit cell is shown in Fig. 1. The Mn-atoms at the corner and the face center have inequivalent magnetic moments and are ferrimagnetically coupled.21 Although the easy axis of bulk Mn4N is along the [111] direction, PMA is repeatedly and reproducibly observed in crystalline Mn4N films.22–28 As a result, ferrimagnetic Mn4N thin films have also attracted increasing interest in spintronics applications. Compared to the amorphous RE-TM ferrimagnetic alloys, the Mn4N system has better thermal stability for two main reasons. First, in most thin film studies, it has been shown that the anti-perovskite crystal structure is formed at 400 °C–450 °C. No loss of PMA is reported in thin films after annealing, which is an encouraging outcome. Second, there is no known evidence for any structural phase transformation on cooling to room temperature. Thus, we interpret that the anti-perovskite crystal structure is tolerant of high-temperature device fabrication processes (unlike the GdCo amorphous alloys). As noted earlier, the emergence of PMA is an important magnetic property for its use in spintronic applications. One of the plausible reasons for the PMA in Mn4N thin films could be attributed to the deviation of the out-of-plane lattice constant (c) to the in-plane lattice constant (a) (c/a) ratio from 122–28 due to the in-plane epitaxial strain. A recent study also showed that the anisotropy energy is correlated with the c/a ratio.27 We note that further studies are warranted to understand the interplay of magnetic and strain effects on the PMA and this is beyond the scope of this paper.

FIG. 1.

Mn4N crystal structure.

FIG. 1.

Mn4N crystal structure.

Close modal

Several groups have grown crystalline Mn4N epitaxial films (30 nm–100 nm) on MgO, SrTiO3, or LaAlO3 substrates by magnetron sputtering,22,23 molecular beam epitaxy (MBE),24–27 and pulsed laser deposition (PLD),28 and they reported similar c/a ratios of ∼0.99. The reported uniaxial magnetic anisotropy constant (Ku) and Ms of those Mn4N films were about 0.5 Merg/cc–1 Merg/cc and 50 emu/cc–100 emu/cc, respectively,22–28 comparable with the data observed for amorphous GdCo thin films (Ku ∼ 0.25 Merg/cc and Ms ∼ 50 emu/cc).12 

To date, however, only a few groups have grown sub-20 nm Mn4N thin films with good magnetic hysteresis [M(H)] loops, which has the remanent magnetization Mr to Ms ratio larger than 0.5 (Mr/Ms > 0.5), by MBE.26,27 For those reported Mn4N films by sputtering, the thicknesses were 30 nm22 to 100 nm,23 some of them even up to hundreds nanometers.29,30 Similar results on sputter deposited sub-20 nm Mn4N thin films have not been reported. Compared to MBE, sputter deposition is a more widely adopted method in CMOS technology. In this work, we have epitaxially grown sub-20 nm Mn4N thin films on the MgO (001) substrate with PMA by reactive sputtering. The effect of MgO substrate morphology on the quality of Mn4N films is studied. We then compare the magnetic properties of the Mn4N films on the MgO substrate from experiment and first principles-based density functional theory (DFT) calculations. The main contribution of this paper lies in the demonstration of the growth of high magnetic quality sub-20 nm crystalline Mn4N films on the MgO substrate using reactive sputtering that is more promising for scale-up production.

Mn4N thin films with nominal thicknesses of ∼15 nm and 10 nm were deposited on MgO(001) 5 × 5 × 0.5 mm3 substrates by reactive rf-sputtering at 400 °C and 450 °C substrate temperatures and a base pressure of 7 × 10−8 Torr. The flow rates of Ar and N2 gases were controlled by a mass flow meter, and we maintained a flow rate Ar:N2 ratio of 93:7. A 3 nm Pt capping layer was deposited on the Mn4N layer at room temperature to prevent oxidization. Before loading the MgO (001) substrate into the vacuum chamber, it was wet-cleaned with 2% diluted Hellmanex III alkaline detergent, acetone, and isopropanol and sealed in a vacuum tube with pressure 30 mTorr. The MgO substrate was baked at high temperature 1000 °C or 1100 °C for 4 h. Then, the substrates were annealed inside the chamber at 500 °C for 1 h to remove the surface contaminations. The Mn target was pre-sputtered with Ar gas for 20 min to remove the surface oxide. The surface roughness of the MgO substrate was measured using atomic force microscopy (AFM). The deposition rate was measured by x-ray reflectometry (XRR). The film compositions were determined with x-ray photoelectron spectroscopy (XPS). The film thickness was verified by XRR with an XRR simulation/calculation program (Rigaku, GXRR).31 

The epitaxial growth of the Mn4N crystal layer was demonstrated using x-ray diffraction (XRD) with Cu-Kα radiation. The magnetic properties of the samples were measured at room temperature with vibrating sample magnetometry (VSM). The diamagnetic component of the substrate was deduced from the slope of raw M-H curves at a large H region and subtracted from the raw data. Ku was calculated from effective anisotropy K with the following equations:

Ku=K+2πMs2,
(1)
K=0Measy(H)Mhard(H)dH,
(2)

where Measy and Mhard are the magnetization in the out-of-plane applied field and in-plane applied field, respectively.

Ab initio electronic structure calculations were carried out in the Density Functional Theory (DFT) framework using the plane-wave pseudopotential Quantum ESPRESSO code.32,33 Core and valence electrons were treated using the ultrasoft pseudopotential method.34,35 The exchange-correlation functionals were described using the Perdew–Burke–Ernzerhof parameterization of the generalized gradient approximation modified for solids (PBEsol).36 The plane-wave cutoff energy was set to 60 Ry, and a Γ-centered Monkhorst–Pack k-point mesh of 12 × 12 × 12 was used to sample the Brillouin zone.37 Lattice parameters were fixed to an experimentally measured value (c/a = 0.987), and Ms was calculated using the formula Ms=μtotalV, where μtotal is the absolute value of the total magnetization from DFT calculations given in Bohr magnetons and V is the unit cell volume.38 Self-consistent spin-polarized calculations were performed for collinear spin structures with no spin–orbit coupling term in the Hamiltonian. The Mn-atoms (Mn I) located in the cell corners with coordinates (0, 0, 0) were ferromagnetically coupled to Mn-atoms (Mn IIa) located at (0.5, 0.5, 0). Both Mn I- and Mn IIa-atoms were anti-ferromagnetically coupled to the Mn-atoms (Mn IIb) in the face centers with coordinates (0, 0.5, 0.5) and (0.5, 0, 0.5). Since the total atomic magnetic moments at all three Mn-sites do not cancel each other out (Mn I = 3.47 µB, Mn IIa = 0.75 µB, and Mn IIb = −2.36 µB), the ground state is a ferrimagnet.

Table I lists the five samples (S1–S5) that were examined in this study. We begin by investigating different substrate annealing temperatures to understand the impact of the MgO surface on the epitaxial growth of Mn4N. S1–S3 were deposited at 400 °C, and S4 and S5 were deposited at a higher temperature of 450 °C. The film thickness was determined by XRR measurement. In the case of S2–S5, the MgO substrates were pre-baked before loading into the sample deposition chamber.

TABLE I.

List of Mn4N thin films on MgO substrates.

MgO annealingMgO annealingFilmDeposition
Sample No.temperaturetimethickness (nm)temperature (°C)
S1 No annealing (RT) … 16.6 ± 0.6 400 
S2 1000 °C 4 h 16.8 ± 0.5 400 
S3 1100 °C 4 h 16.8 ± 0.5 400 
S4 1100 °C 4 h 16.8 ± 0.5 450 
S5 1100 °C 4 h 11.5 ± 0.6 450 
MgO annealingMgO annealingFilmDeposition
Sample No.temperaturetimethickness (nm)temperature (°C)
S1 No annealing (RT) … 16.6 ± 0.6 400 
S2 1000 °C 4 h 16.8 ± 0.5 400 
S3 1100 °C 4 h 16.8 ± 0.5 400 
S4 1100 °C 4 h 16.8 ± 0.5 450 
S5 1100 °C 4 h 11.5 ± 0.6 450 

The film thicknesses measured by the XRR technique indicated that the thicknesses of all five thin film samples were less than 20 nm (see Table I). The analysis of XRR results is shown in Fig. 1 of the supplementary material. S1–S4 had a thickness of 16.8 ± 0.5 nm, whereas S5 had a thickness of 11.5 ± 0.6 nm. Figure 2(a) shows the normalized M-H curves with the out-of-plane applied magnetic field for 16.8 nm Mn4N films on different heat-treated MgO substrates (S1–S3). The M-H curves show a strong dependence on the annealing temperature. All out-of-plane hysteresis loops were open, indicating PMA. The substrate annealing temperature of S3 (1100 °C) is higher than that of S2 (1000 °C), whereas S1 is unannealed. The squareness of the M(H) hysteresis loop improves from S1–S3. Figure 2(b) shows the dependence of the substrate annealing temperature on Mr/Ms. The Mr/Ms ratio increased from 0.47 to 0.72 as the annealing temperature increased from RT to 1100 °C. The increment of Mr/Ms is attributed to the improved epitaxial growth of the thin film. High-temperature annealing not only decomposes the Mg(OH)2 and MgCO3 contaminants but also reconstructs the surface of the MgO substrate.39,40 During this process, the substrate forms an atomically smooth surface,39,40 which is beneficial for epitaxial growth.41 The difference between S1 and S3 is mainly attributed to the different substrate annealing temperatures. Thin films annealed at 1100 °C produced a smoother surface than 1000 °C. The surface morphology as characterized by the atomic force microscopy revealed an average root-mean-square (rms) roughness of 0.592 nm, 0.308 nm, and 0.278 nm for the as-received, 1000 °C annealed, and 1100 °C annealed thin films, respectively (see Fig. 2 of the supplementary material).

FIG. 2.

(a) Normalized out-of-plane M(H) loops of 16.8 nm Mn4N films deposited at 400 °C on MgO substrates annealed at different temperatures: S1 (red) was unannealed, S2 (green) was annealed at 1000 °C, and S3 (blue) was annealed at 1100 °C; (b) Mr/Ms ratio of the 16.8 nm Mn4N film (S1–S3) deposited at 400 °C on MgO substrates, which have been annealed at different temperatures; (c) normalized out-of-plane M(H) loops of 16.8 nm (S4, cyan) and 11.5 nm (S5, brown) thick Mn4N films deposited at 450 °C on MgO substrates, which have been annealed at 1100 °C; and (d) out-of-plane and in-plane M(H) loops of the S4 Mn4N film.

FIG. 2.

(a) Normalized out-of-plane M(H) loops of 16.8 nm Mn4N films deposited at 400 °C on MgO substrates annealed at different temperatures: S1 (red) was unannealed, S2 (green) was annealed at 1000 °C, and S3 (blue) was annealed at 1100 °C; (b) Mr/Ms ratio of the 16.8 nm Mn4N film (S1–S3) deposited at 400 °C on MgO substrates, which have been annealed at different temperatures; (c) normalized out-of-plane M(H) loops of 16.8 nm (S4, cyan) and 11.5 nm (S5, brown) thick Mn4N films deposited at 450 °C on MgO substrates, which have been annealed at 1100 °C; and (d) out-of-plane and in-plane M(H) loops of the S4 Mn4N film.

Close modal

To explore further improvement, we increased the deposition temperature to 450 °C while retaining the substrate annealing temperature and time at 1100 °C and 4 h. It has been shown that the squareness of the Mn4N film can be improved by increasing deposition temperature.29We found that the Mr/Ms ratio of these samples, S4 and S5, increased further to near 1, as shown in Fig. 2(c). The difference between S4 and S5 is in the film thickness: The S4 and S5 thin films were 16.8 ± 0.5 nm and 11.5 ± 0.6 nm thick, respectively (see Table I). S5, the thinnest film, kept a high Mr/Ms ratio of ∼0.8, which is an encouraging outcome.

Figure 2(d) shows the out-of-plane and in-plane M-H curves for the S4 film. Ms and Ku for the S4 film are 43 ± 1 emu/cc and 0.70 Merg/cc, respectively. All five samples had similar Ms (40 emu/cc–60 emu/cc). Although Ms of these films are smaller than the reported Ms (∼100 emu/cc) in thicker films (30 nm–100 nm),22–25,28 they are comparable with that of the MBE-grown sub-20 nm Mn4N film on MgO, 50 emu/cc–80 emu/cc.26,27 The lower Ms of sub-20 nm Mn4N films may be due to surface oxidation.25 

Figure 3 shows the out-of-plane 2θ-θ XRD profile (a) and φ scan (b) of S3. Besides the MgO substrate peaks, only the Mn4N (002) peak is observed in the 2θ-θ XRD profile, which indicates the Mn4N (00l) orientation is parallel to MgO (00l). In the XRD 360° φ scan, both Mn4N and MgO show four peaks with 90° interval, which correspond to (202), (022), (−202), and (0–22). The overlap of Mn4N peaks and MgO peaks in φ-scan confirms their epitaxial relationship to be MgO (001)[100]//Mn4N (001)[100]. The out-of-plane lattice constant c and in-plane lattice constant a deduced from the (002) and (202) diffraction peaks are 0.386 nm and 0.391 nm, respectively. Therefore, the c/a ratio was calculated to be 0.987, which is close to the value reported earlier.22–28 

FIG. 3.

(a) 2θ-θ profile of S3 and the MgO substrate annealed at 1100 °C. (b) φ scan of Mn4N (101) and MgO (101) peaks.

FIG. 3.

(a) 2θ-θ profile of S3 and the MgO substrate annealed at 1100 °C. (b) φ scan of Mn4N (101) and MgO (101) peaks.

Close modal

Based on the experimental lattice constant, DFT predicts an Ms value of 153 emu/cc, which is larger than our experiment results. However, DFT is a zero-temperature electronic structure calculation. Li et al. have shown that Ms of the bulk Mn4N would decrease as a function of increasing measurement temperature.42 DFT also does not consider the possible effects of the mixing layer at the interface as well as defects. Future investigation will address the temperature dependence of saturation magnetization and magnetic anisotropy energy.

We have grown sub-20 nm ferrimagnetic Mn4N epitaxial thin films with high Mr/Ms ratios at 400 °C–450 °C substrate temperatures on MgO substrates by reactive sputtering. The quality of the epitaxial growth was optimized by ex situ pre-annealing of MgO substrates at temperatures as high as 1100 °C. The annealing was found to reduce the surface roughness of the MgO substrates at the atomic level, which was found to improve the quality of the out-of-plane magnetization hysteresis loops. The present results established the Mn4N thin film as a thermally stable ferrimagnet for further investigation as a promising skyrmion-based spintronic material.

See the supplementary material for the XRR results of the film and the AFM results of the MgO substrate surface.

This work was supported by the DARPA Topological Excitations in Electronics (TEE) program (Grant No. D18AP00009). The content of the information does not necessarily reflect the position or the policy of the Government, and no official endorsement should be inferred. Approved for public release; distribution is unlimited.

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
A.
Fert
and
F. N.
Van Dau
, “
Spintronics, from giant magnetoresistance to magnetic skyrmions and topological insulators
,”
C. R. Phys.
20
(
7–8
),
817
831
(
2019
).
2.
S.
Woo
,
K.
Song
,
H.
Han
,
M.
Jung
,
M.
Im
,
K.
Lee
,
K. S.
Song
,
P.
Fischer
,
J.
Hong
,
J. W.
Choi
 et al, “
Spin-orbit torque-driven skyrmion dynamics revealed by time-resolved X-ray microscopy
,”
Nat. Commun.
8
,
15573
(
2017
).
3.
K.
Hals
and
B.
Arne
, “
Spin-orbit torques and anisotropic magnetization damping in skyrmion crystals
,”
Phys. Rev. B
89
,
064426
(
2014
).
4.
N.
Romming
,
C.
Hanneken
,
M.
Menzel
,
J. E.
Bickel
,
B.
Wolter
,
K.
Bergmann
,
A.
Kubetzka
, and
R.
Wiesendanger
, “
Writing and deleting single magnetic skyrmions
,”
Science
341
,
6146
(
2013
).
5.
S. Z.
Lin
,
C.
Reichhardt
,
C. D.
Batista
, and
A.
Saxena
, “
Particle model for skyrmions in metallic chiral magnets: Dynamics, pinning, and creep
,”
Phys. Rev. B
88
,
019901
(
2013
).
6.
T.
Schulz
,
R.
Ritz
,
A.
Bauer
,
M.
Halder
,
M.
Wagner
,
C.
Franz
,
C.
Pfleiderer
,
K.
Everschor
,
M.
Garst
, and
A.
Rosch
, “
Emergent electrodynamics of skyrmions in a chiral magnet
,”
Nat. Phys.
8
,
301
304
(
2012
).
7.
A.
Fert
,
N.
Reyren
, and
V.
Cros
, “
Magnetic skyrmions: Advances in physics and potential applications
,”
Nat. Rev. Mater.
2
,
17031
(
2017
).
8.
K.
Everschor-Sitte
,
J.
Masell
,
R. M.
Reeve
, and
M.
Kläui
, “
Perspective: Magnetic skyrmions—Overview of recent progress in an active research field
,”
J. Appl. Phys.
124
,
240901
(
2018
).
9.
S.
Woo
,
K.
Litzius
,
B.
Krüger
,
M.-Y.
Im
,
L.
Caretta
,
K.
Richter
,
M.
Mann
,
A.
Krone
,
R. M.
Reeve
,
M.
Weigand
 et al, “
Observation of room-temperature magnetic skyrmions and their current-driven dynamics in ultrathin metallic ferromagnets
,”
Nat. Mater.
15
,
501
506
(
2016
).
10.
A.
Soumyanarayanan
,
M.
Raju
,
A. L.
Gonzalez Oyarce
,
A. K. C.
Tan
,
M.-Y.
Im
,
A. P.
Petrović
,
P.
Ho
,
K. H.
Khoo
,
M.
Tran
,
C. K.
Gan
 et al, “
Tunable room-temperature magnetic skyrmions in Ir/Fe/Co/Pt multilayers
,”
Nat. Mater.
16
,
898
904
(
2017
).
11.
F.
Büttner
,
I.
Lemesh
, and
G. S. D.
Beach
, “
Theory of isolated magnetic skyrmions: From fundamentals to room temperature applications
,”
Sci. Rep.
8
,
4464
(
2018
).
12.
L.
Caretta
,
M.
Mann
,
F.
Büttner
,
K.
Ueda
,
B.
Pfau
,
C. M.
Günther
,
P.
Hessing
,
A.
Churikova
,
C.
Klose
,
M.
Schneider
 et al, “
Fast current-driven domain walls and small skyrmions in a compensated ferrimagnet
,”
Nat. Nanotechnol.
13
,
1154
1160
(
2018
).
13.
K.
Ueda
,
A. J.
Tan
, and
G. S. D.
Beach
, “
Effect of annealing on magnetic properties in ferrimagnetic GdCo alloy films with bulk perpendicular magnetic anisotropy
,”
AIP Adv.
8
,
125204
(
2018
).
14.
F.
Nabki
,
T. A.
Dusatko
,
S.
Vengallatore
, and
M. N.
El-Gamal
, “
Low-stress CMOS-compatible silicon carbide surface-micromachining technology—Part I: Process development and characterization
,”
J. Microelectromech. Syst.
20
,
720
729
(
2011
).
15.
H.
Yang
,
A.
Thiaville
,
S.
Rohart
,
A.
Fert
, and
M.
Chshiev
, “
Anatomy of Dzyaloshinskii-Moriya interaction at Co/Pt interfaces
,”
Phys. Rev. Lett.
115
,
267210
(
2015
).
16.
X.
Ma
,
G.
Yu
,
X.
Li
,
T.
Wang
,
D.
Wu
,
K. S.
Olsson
,
Z.
Chu
,
K.
An
,
J. Q.
Xiao
,
K. L.
Wang
 et al, “
Interfacial control of Dzyaloshinskii-Moriya interaction in heavy metal/ferromagnetic metal thin film heterostructures
,”
Phys. Rev. B
94
,
180408(R)
(
2016
).
17.
A.
Hrabec
,
N. A.
Porter
,
A.
Wells
,
M. J.
Benitez
,
G.
Burnell
,
S.
McVitie
,
D.
McGrouther
,
T. A.
Moore
, and
C. H.
Marrows
, “
Measuring and tailoring the Dzyaloshinskii-Moriya interaction in perpendicularly magnetized thin films
,”
Phys. Rev. B
90
,
020402(R)
(
2014
).
18.
R.
Ramaswamy
,
J. M.
Lee
,
K.
Cai
, and
H.
Yang
, “
Recent advances in spin-orbit torques: Moving towards device applications
,”
Appl. Phys. Rev.
5
,
031107
(
2018
).
19.
C. T.
Ma
,
Y.
Xie
,
H.
Sheng
,
A. W.
Ghosh
, and
S. J.
Poon
, “
Robust formation of ultrasmall room-temperature Neél skyrmions in amorphous ferrimagnets from atomistic simulations
,”
Sci. Rep.
9
,
9964
(
2019
).
20.
Y.
Quessab
,
J.
Xu
,
C. T.
Ma
,
W.
Zhou
,
G. A.
Riley
,
J. M.
Shaw
,
H. T.
Nembach
,
S. J.
Poon
, and
A. D.
Kent
, “
Tuning interfacial Dzyaloshinskii-Moriya interactions in thin amorphous ferrimagnetic alloys
,”
Sci. Rep.
10
,
7447
(
2020
).
21.
W. J.
Takei
,
R. R.
Heikes
, and
G.
Shirane
, “
Magnetic structure of Mn4N-type compounds
,”
Phys. Rev.
125
,
1893
(
1962
).
22.
S.
Isogami
,
K.
Masuda
, and
Y.
Miura
, “
Contributions of magnetic structure and nitrogen to perpendicular magnetocrystalline anisotropy in antiperovskite ɛ−Mn4N
,”
Phys. Rev. Mater.
4
,
014406
(
2020
).
23.
K.
Kabara
and
M.
Tsunoda
, “
Perpendicular magnetic anisotropy of Mn4N films fabricated by reactive sputtering method
,”
J. Appl. Phys.
117
,
17B512
(
2015
).
24.
Y.
Yasutomi
,
K.
Ito
,
T.
Sanai
,
K.
Toko
, and
T.
Suemasu
, “
Perpendicular magnetic anisotropy of Mn4N films on MgO(001) and SrTiO3(001) substrates
,”
J. Appl. Phys.
115
,
17A935
(
2014
).
25.
A.
Foley
,
J.
Corbett
,
A.
Khan
,
A. L.
Richard
,
D. C.
Ingram
,
A. R.
Smith
,
L.
Zhao
,
J. C.
Gallagher
, and
F.
Yang
, “
Contribution from Ising domains overlapping out-of-plane to perpendicular magnetic anisotropy in Mn4N thin films on MgO(001)
,”
J. Magn. Magn. Mater.
439
,
236
244
(
2017
).
26.
T.
Gushi
,
M.
Jovičević Klug
,
J.
Peña Garcia
,
S.
Ghosh
,
J.-P.
Attané
,
H.
Okuno
,
O.
Fruchart
,
J.
Vogel
,
T.
Suemasu
,
S.
Pizzini
 et al, “
Large current driven domain wall mobility and gate tuning of coercivity in ferrimagnetic Mn4N thin films
,”
Nano Lett.
19
(
12
),
8716
8723
(
2019
).
27.
T.
Hirose
,
T.
Komori
,
T.
Gushi
,
A.
Anzai
,
K.
Toko
, and
T.
Suemasu
, “
Strong correlation between uniaxial magnetic anisotropic constant and in-plane tensile strain in Mn4N epitaxial films
,”
AIP Adv.
10
,
025117
(
2020
).
28.
X.
Shen
,
A.
Chikamatsu
,
K.
Shigematsu
,
Y.
Hirose
,
T.
Fukumura
, and
T.
Hasegawa
, “
Metallic transport and large anomalous Hall effect at room temperature in ferrimagnetic Mn4N epitaxial thin film
,”
Appl. Phys. Lett.
105
,
072410
(
2014
).
29.
K. M.
Ching
,
W. D.
Chang
,
T. S.
Chin
,
J. G.
Duh
, and
H. C.
Ku
, “
Anomalous perpendicular magnetoanisotropy in Mn4N films on Si(100)
,”
J. Appl. Phys.
76
,
6582
(
1994
).
30.
S.
Nakagawa
and
M.
Naoe
, “
Preparation and magnetic properties of Mn4N films by reactive facing targets sputtering
,”
J. Appl. Phys.
75
,
6568
(
1994
).
31.
GXRR Version 2.0.3.0, Rigaku Corporation, Tokyo, Japan.
32.
P.
Giannozzi
,
S.
Baroni
,
N.
Bonini
,
M.
Calandra
,
R.
Car
,
C.
Cavazzoni
,
D.
Ceresoli
,
G. L.
Chiarotti
,
M.
Cococcioni
,
I.
Dabo
,
A.
Dal Corso
 et al, “
QUANTUM ESPRESSO: A modular and open-source software project for quantum simulations of materials
,”
J. Phys.: Condens. Matter
21
,
395502
(
2009
).
33.
P.
Giannozzi
,
O.
Andreussi
,
T.
Brumme
,
O.
Bunau
,
M.
Buongiorno Nardelli
,
M.
Calandra
,
R.
Car
,
C.
Cavazzoni
,
D.
Ceresoli
,
M.
Cococcioni
 et al, “
Advanced capabilities for materials modelling with Quantum ESPRESSO
,”
J. Phys.: Condens. Matter
29
,
465901
(
2017
).
34.
D.
Vanderbilt
, “
Soft self-consistent pseudopotentials in a generalized eigenvalue formalism
,”
Phys. Rev. B
41
,
7892(R)
(
1990
).
35.
A.
Dal Corso
, “
Pseudopotentials periodic table: From H to Pu
,”
Comput. Mater. Sci.
95
,
337
350
(
2014
).
36.
J. P.
Perdew
,
A.
Ruzsinszky
,
G. I.
Csonka
,
O. A.
Vydrov
,
G. E.
Scuseria
,
L. A.
Constantin
,
X.
Zhou
, and
K.
Burke
, “
Restoring the density-gradient expansion for exchange in solids and surfaces
,”
Phys. Rev. Lett.
100
,
136406
(
2008
).
37.
H. J.
Monkhorst
and
J. D.
Pack
, “
Special points for Brillouin-zone integrations
,”
Phys. Rev. B
13
,
5188
(
1976
).
38.
N. A.
Spaldin
,
Magnetic Materials: Fundamentals and Applications
, 2nd ed. (
Cambridge University Press
,
United Kingdom
,
2011
).
39.
D. K.
Aswal
,
K. P.
Muthe
,
S.
Tawde
,
S.
Chodhury
,
N.
Bagkar
,
A.
Singh
,
S. K.
Gupta
, and
J. V.
Yakhmi
, “
XPS and AFM investigations of annealing induced surface modifications of MgO single crystals
,”
J. Cryst. Growth
236
,
661
666
(
2002
).
40.
A. H.
Youssef
,
G.
Kolhatkar
,
A.
Merlen
,
R.
Thomas
, and
A.
Ruediger
, “
Surface preparation and the evolution of atomically flat step terrace morphology of MgO single crystals
,”
AIP Adv.
8
,
095025
(
2018
).
41.
A. F.
Degardin
,
F.
Houze
, and
A. J.
Kreisler
, “
MgO substrate surface optimization for YBaCuO thin film growth
,”
IEEE Trans. Appl. Supercond.
13
(
2
),
2721
2724
(
2003
).
42.
C.
Li
,
Y.
Yang
,
L.
Lv
,
H.
Huang
,
Z.
Wang
, and
S.
Yang
, “
Fabrication and magnetic characteristic of ferrimagnetic bulk Mn4N
,”
J. Alloys Compd.
457
,
57
60
(
2008
).

Supplementary Material