Introduction

Recently, combining semiconducting and magnetic properties within the same material changed from being a concept1 to being one of the main progress vectors for spintronics2, in particular when combining magnetism with topological properties3. With the discovery of ferroelectric Rashba semiconductors (FERS)4, magnetic doping in dilute magnetic semiconductors (DMS) has opened pathways to exploit the electron spin associated with the Rashba-Zeeman type spin splitting of the electronic states5,6,7. To date the primary FERS representative is α-GeTe; with only two atoms per unit cell and ≈0.3 Å displacement along the \(\left\langle 111\right\rangle\) direction between the Ge and Te atoms, it is arguably the simplest room temperature ferroelectric semiconductor8,9. The combination of ferroelectric order and large spin-orbit interaction (SOI) yields a switchable Rashba-type spin structure of the bulk states10, whereby the states are fully spin polarised around the valence band maximum11. When doped with Mn, magnetic order is induced in Ge1−xMnxTe while the ferroelectric order remains for dopings below x = 25%, rendering it a multiferroic semiconductor with broken inversion symmetry6,12,13,14. Moreover, the collinear alignment of the magnetisation and ferroelectric polarisation axes in Ge0.87Mn0.13Te ensures magnetoelectric coupling in the system10. Conversely, as there is now a cohesive picture between Berry curvature and emergence of anomalous Hall effect in non-centrosymmetric magnetic semiconductors15, the current-induced magnetization switching mechanism in this material points to a bulk Rashba-Edelstein effect7. Both spectroscopic6,10 and transport studies7 indicate that the key ingredient for the magnetoelectric functionality of Ge1−xMnxTe is its carrier-induced magnetism16,17,18,19,20 mediated by the valence band6,7. Furthermore, multiferroic Ge1−xMnxTe is proposed to be a unique platform to explore novel phenomena such as nonreciprocal electric transport21 or magnetic skyrmionic polarons22.

In this work, we combine a range of spectroscopic techniques based on X-ray magnetic circular dichroism (XMCD), near-edge X-ray absorption fine structure spectroscopy (NEXAFS) and state of the art modeling (see Methods I–IV), to reveal that Ge0.87Mn0.13Te is a bulk ferrimagnet (FiM). This is corroborated by complementary low-energy muon spectroscopy (LE-μSR) as a local probe to characterize the temperature-dependent magnetic transitions on the nanometer scale and SQUID magnetometry to access the macroscopic magnetisation. The results testify that the FiM order builds up below Tc1 = 100 K on top of a paramagnetic (PM) background and that below Tc2 = 40 K the system behaves as a spin-glass. By developing new analysis tools based on magnetostochastic resonance (MSR) switching we address the Ge0.87Mn0.13Te magnetism in the static and dynamical regimes. The ensemble of experimental techniques and theory yields a holistic view of a correlated spin glass with magnetization switching that can be driven by nanoampere pulsed currents using stochastic resonance. In particular we show that this switching is collective due to glassiness, and that the magnetic switching can be turned on and off by small changes in pulse periodicity.

Results

Ferrimagnetic switching

Figure 1 a shows XMCD spectra measured at 10 K in total electron yield (TEY) mode while applying magnetic fields up to ±6 T. As will be discussed later, the dichroic signal does not saturate at our maximum B-fields due to a paramagnetic spin-glass state, similar to the Cd1−xMnxTe DMS system23. The top-inset is a zoom-in into the MnL3 absorption edge, showing two distinct XMCD features separated by ≈0.5 eV. A time series of XMCD spectra for selected B-fields are shown in Fig. 1b, revealing that these features fluctuate, eventually completely and spontaneously reversing at zero B-field (Fig. 1c). The shape of the dichroic signal at the MnL3-edge can be understood from the combination of a paramagnetic background superposed with ferrimagnetic contributions (Fig. 1d). The fluctuating spectral weight in Fig. 1b indicates that the switching continues for applied fields even above 4 T. From detailed analysis of an extensive data set obtained from several experimental runs and different samples, we find that the switching between the two magnetic orientations appears to be regular and does not show the telegraph noise expected for a system which randomly switches between two magnetic- or spin-states24,25. As surprising as the switching itself, is the fact that it occurs simultaneously, within the time resolution of the experiment, over the square millimetre area probed by the photon beam spot.

Fig. 1: Switching of ferrimagnetic order in XMCD.
figure 1

a XMCD spectra for selected applied B-fields. Top inset: zoom into the Mn L3 edge, showing two distinct spectral weight from substitutional Mns and interstitial Mni atoms. Bottom inset: background subtracted XAS spectra for ± μ light polarizations, measured at B = 6 T (10 K), and overlaid with multiplet calulations. b Series of MnL3 XMCD spectra taken at ≈6 min intervals for selected B-fields. c Spontaneous switching recorded in two consecutive XMCD datasets (red/blue) in the absence of applied B-field, shown at the MnL3 (main graph) and MnL2 (inset) absorption edges. d MnL3-edge model based on 4 Gaussian functions simulating Mns and Mni spectral contributions. For PM the Mns and Mni features are aligned, with their amplitude dependent on the B-field. The FiM consists of oppositely aligned peaks with different amplitude. M and MSW refer to the reference and switched state respectively.

To reiterate, the dichroism at MnL3 systematically shows two opposite magnetic moments from different types of Mn atoms, with absorptions at hν = 638.8 and 639.3 eV, simultaneously switching after about every energy scan across the MnL3 absorption edge, in the absence or presence of an externally applied field. Notably, the opposite magnetic moments are not exactly equal in magnitude, indicating a non-balanced ferrimagnetic order from two distinct Mn-sites. The spontaneous switching is further corroborated by the fact that we obtain an equivalent dichroism effect by measuring spectra with the same circular polarization (see Supplementary Fig. 1a). This also indicates that the light helicity is not the driving force for the switching. At this point, the following questions arise:

  • What is the difference between the two Mn-sites?

  • Why are their magnetic moments spontaneously reversing in time over macroscopic length scales?

  • What is the microscopic origin of the magnetization switching?

  • How to control the switching?

To address these questions, we first identify the different Mn contributions with our theoretical calculations, and then proceed to explain the spontaneous magnetic switching in terms of magneto-stochastic resonance, whereby periodic time fluctuations in the current associated with photoexcited electrons interfere with large TEY pulses during the XMCD measurements to yield a global resonant magneto-stochastic switching of the magnetisation. Finally, we show that the collective behavior originates from a glassy spin state.

Magnetic ground state properties

In accordance with our earlier resonant angle-resolved photoemission (ARPES) studies on Ge1−xMnxTe6 and further XMCD studies on other Mn-doped DMS systems26,27,28, we identify the magnetic moments in Ge1−xMnxTe to originate from substitutional (Mns) and interstitial (Mni) atoms. However, in contrast to other related systems, we find that in Ge0.87Mn0.13Te the characteristic ≈0.5 eV energy separation between Mns and Mni is not due to Mn segregation or non-magnetic surface Mn-oxides, nor to a depletion of Mn in the near-surface layer. Instead, the energy shift is an intrinsic bulk property that we also see in bulk-sensitive fluorescence yield and in NEXAFS analysis from the Mn-K edge (see Supplementary Fig. 1b, c, respectively). From these data, we estimate the presence of Mns and Mni atoms in the GeTe host lattice, with a Mns:Mni occupancy close to 2:1. Based on this experimental finding we calculate the Ge0.87Mn0.13Te magnetic ground state properties using density functional theory, where the Mn-disorder is implemented within the coherent potential approximation (CPA) alloy theory (see Methods).

Figure 2 describes how the distinct Mn incorporation into the GeTe lattice induces magnetic frustration due to the various competing exchange interactions, while the ferroelectric crystal structure is retained. For 13% Mn-doping we expect nearly 1 Mn per 8 unit cells as schematically depicted in Fig. 2a. Our CPA calculations are based on an “infinite” Ge0.87Mn0.13Te lattice consisting of 8.9% probability occupation of Mns substituted on Ge lattice sites (Mn1) and two types of interstitial Mni atoms, Mn2 and Mn3, each one with 2.2% probability occupation. The most prominent contribution to the magnetic order comes from the Heisenberg exchange energy between Mn2,3 and between Mn1,2/3 atoms, whereas the interaction between substitutional atoms (Mn1,1) is damped due to the presence of the interstitial Mn. The corresponding magnetic exchange couplings Ji,j are presented in Fig. 2b as a function of the distance Ri,j/a between atoms, where a is the lattice parameter and i, j = 1, 2, 3. The positive exchange constants (Ji,j > 0) favor ferromagnetic (FM) order, whereas negative values (Ji,j < 0) favor antiferromagnetic (AFM) order (see Methods).

Fig. 2: Frustrated magnetic interactions in Ge0.87Mn0.13Te.
figure 2

a Illustration of Mn doping in the rhombohedrally distorted α-GeTe unit cell with one substitutional Mns and two potential interstitial Mni dopants. b Exchange couplings Ji,j, between Mni-Mnj atoms with all three types of Mn atoms present (full markers), and with only substitutional Mn1 atoms (empty marker). Distances are normalized to the lattice parameter a. Inset: Extension to smaller atomic distances. c DMI values for inter- and intra-Mn site coupling. The Di,j,x component is summarized as xi, j and similar for y and z (x, y, z index directions are denoted in a). Dominant interactions are indicated in the graph, others listed in legends. d Comparison between XMCD experiment (red dotted line) and CPA theory, in which spectral contributions from individual Mn atoms result in a final XMCD signal (dashed line).

The co-existence of FM and AFM order in a DMS system is one of the main ingredients for magnetic frustration in a canonical spin glass, which in our case is further enhanced by Dzyaloshinskii-Moriya interaction (DMI). DMI is characteristic to non-centrosymmetric systems with large spin-orbit coupling (SOC) as clearly applies for Ge1−xMnxTe, where it promotes non-coplanar (canted) arrangements between spin states. As shown by Fig. 2c, DMI is highly anisotropic in both inter- and intra-Mn interactions, amongst which x2, 3, y2, 3 and z2, 2 appear as most prominent (see Methods). Both DMI and all Ji,j have fluctuating Ri,j/a dependences. Furthermore, the DMI is non-negligible compared to the Ji,j especially for larger length scales, leading to magnetic frustration at different length scales.

As shown in the inset of Fig. 2b, for short length scales there is a competition between FM and AFM contributions, the relative magnitudes of which explain why Ge0.87Mn0.13Te stabilizes in a ferrimagnetic ground state. Since CPA allows us to derive spectral contributions from individual Mn-atoms (Fig. 2d), the resulting theoretical magnetic dichroism reproduces the measured spectra very well if we take into account that the majority of Mns contribute to the PM background once the magnetic interactions stabilize, as detailed in Supplementary Fig. 2.

The described ferrimagnetic ground state, based on three distinct interacting Mn lattice sites, has been elusive in previous theoretical20,29 and experimental studies16,18,30 because only substitutional Mns-sites were considered (empty markers in Fig. 2b). The oscillatory long-range Ji,j exchange of the Mns-sites is damped due to the presence of interstitial Mni atoms (filled green markers in Fig. 2b). Furthermore, when probed under applied B-field (Fig. 1a), the opposite FiM contributions are placed on a large PM background due to which the Mni-moments partially align with those of Mns. Only the magnetization switching dynamics shown in Fig. 1b allows us to distinguish the two sites under applied magnetic field.

In order to assess the energetics of the FiM switching, we compare in Fig. 3a the energy scale of the rotation of Mni spins with respect to the Mns spin held fixed along the [111] axis by the magnetocrystalline anisotropy energy (MAE, see Methods). The combination of both energy profiles constitutes a double-well uniaxial potential with a MAE barrier of ≈0.1 meV, comparable to that of ferromagnetic 3d metals31. Having identified the general magnetic order and its bistable potential model, we now turn to the dynamics and the unexpected magnetic switching.

Fig. 3: Magnetostochastic resonance driving sources.
figure 3

a Stochastic resonance switching on a symmetric double-well qualitatively modelled by FiM exchange energy (grey downward triangles) and magnetocrystalline anisotropy energy (red upward triangles). The potential is subject to a synchronised periodic drive TΩ (red-yellow trace), enabling the switching of the otherwise stable FiM state. b FiM switching cartoon: the circles illustrate the MSR harnessing concept with periodic MnL3 TEY pulses indicated by grey and red/blue flash symbols, respectively. Periodic pulses with entrained P(t) signal (red-yellow circle) enable the FiM switching in a statistical average. c Sinusoidal fit (red-yellow trace) of a time evolution of normalized TEY periodic drive P(t) measured at 10 K, B = 0, and hν = 638.8eV (grey trace). Superimposed red and blue spikes represent transient TEY pulse at the MnL3 edge that would occur during consecutive XAS1, XAS2, etc. scans. d Typical time series of periodic (green/top) and aperiodic (grey/bottom) XAS scans which controls the magnetostochastic resonance switching (MSR-on/off). Empty spaces between the XAS scans is the time needed to move the monochromator back to the initial energy.

Magnetostochastic resonance switching

The dynamical spontaneous magnetisation reversal in our system is explained by the mechanism of stochastic resonance (SR). Generally speaking, SR is a phenomenon applicable to nonlinear systems whereby a weak signal is amplified through an entraining periodic signal or even noise. Besides the bistable potential, SR requires two additional ingredients32: (i) a weak periodic input and (ii) a source of noise, or amplitude modulated signal33 entrained with the periodic input. As will be elaborated in detail below, the first ingredient is provided by periodic TEY oscillations P(t) upon Ge0.87Mn0.13Te x-ray illumination, whereas the second η(t) originates from the short TEY pulse in x-ray absorption during the scan over the MnL3 absorption edge, both subjected to thermal noise. The sum of these induce the observed transition between the two FiM equilibria as depicted in Fig. 3a, b.

The periodic drive P(t) represents a continuous sinusoidal signal with TΩ ≈ 5 min (Fig. 3c). These systematically observed oscillations are present also while performing hν scans; they are not related to the periodic top-up current refill of the synchrotron and do not depend on x-ray polarization and applied B-field (see Supplementary Fig. 3). Therefore, our understanding is that they originate from charging and discharging inside a capacitor circuit with Ge0.87Mn0.13Te as dielectric, whereby the primary charging effect is caused by charge separation due to a steady-state bulk photovoltaic effect typical to materials with broken inversion symmetry34. For the GeTe host material this shift current effect is even stronger compared to conventional oxide ferroelectrics35. Moreover, the system possesses self-poling properties36 and negative capacitance37, leading to thermodynamically unstable transient effects38.

The photoelectrons contributing to the TEY current leave the sample following the shortest path. On average this is along the normal to the sample plane, i.e. along the (Ge,Mn)Te [111] direction. This current direction is collinear with both the perpendicular magnetic anisotropy, as well as the ferroelectric polarisation. In line with the exertion of SOT by the inverse Rashba-Edelstein effect7, the SOT-induced switching in our case is triggered by the TEY current, especially the short pulse η(t) at the MnL3-edge (Fig. 3c). In fact, in the dichroic signal we never observe anomalous or discontinuous spectral shapes, indicating that the switching event most likely occurs directly after the MnL3-edge.

Harnessing the magnetostochastic resonance switching

The phenomenological description of magnetostochastic resonance (MSR) applied to magnetization dynamics typically predicts a million times higher switching probability for bistable magnetic systems39. Consistent with these predictions, even though the geometry and time scales are different, for Ge0.87Mn0.13Te the current-driven magnetisation switching in the milliampere regime7 is achieved under MSR with a nanoampere TEY pulse at the MnL3 absorption edge, and thus with unprecedented efficiency of low-current density pulses compared to typical spin-orbit torque systems. The TEY current is typically in the 10-100 nA range. In ref. 7 the typical current density used to switch the magnetization is ≈6 × 106 Acm−2. This huge current density is applied within a pulse current of 1 ms. On the other hand, in our case the TEY current from a sample area of ≈1 mm2 amounts to 100 × 10−9/(0.1 × 0.1) ≈ 1 × 105 Acm−2, within a period of time during the on-the-fly energy scan40 across the Mn L3-edge (ca.1–2 s). Considering the pulse duration, a coarse estimate between the two experiments yields a current density ratio of 6 × 106/(2000*10−5) ≈ 3 × 108. This indicates that for the magnetic switching, the current density per unit time is in our case >6 orders of magnitude lower compared to ref. 7. It should be noted that in conventional slow point-by-point scans the FiM switching is not observed. This underlines the relevance of dynamics in the magnetostochastic resonance switching.

A typical measurement series is represented by the green top traces in Fig. 3d, where the resulting TEY signal is a two-frequency system in which P(t) realizes frequency modulation and η(t) implements amplitude modulation as a deterministic additive noise. These two driving frequencies, schematically depicted in Fig. 3c, exemplify stochastic resonance in two-frequency signal systems33. With such a combination of periodic and aperiodic driving frequencies the SR can be harnessed, e.g. controllably suppressed by detuning the frequencies from their optimal setting, as schematically depicted by aperiodic gray flash symbols in Fig. 3d. We achieved this by measuring XMCD data with different photon energy ranges (grey bottom traces in 3d), thereby moving away from the optimal synchronization with driving frequency \({t}_{1}={t}_{2}=\frac{1}{2}{t}_{{{\Omega }}}\)32, where t1,2 is the time needed to measure one XAS spectrum, and tΩ one full XMCD dataset (i.e. two XAS spectra measured with opposite circular polarisations).

Figure 4 summarizes the MSR harnessing with two synchronized loops with aperiodic (MSR-off) and periodic (MSR-on) additive noise η(t). The data indicates that there is an immediate change from the MSR-off regime without FiM switching to FiM switching under the MSR-on regime. We emphasize that under MSR-on, the FiM switching holds only in the statistical average, quantified via the residence time distribution N(t)32. We quantified N(t) by counting the multiples of tΩ between subsequent FiM switching occurrences, leading to a sequence of Gaussian-like peaks with exponentially decreasing envelope as seen in Fig. 4d at 5 K. In qualitative agreement with SR predictions32, the peak distributions under thermal noise shifts to shorter times, as seen at 10 K. The N(t) distribution envelope is found to depend on the resonance conditions through the driving period tΩ (see Supplementary Fig. 4a), which further confirms the MSR origin of the energy efficient FiM switching with drastically reduced currents.

Fig. 4: Harnessing switching dynamics with magnetostochastic resonance.
figure 4

a Stacked plot of MnL3 XMCD spectra measured in MSR-off regime directly followed by MSR-on regime. b XMCD MSR-off spectra averaged for individual applied B-fields as depicted in the inset. Top cartoon illustrates that under aperiodic MnL3 TEY pulses the MSR switching is disabled. c Stacked plot of MSR-on XMCD spectra. In (a) and (c) blue/red traces reflect the original/switched FiM state, grey traces reflect totally quenched magnetic state. d Statistical FiM switching quantification by residence time distributions N(t), fitted with Gaussian functions at 5 and 10 K.

The observed SR switching allows us to sharpen the microscopic understanding of the Ge0.87Mn0.13Te magnetic order. To this end we analyze the orbital mL and spin mS contributions from sum rules41,42. Because the magnitude and sign of the MSR-on XMCD signal changes in time, the directional sense of the magnetic moments with respect to the beam axis is changing as well, occasionally yielding a totally quenched moment Mtot = mS + mL such as in the grey traces in Fig. 4c. To avoid uncertainties imposed by this, we therefore focus on the MSR-off regime for the B-field dependence of mL and mS contributions, as summarized in Fig. 5. The theoretically obtained mL of -0.032 μB/(Mn atom) is independent of the magnitude of the applied field and the total magnetisation Mtot is expected to be dominated by mS, which takes a value of mS ≈ 2.1 μB/(Mn atom) at saturation. This trend is well observed in Fig. 5a where the applied B-field drives mS, whereas mL is rather constant. That the local Mn-moments deduced from the sum rules (0.13 ± 0.02 μB/Mn atom) for B = 0 differ from the theoretical values is due to the fact that most Mn-moments from the PM background do not contribute to the measured moments. Table 2 in the Methods section compares the calculated Mn-moments with experimental data in the MSR-off regime.

Table 1 Ge0.867Mn0.133Te R3mR (No. 160) space group crystal structure with atomic positions and occupancies
Fig. 5: Microscopic and macroscopic magnetisation in XMCD and SQUID.
figure 5

a Time series of spin (mS) and orbital (mL) moments showing their dependence on applied B-field under the MSR-off conditions after zero-field cooling (ZFC). b Out-of-plane SQUID hysteresis for Ge0.87Mn0.13Te grown on InP(111).

Table 2 Spin (mS) and orbital (mL) magnetic moments obtained from CPA theory vs XMCD

The macroscopic perspective of the magnetic order in Ge0.87Mn0.13Te is given by SQUID magnetometry as shown in Fig. 5b. A clear hysteresis loop is observed with a narrow waist and low coercive field. A remarkable feature are the dip-peak structures around zero field indicated by arrows. Similar non-monotonic features were observed in the magnetic field dependence of the Ge0.87Mn0.13Te anomalous Hall component, which were attributed to noncoplanar spin textures with scalar spin chirality7. This consideration is supported by the magnetic frustration due to different Ji,j and DMI length scales summarized in Fig. 2c, d. Interestingly, the non-monotonic features in SQUID, appearing under decreasing applied B-field, suggest the possible formation of frustrated spin textures like magnetic skyrmions43.

Correlated spin glass

In addition, the temperature dependence of the magnetization shown in Fig. 6b does not follow the typical \({(1-{(T/{T}_{c})}^{\alpha })}^{\beta }\) magnetization dependence with a single transition temperature, but rather a steadily diminishing magnetization with various critical temperatures indicated in the figure. The splitting between zero-field-cooled (ZFC) and field-cooled (FC) curves is a further clear indication of competing energy scales and spin orders, and of possible spin excitations in magnetic glassy-phases44. LE-μSR is the method of choice to characterize magnetically ordered volumes inside such a spin-glass from a local point of view.

Fig. 6: Correlated spin glass state observed by LE-μSR, SQUID, and switching dynamics.
figure 6

a Temperature-dependent average local field B (top), muon damping rate λ (middle), volume fraction of the paramagnetic regions (bottom), obtained from fits of the LE-μSR data. Error bars show statistical errors as obtained from the Musrfit program. b Temperature dependent comparison of the μSR local mean field, bulk magnetization M measured by SQUID magnetometry, and XMCD switching statistics for ZFC (green markers) and FH-FC (red markers). The pictographs illustrate a muon stopping in FiM clusters embedded in a PM sea. As the clusters grow in size at lower temperatures, the average internal B-field shows a negative shift indicated by the arrow in (a). The FiM switching statistical error margins are binomial confidence intervals (obtaining x-switches in n trials gives a \(\sqrt{(x/n*(1-x/n)/n)}\) error). The statistically most relevant FiM switching, summarized from three independent beamtimes from fresh samples, was measured at 5 K (ZFC), that is why the smallest error bar in FiM switching probability is achieved under these measurement conditions.

In Ge0.87Mn0.13Te we find that muons stopping in magnetic regions of the sample depolarize almost immediately and do not contribute to the measured signal. However, muons that land in paramagnetic regions experience predominantly the applied field (≈5 mT). The polarization of these muons can be used to extract the three quantities presented in Fig. 6a, namely, (i) the local mean field B sensed in the paramagnetic regions (top); (ii) the damping rate λ of the oscillations which is proportional to the width of local field distribution (middle); and (iii) the volume fraction of the paramagnetic regions (bottom). All three parameters start to change at the onset of magnetic order (≈100 K). Note that the paramagnetic (magnetically ordered) volume fraction decreases (increases) gradually at the onset of magnetic order, clearly showing the gradual nature of this transition. This is accompanied by a decrease in B, indicating that the magnetic regions have a net magnetic moment producing an increasing demagnetizing fields in the paramagnetic regions with decreasing temperatures. Below ≈40 K, most of the sample volume becomes magnetic, i.e., a long range magnetic order is established. The combination of bulk magnetization and LE-μSR data confirms the presence of a net magnetic moment imposed by the FiM order below Tc1 ≈ 100 K, followed by a spin-glass transition below Tc2 ≈ 40 K, where the coupling between the FiM clusters becomes strong enough for the sample to undergo a transition to a cluster spin-glass long range order.

The combined set of experimental and theoretical results allows us to rationalize the MSR switching mechanism. For a broader perspective Fig. 6b compares the temperature dependence of μSR local mean field, bulk magnetization, and XMCD switching statistics. As expected with the onset of the long-range order, the spin-freezing temperature Tc2 marks a minimum rather than an onset in FiM switching probability. Strikingly, above Tc2, the switching events dramatically increase along with the decreasing size of the FiM clusters, as highlighted by the red shaded region. This trend appears to be independent on whether the system was magnetized under ZFC or field-heated field-cooling (FH-FC) indicated by green and red markers, respectively. Moreover, at T = 65 K the XMCD asymmetry sometimes reaches values as high as 60% in contrast to the expected increase of magnetic disorder at elevated temperatures (Supplementary Fig. 4). This is reminiscent of the spontaneous thermally induced magnetic order in elemental spin glasses45, which in our case begins to morph between individual FiM clusters as indicated by pictographs in Supplementary Fig. 4b. The magnetic reordering on a mesoscopic scale and the magnetic frustration highlighted in the theory section are therefore related to the glassy behavior. Importantly, our XMCD results show that the FiM switching with nanoampere currents is collective over macroscopic dimensions. This indicates that the magnetic order has to be correlated across local magnetic interactions on which the magnetic spin textures spread like a collective excitation through the sample. This is exactly what is expected for a correlated cluster spin-glass46,47. Thus, our experimental results illustrate efficient collective spin dynamics in a correlated spin glass.

In conclusion, the multiferroic Ge1−xMnxTe system presents an outstanding platform for a great variety of complex intriguing physical phenomena which we decipher using an holistic approach: XMCD and theory confirm a bistable ferrimagnetic ground state, whereas the collective switching of this state in XMCD, together with μSR and magnetometry, indicate that this switching shapes up inside a cluster spin glass. For such systems we have demonstrated a unique new concept to harness magnetostochastic switching, which allowed us to describe novel magnetization dynamics. Under this MSR drive the current needed for switching of the magnetic state is reduced by many orders of magnitude compared to typical spin-orbit torque systems, allowing for the design of energy efficient non-volatile electronics. In such a design, the synchrotron radiation induced currents would be replaced by electric currents generated from electrical contacts or by using a laser source. On a more fundamental level, our results open up new avenues for studying collective spin dynamics in correlated spin glass systems.

Methods

Sample preparation

We have grown 200 nm Ge0.87Mn0.13Te thin films on InP(111) and BaF2(111) substrates by molecular beam epitaxy. The choice for 13% Mn-doping is that for such a doping the Zeeman splitting is saturating6. A protective stack of amorphous Te- and Se-capping layers with a total thickness of ≈20 nm was used to avoid surface oxidation and degradation. Before mounting into the XTREME beamline sample holder48, it was completely removed in a ultrahigh vacuum chamber by annealing the samples for 30–45 min at 250 °C; and recapped with ≈2 nm amorphous carbon by carbon-thread evaporation within the same vacuum chamber.

Experimental methods

The XMCD setup at the XTREME beamline48 allows us to measure magnetic moments from the averaged orbital and spin contribution projected along the out-of-plane magnetization M, which is collinear to the ferroelectric polarization P along the [111] direction6. Raw TEY and TFY signals measured with Keithley 428 current amplifiers were normalized with the I0 photocurrent from the XTREME beamline refocussing mirror. All the signals were sampled from the current amplifier analog outputs, and further processed with a software control system to synchronize the on-the-fly scan for absorption measurements with the undulator40. All XMCD spectra, except those in Supplementary Fig. 1, were obtained by taking the difference of (μ+ - μ-), where μ+ and μ- represent the XAS spectra measured by TEY, or TFY, with right- and left-handed circular light polarisation. The XMCD measurements were recorded as on-the-fly scans for a given magnetic field up to 6 T either parallel or anti-parallel to the beam direction. All data presented were measured during three independent beamtimes on fresh samples prepared in exactly the same way. The same samples were used for further NEXAFS (SuperXAS beamline at Swiss Light Source) and SQUID magnetometry.

Special extra-large samples capped with Au were prepared for low energy muon spin relaxation experiment (LE-μSR) at the μE4 muon beamline at the Paul Scherrer Institut49. LE-μSR time spectra are measured with two positron detectors placed opposite to each other for the considered samples. Temperature dependent asymmetry spectra were obtained under an external magnetic field of 5 mT applied perpendicular to the initial muon spin polarization direction. The data indicate magnetically ordered islands with a net magnetic moment aligned with the applied field, embedded in a sea of paramagnetic environment. As these islands grow in size, their magnetic moment along the applied field increases and the strength of their dipolar field in the paramagnetic regions increases. This dipolar field is primarily opposite to the direction of the applied field and hence produces the observed negative shift in B(T) as seen in Fig. 6a. At the lowest temperatures, the remaining 5 mT oscillation is not due to the sample. Instead, it originates from the background contribution50.

In order to achieve comparable local probe conditions with the XMCD photoelectron escape depth (≈10 nm), the muon implantation energy was adjusted to set their stopping distribution just below the sample surface. Under magnetization, the asymmetry of the transverse field the muon precession decays. The main quantity determining the LE-μSR spectra in Fig. 6a is the muon spin precession amplitude decay from asymmetric emission of positrons, analyzed with Musrfit program51.

Crystal structure and computational details

Using first-principles calculations within the density functional theory, we have investigated the ground state magnetic properties of bulk Ge0.867Mn0.133Te. The calculations were carried out using the multiple scattering KKR Green function method as implemented in the spin-polarized fully relativistic Korringa-Kohn-Rostoker (SPRKKR) code52. The exchange and correlation effects were incorporated within the LDA framework53. Brillouin zone integrations were performed on a 39 × 39 × 39 dense mesh of k-points. The angular momentum expansion up to l\({}_{\max }\) = 4 has been used for each atom. The energy convergence criterion and coherent potential approximation (CPA) tolerance has been set to 10−7 Ry. The potential is constructed in full potential geometry. The influence of chemical disorder in Ge0.867Mn0.133Te can be estimated by means of the CPA alloy theory54,55 implemented in the SPR-KKR method.

We consider two types of Mni atoms denoted as Mn2 and Mn3 in Fig. 2a, which occupy allowed Wyckoff positions within the Ge0.87Mn0.13Te space group (see Table 1), derived from the experimental data as determined from the room temperature X-ray powder diffraction measurements56.

In addition we also performed 2 × 2 × 2 super-cell Quantum ESPRESSO pseudopotential relaxation calculations with the Mn atomic positions summarized in Table 1. The results confirmed that AFM is the energetically most stable configuration. The structural models from the super-cell calculations were used to model XAS and XMCD calculations, which in turn where consistent with our CPA model summarized in Fig. 2d. Moreover, the structural relaxations around the Mn atoms showed negligible impact on the magnetic structure and shape of the XAS spectra.

Finally, we also performed multiplet calculations based on localized Mn2+3d with parameters reported by Sato et al.17. The comparison with measured Ge0.87Mn0.13Te XAS spectra in Supplementary Fig. 1c and Fig. 2a confirms that our XAS and XMCD spectra feature localized Mn2+3d states. Within the SPRKKR multiple scattering methodology we consider an average of 4.72 holes for Mns and 3.73 holes for Mni unoccupied d states with a correction factor of 1.47 for compensating the jj-mixing57. However, as reported by Sato et al.17, multiplet calculations suffer from intensity deviations between the calculations and experiments. For a better comparison with experimental data we thus relied on CPA calculations.

Because the CPA enables us to address spectral contributions from individual Mn-atoms, we tentatively modelled the resulting magnetic dichroism as a simple sum of all Mn contributions in saturation, as shown in Supplementary Fig. 2d. When reducing the Mns spectral contribution by a factor of 3, we find that the resulting XMCD signal (dashed line in Fig. 2d) reproduce the measured data very well, whereas multiplet calculations with \({{{{{{{\rm{M{n}}}}}}}_{s}^{2+}}}\) atoms fail to reproduce the exact shape of the dichroism effect and lead to a spectral shift indicated by the black arrow (see Supplementary Fig. 2d). The factor 3 reduction of the Mns spectral contribution is rationalized by observing a significant decrease in Mns dichroism between the states right after the FH-FC@6T at 10 K, and recorded later during MSR-on conditions (vertical arrow in Supplementary Fig. 2b). With every FiM switching event at low temperatures some of the Mns becomes part of the paramagnetic spin glass background; and when around one third of the original dichroism effect is left, this process appears to saturate (Supplementary Fig. 2c). However, on increasing the temperature from 10 → 45-65 K, we observe a significant increase in the Mns dichroism, which we attribute to self-induced magnetization stimulated by MSR along with the paramagnetic spin glass background as schematically depicted in Supplementary Fig. 4b.

Magnetic ground state calculations

Figure 2 a is a cartoon description of various Mn-Mn exchange interactions between individual S1, S2 spin states considered in our model Hamiltonian, where Mn impurities act at the same time as random magnetic moments and as acceptors producing the carriers. For the Ge0.87Mn0.13Te magnetic ground state we consider the general spin Hamiltonian in the following form:

$$H=-\frac{1}{2}\mathop{\sum}\limits_{ij}{J}_{i,\, j}{{{{{{{{\bf{S}}}}}}}}}_{{{{{{{i}}}}}}}\cdot {{{{{{{{\bf{S}}}}}}}}}_{{{{{{{j}}}}}}}+\mathop{\sum}\limits_{ij}{{{{{{{{\bf{D}}}}}}}}}_{{{{{{{i,\, j}}}}}}}\cdot ({{{{{{{{\bf{S}}}}}}}}}_{{{{{{{i}}}}}}}\times {{{{{{{{\bf{S}}}}}}}}}_{{{{{{{j}}}}}}})+\mathop{\sum}\limits_{i}{K}_{i}{{{{{{{{{\bf{S}}}}}}}}}_{{{{{{{i}}}}}}}}^{2}$$
(1)

The first term represents the Heisenberg exchange energy with Si and Sj unit vectors having directions corresponding to local magnetic moments on sites i and j. Ji,j > 0 and Ji,j < 0 prefer FM and AFM spin configurations, respectively. The second term originates from the antisymmetric part of the interaction matrix, also termed as antisymmetric exchange or Dzyaloshinskii-Moriya interaction (DMI). Finally the Ki term in the final term accounts for the magnetocrystalline anisotropy energy (MAE).

The isotropic Ji,j exchange between localized S1, S2 spin states depends on the distance between the next-nearest Mn dopants. In contrast to earlier theoretical Ge1−xMnxTe studies considering only Mns atoms20, another type of interaction is coming from the ∑i,jDi,j (Si × Sj) Dzyaloshinskii-Moriya interaction obtained by sampling the magnetization deflection between [111] and [\(\overline{111}\)] easy magnetization axis directions. Below we discuss the contributions of individual terms to the Ge0.87Mn0.13Te magnetic ground state.

Isotropic exchange Interactions

The magnetic exchange coupling parameters (Ji,j) are based on the real space approach by using the theory proposed by Liechtenstein et al.58. This approach employs the “magnetic force theorem” to determine Ji,j by assessing the total energy change related to an infinitesimal rotation of the magnetic moments located at the atomic sites i and j. The energy change can be related to the exchange coupling parameters Ji,j as:

$${J}_{i,j}^{}=\frac{1}{4\pi }{\int}^{{E}_{F}}dEImT{r}_{L}\{{{{\Delta }}}_{i}^{}{\tau }_{\uparrow }^{i,j}{{{\Delta }}}_{j}^{}{\tau }_{\downarrow }^{j,i}\}$$
(2)

where τ is the scattering path operator, Δi is the difference in the inverse single site scattering t matrices for spin up and spin down electrons, and TrL is the trace of scattering matrices over the orbital indices L = (l, m).

The Ji,j calculations are performed within a cluster of radius 3a, where a is the lattice parameter. We neglect all interactions involving Ge and Te atoms and consider only those between Mn atoms which are found to host significant localized magnetic moments/atom. Figure 2a of the main text shows three different types of Mn atoms inside the Ge0.867Mn0.133Te primitive unit cell, classified as substitutional Mns and interstitial Mni atoms, respectively.

The physical mechanism behind the magnetic ground state has been attributed to pd-exchange coupling20,59, combined with d-d magnetic interactions17. Consistently with our theoretical predictions, the pd-exchange coupling is relatively weak, but long ranged. Moreover, for Ge1−xMnxTe the pd-exchange constant is negative17, stemming from AFM exchange interaction between Mn2+ states. In agreement with our Ge1−xMnxTe magnetic ground state description, this negative term naturally materializes between Mni-Mns which further confirms that Ge1−xMnxTe is a ferrimagnetic rather than a ferromagnetic DMS16,17,18,19,20.

Dzyaloshinskii-Moriya interaction

The DMI is an anisotropic chiral exchange interaction between localized spins and has a net contribution only in systems without structural inversion symmetry and with the presence of an indirect or long-range exchange interaction. Ge0.87Mn0.13Te satisfies these two conditions, which is also evident from the isotropic exchange interactions described in the previous section. The origin of DMI is found in the spin-orbit coupling (SOC) which acts as a perturbation on localized orbital states. Given two neighboring spins S1 and S2, the DMI energy can be described by -D1,2 (S1 × S2). Therefore, the direction of the relative S1 and S2 rotations can be clockwise or counter-clockwise, providing the information about helicity in the case of spin spirals. This expression is part of a generalized exchange interaction where the DMI term is related to the exchange constant J of the direct Heisenberg exchange − J(SiSj). However, contrary to the latter which favors collinear alignment, the DMI promotes an orthogonal arrangement between Si and Sj, with a chirality imposed by the direction of Di,j. The resulting spin helicity is uniform, meaning that clockwise or counterclockwise rotation of spins are energetically identical. Data in Fig. 2c show all the components of Di,j (i.e. Di,j,x,Di,j,y,Di,j,z) interaction, showing robust exchange up to a cluster size of 3.0a. The average DMI energy is comparable to the Ji,j exchange energy for larger distances; and a factor 4 higher than the magnetic anisotropy energy (MAE) discussed below.

The effective ratio between the exchange interaction and DMI for a given pair i, j was calculated in the mean field approach as

$${\left(\frac{J}{D}\right)}_{{{{{{{{\rm{eff}}}}}}}}}=\sum \left({C}_{i}{C}_{j}\frac{{J}_{i,j}}{{D}_{i,j}}\right)$$
(3)

where the summation runs over all the neighbours up to a cluster of radius 3.5a where a is the lattice parameter. The factors Ci and Cj represent the occupation of the Mn atom at a given site. This yields 19.933, 20.7919, and 2.4275 for the Mn pairs i, j = 1, 1, 1, 2, and 2, 3, respectively.

Magnetic anisotropy energy

The magnetic ground state of FiM exchange energy (grey markers in Fig. 3a) has been explored by determining the total energy corresponding to the relative spin orientation of Mn2 and Mn3 spins with respect to the substitutional Mn1 spins. The Mn1 spin was frozen along the [111] quantization axis and the Mn2, Mn3 spins were rotated along the out-of-plane [111] direction starting from θ = 0, which corresponds to the FM state with all spins pointing in the same direction. However, the self consistency convergence in calculations can be achieved only for a limited θ range between 40°-320°. Nevertheless, the total energy of the system reaches a minimum with Mn1 spins set to 0° and Mn2,3 spins to 180°, which confirms the FiM order. The red markers in Fig. 3a show the MAE energy variation as a function of magnetization angle θ, evaluated as a difference between the fully relativistic total energies calculated for quantization axes [111] and axis orthogonal to \(\bar{[111]}\). Both FiM exchange and MAE consider a magnetoelectric ground state with ferroelectric distortion and the easy magnetization axis alignment along the [111] or [\(\bar{1}\bar{1}\bar{1}\)] directions. The calculated MAE of the Ge0.867Mn0.133Te is 0.101 meV/f.u. and represents the energy barrier between the easy and hard magnetic axes. Finally, the energy landscape obtained by combination of MAE and FiM ground state was used to model the FiM double-well potential as depicted in Fig. 3a. The MSR-driven FiM switching contracts the states into one of the two wells. The bistable potential depicted can be expressed as \(U(x)=-\frac{1}{2}a{x}^{2}+\frac{1}{4}b{x}^{4}\) with minima located at ± 90° and barrier height ΔV = a2/4b.