CoPS3 stands out in the family of the van der Waals antiferromagnets XPS3 (X = Mn, Ni, Fe, and Co) due to the unquenched orbital momentum of the magnetic Co2+ ions, which is known to facilitate the coupling of spins to both electromagnetic waves and lattice vibrations. Here, using a time-resolved magneto-optical pump–probe technique, we experimentally study the ultrafast laser-induced dynamics of mutually correlated spins and lattice. It is shown that a femtosecond laser pulse acts as an ultrafast heater and, thus, results in the melting of the antiferromagnetic order. At the same time, the resonant pumping of the 4T1g4T2g electronic transition in Co2+ ions effectively changes their orbital momentum, giving rise to a mechanical force that moves the ions in the direction parallel to the orientation of their spins, thus generating a coherent Bg phonon mode at the frequency of about 4.7 THz.

Since the seminal discovery of ultrafast demagnetization in Ni,1 the ultrafast manipulation of magnetism with ultrashort pulses of light has evolved into a fascinating research topic of nonequilibrium magnetism, with examples ranging from excitation of collective magnetic modes2–6 to light-driven magnetic phase transitions7–9 and switching of spin orientation.10–12 The recent resurgence of interest in two-dimensional (2D) van der Waals (vdW) materials hosting intrinsic long-range magnetic orders has offered a novel playground for investigating these phenomena in systems where the interplay between structural and magnetic orders plays a pivotal role.13–17 Understanding the nonequilibrium dynamics in vdW magnets particularly promises to provide important insights into the fundamental physics of spin–lattice interactions and spin relaxation in ultimately thin magnets. Moreover, the inherently strong light–matter interactions, typical of vdW materials,13,18–24 open up exciting possibilities for efficient manipulation of magnetism on the ultrafast timescale, promising novel energy-efficient data processing devices for future spintronics and magnonics applications.25,26

Among the various vdW magnets, transition metal thiophosphates, XPS3 (X = Fe, Ni, Mn, and Co), form a unique class of 2D antiferromagnets (AFMs) with intralayer AFM order on a honeycomb lattice. In addition to the 2D AFM order, previous studies of XPS3 have uncovered strongly coupled spin and charge orders,27,28 highly anisotropic excitons,29 magneto-electric coupling,30,31 signatures of the Berezinskii–Kosterlitz–Thouless (BKT) transition,32 and strong electron correlations.33,34 The family has also recently attracted a lot of attention for its ultrafast control of magnetism. In particular, it has been shown that a sudden perturbation of electron orbital momentum via resonant pumping of specific electronic transitions opens up new ways to control spins and lattice at an ultrafast timescale.19,20,34,35 CoPS3 stands out in the XPS3 family due to the Co2+ ions, characterized by the large spin and unquenched orbital momentum. The presence of such ions in magnetic materials usually results in a strong coupling of the spins to the lattice,36,37 large magnetocrystalline anisotropy,38 high frequencies of magnetic resonance,39 and strong photomagnetic effects.11,36,37 Furthermore, it has (not sure about “been”)recently suggested that the high-spin d7 configuration of Co2+-based compounds may host a dominant Kitaev interaction.40–42 However, despite intense studies of the XPS3 family, the ultrafast optical control of magnetism and lattice in CoPS3 remains unexplored.

Here, we address these shortcomings by first introducing optical magnetic linear dichroism (MLD) as an efficient means to probe the AFM order in CoPS3 and then employing time-resolved magneto-optical pump–probe spectroscopy to detect the ultrafast light-induced dynamics of spins and lattice in this AFM compound. Comparison of the results with earlier studies for MnPS3 and FePS3 reveals that the presence of the large unquenched orbital momentum in CoPS3 results in a substantially higher MLD and much faster laser-induced melting of the spin order. Moreover, the unquenched momentum leads to exceptionally strong spin–lattice coupling and provides an opportunity for highly efficient optical control of the lattice dynamics by the selective pumping of specific orbital transitions in magnetic Co2+ ions.

CoPS3 is a layered van der Waals AFM, characterized by a weak coupling between the adjacent crystal and magnetic layers, which can be viewed approximately as a quasi-2D antiferromagnet.43 The intralayer spin ordering below TN = 120 K results in the formation of ferromagnetic “zigzag” chains along the a-axis, while the adjacent chains in the direction of the b-axis are AFM coupled (see Fig. 1(a). Introducing individual magnetizations of the adjacent chains as M1 and M2, the AFM order in CoPS3 can be naturally characterized by a Néel vector L, such that L = M2M1. Inelastic neutron scattering shows that CoPS3 has sizable easy-axis single-ion anisotropy, which defines the orientation of the spins and the resultant Néel vector along the crystallographic a-axis.40 The formation of spin chains, combined with a strong spin–lattice interaction of Co2+ ions, leads to a reduction of the crystal symmetry, such that the point group of a single CoPS3 layer changes from D3d to C2h.16 Above TN in the paramagnetic (PM) phase, the crystal lattice is characterized by a six-fold rotational symmetry. Below TN, the ordering of spins in chains leads to a compression of a and an elongation of b lattice parameters, effectively elongating the hexagons formed by the Co2+ ions in the direction perpendicular to the spin chains. The structural changes that accompany the magnetic transition strongly affect the vibrational (phonon) spectrum of CoPS3. Recent Raman studies show that upon spin ordering, several otherwise double degenerate Eg phonon modes lose their degeneracy and split into a pair of nondegenerate Ag and Bg ones, of which the frequencies demonstrate an anomalous behavior below TN.16 

FIG. 1.

(a) Schematics of the AFM spin and crystal structure of CoPS3 in the antiferromagnetic and paramagnetic phases. (b), (c) The normalized transmission of light as a function of linear polarization angle ϑ measured below and above TN, respectively. The angle ϑ = 0 corresponds to the orientation of the polarization plane along the a-axis. (d) Magnetic linear dichroism (αMLD) of CoPS3 measured as a function of temperature. The solid line is a guide to the eye. (e) Rotation of the light polarization plane θ as a function of temperature. The solid line is a guide to the eye. The photon energy of the probe light is 1.55 eV.

FIG. 1.

(a) Schematics of the AFM spin and crystal structure of CoPS3 in the antiferromagnetic and paramagnetic phases. (b), (c) The normalized transmission of light as a function of linear polarization angle ϑ measured below and above TN, respectively. The angle ϑ = 0 corresponds to the orientation of the polarization plane along the a-axis. (d) Magnetic linear dichroism (αMLD) of CoPS3 measured as a function of temperature. The solid line is a guide to the eye. (e) Rotation of the light polarization plane θ as a function of temperature. The solid line is a guide to the eye. The photon energy of the probe light is 1.55 eV.

Close modal
To be able to employ an all-optical pump-and-probe technique for studies of ultrafast magnetism in CoPS3, one must find an efficient method of the optical detection of spin order in the AFM phase. Earlier, it was reported that several XPS3 materials possess a large magnetic linear dichroism (MLD) in the AFM phase.13,44 Moreover, according to Refs. 13, the strength of the dichroism seems to scale with the orbital moment of the magnetic ion. The MLD is the smallest in MnPS3, where the orbital momentum of the magnetic Mn2+ ions is quenched, and the largest in FePS3, where the orbital momentum of Fe2+ is known to be non-zero. To reveal the strength of this magneto-optical effect in CoPS3, with an even larger orbital momentum associated with the Co2+ ions, we measured the transmittance of linearly polarized probe light at a photon energy of 1.55 eV as a function of the angle ϑ between the electric field of light and the a-axis of the crystal. Figures 1(b) and 1(c) show the polarization dependencies below (T = 7 K) and above (T = 300 K) the TN. It is seen that the normalized transmittance is strongly anisotropic if the material is in the AFM state. Measuring the difference between the intensities Ia and Ib of the transmitted light polarized along the a- and the b-axes, respectively, we define MLD as
(1)

Although MLD, in general, strongly depends on the probe photon energy, already in our experiment its value reaches 30%. This is almost an order of magnitude larger than the largest MLD reported for FePS3.13 To define the origin of the MLD in CoPS3, we studied the effect as a function of the probe photon energy. As shown in the supplementary material Fig. S1, two well-pronounced MLD bands appear with the maxima centered at the photon energies of about 1.69 and 2.34 eV. These energies closely match the energies of the so-called d–d transitions of Co2+ ions that are responsible for the change of the orbital state of the magnetic ion.45 This observation, thus, indicates that the d–d transitions are the origin of MLD in CoPS3, in agreement with previous studies.13,19,20,35

Figure 1(d) shows MLD in CoPS3 as a function of temperature T. It is seen that αMLD reduces significantly when approaching TN from below. To quantify the critical behavior of αMLD, we fitted the temperature dependence of αMLD in the AFM phase using a power law αMLD∝|TNT|2β. The results, shown in Fig. 1(d), yield 2β = 0.60 ± 0.01. This value agrees well with the critical exponent characterizing a temperature scaling of the decay in the intensity of the neutron Bragg diffraction peak from the underlying AFM order in CoPS3 which was found to be 2β = 0.60.43 As the Bragg intensity scales with the Néel order as L2, its comparison with our results suggests that αMLD also scales as L2. In theory, MLD should be quadratic concerning L, but experimentally, this is not always the case, especially in AFMs with strong piezomagnetism.46 At the same time, MLD is known to be one of the most universal effects to probe the AFM order using light in a broad spectral range, including THz,47 visible,48,49 and x-rays.50 Yielding β ≈ 0.30 for the critical exponent characterizing the temperature dependence of the AFM order parameter, our results further confirm that the spin order in CoPS3 is best described by the 3D Ising model.51 Note, the dichroism, as defined in this work, does not depend on the sample thickness and, thus, can be potentially employed to probe AFM spin order in CoPS3 down to a single layer. At the same time, we observed that in the finite size CoPS3 samples studied in our experiments, the linear dichroism does not completely vanish above TN; see Fig. 1(d). This is because the ideal six-fold rotational symmetry of the honeycomb lattice, present in a single-layer form, vanishes in bulk crystals due to the displacement of the stacked layers along the a-axis.31 

In our experiments, to probe the anisotropy of the optical properties induced by the spin order, we employ the fact that if the polarization of the incoming light does not orient along and/or perpendicularly to the Néel vector (the a- and b-axes, respectively), the MLD will result in rotation of the polarization plane for light propagating through the sample. In our case, the light, initially polarized at 45° with respect to both the a- and b-axes, upon propagation through a 4 µm thick sample experiences a net polarization rotation for an angle reaching about 13° [Fig. 1(e)]. We note that, although the value of the angle is large, it is by far less than expected for an MLD as large as 30%. This difference can be explained by the fact that according to the Kramers–Kronig relations,52,53 linear dichroism is always accompanied by linear birefringence. The latter can modify the polarization state of light from linear to elliptical and even circular and, thus, substantially hampers the measurements of the polarization rotation.54 

To study the light-induced ultrafast dynamics of spins and lattice in CoPS3, we carried out an all-optical pump–probe experiment; see Fig. 2(a). To excite the dynamics, we employed ultrashort (∼50 fs) linearly polarized pulses of light with the photon energy hv tunable in a range from 0.7 to 2.5 eV. This energy range covers the vast majority of the orbital d–d transitions in Co2+ ions.45 To probe the light-driven spin dynamics, we relied on measuring the transient polarization rotation Δθ providing, as we have already shown, access to the Néel order. The polarization rotation was also employed to detect the lattice vibrations (phonons), of which the dynamics are intrinsically highly anisotropic and, thus, contribute to the linear birefringence.18,55 Figure 2(b) shows an example of the light-induced dynamics triggered by the pump pulse at a temperature of T = 77 K when the sample is in the AFM phase. It is seen that, after the excitation, the signal of the transient rotation Δθ suddenly drops, indicating a suppression of the AFM order. Note, that the suppression does not occur instantaneously within the duration of the pump pulse but rather proceeds on a longer timescale τs of about 1.5 ps [Fig. 2(b)]. The quenching of the AFM order is concomitant with a set of coherent high-frequency phonon oscillations featuring frequencies from 3 to 10 THz and dominated by a phonon mode at 4.74 THz; see the inset in Fig. 2(b) and the supplementary material in Fig. S2.

FIG. 2.

(a) Schematics of the time-resolved magneto-optical pump–probe experiment in CoPS3. The pump pulse (blue) is delayed with respect to the probe pulse by a variable time delay Δt. (b) Time-resolved rotation of the probe pulse polarization plane exemplifying pump-induced quenching of the AFM order. The inset shows zoom-in of the background-free dynamics of the coherent Bg phonon mode. The sample temperature is set to 77 K.

FIG. 2.

(a) Schematics of the time-resolved magneto-optical pump–probe experiment in CoPS3. The pump pulse (blue) is delayed with respect to the probe pulse by a variable time delay Δt. (b) Time-resolved rotation of the probe pulse polarization plane exemplifying pump-induced quenching of the AFM order. The inset shows zoom-in of the background-free dynamics of the coherent Bg phonon mode. The sample temperature is set to 77 K.

Close modal

Figure 3(a) shows the dynamics of the pump-induced polarization rotation Δθ, demonstrating the evolution of the light-induced quenching as a function of the temperature T. We fit the time-resolved traces to a single exponential function from which we extract amplitude and quenching time; see the supplementary material S3. It is seen [Fig. 3(b)] that upon approaching TN the amplitude of the quenching is gradually growing and peaks right below TN. Above TN, no significant light-induced magnetic dynamics is seen; the small residual signal is likely of non-magnetic origin and caused by pump-induced changes to the electronic part of the dielectric function.56 The growth of the quenching amplitude is accompanied by the growth of the quenching time τs. Upon approaching TN [Fig. 3(c)], this time increases from 1.6 ps at T = 10 K to a maximum detected time of 5 ps at about T = 118 K. We note that qualitatively similar behavior was also observed when the probing was performed in the reflection geometry; see the supplementary material Fig. S4.

FIG. 3.

Ultrafast light-induced quenching of the Néel order in CoPS3 (a) Time-resolved rotation of the probe polarization Δθ showing pump-induced quenching of the AFM order for various temperatures across TN. The pump photon energy is 0.9 eV. (b) Amplitude of the quenching as a function of temperature T. (c) Quenching time as a function of T.

FIG. 3.

Ultrafast light-induced quenching of the Néel order in CoPS3 (a) Time-resolved rotation of the probe polarization Δθ showing pump-induced quenching of the AFM order for various temperatures across TN. The pump photon energy is 0.9 eV. (b) Amplitude of the quenching as a function of temperature T. (c) Quenching time as a function of T.

Close modal

To establish the origin of the AFM quenching, we varied the photon energy of the pump pulse in the range of several d–d transitions of Co2+ ions. We find that the amplitude of the quenching scales with the absorption does not depend on the origin of the optical transition. The quenching is defined by the amount of heat deposited into CoPS3 by a laser pulse. The origin of the laser-induced quenching of AFM order is, thus, similar to that reported earlier for other materials from the XPS3 family.18 Although laser-induced demagnetization is often described using a three-temperature 3T-model,1 CoPS3 lacks free electrons and the model is not adequate in this case. Instead, the quenching of the AFM order in CoPS3 can be described by a two-temperature 2T model, where the pump photons increase the potential energy of the excited electrons without making them hot. Upon non-radiative recombination of the excited electrons, the latter promptly (<1 ps) transfer the gained photon energy to the lattice.57 The subsequent heat exchange via spin–lattice interaction leads to an increase in the effective spin temperature and consequently to the melting of the spin order.58 This mechanism can explain a substantial increase in the magnitude of the quenching as the temperature approaches TN. Indeed, at higher temperatures, the derivative of the Néel order parameter L with respect to the temperature increases, and the spins become more susceptible to temperature variations; see Fig. 1(d). Remarkably, not only the amplitude of the quenching goes up, but the quenching time τs also experiences a substantial increase in the vicinity of TN. Such a critical slowing down of the transient spin dynamics induced by light is seen in many AFM compounds, including in recently published results on MnPS3 and FePS3.19,35 It has been shown18 that in magnetic insulators the rate of the spin–lattice coupling defining τs scales with the spin-specific heat Cs, such that τs = Cs/gsl, where gsl is the spin–lattice relaxation rate representing the strength of the spin–lattice interaction. As the heat capacity near the Néel temperature is expected to diverge, assuming a weak temperature dependence of gsl near TN, the characteristic time τs is expected to follow the divergence of Cs. Indeed, our experiment shows that τs ∝ |TNT|−α with a critical exponent α ∼ 0.1. Remarkably, this value closely matches the theoretical one of α = 0.1 characterizing the critical scaling of the heat capacity in the 3D Ising model.51,59

We would like to emphasize that the characteristic time for the quenching of the AFM order in CoPS3 (τs = 2 ps) is much shorter than the one reported for other XPS3 compounds and MnPS3, in particular (τs = 30 ps).19 It is well known that the heat capacity Cs scales with the strength of the exchange interaction, which is proportional to the Néel temperature. As CoPS3 and MnPS3 are characterized by similar values of TN, the difference in the quenching time can only be explained by a difference in the magnitude of the spin–lattice relaxation rate gsl. Using the available data for the specific heat capacity,16,60 we estimate gsl to be about 1 × 1014 Wm−3 K−1 and 5 × 1011 Wm−3 K−1 for CoPS3 and MnPS3, respectively. Our experiment, thus, shows that the strength of the coupling can be effectively changed by more than two orders of magnitude by inducing an orbital momentum at the ground state of the magnetic ion. We note that, although the obtained values of the spin–lattice relaxation rates in CoPS3 due to the unquenched momentum of the Co2+ ions are high, this is only in comparison with electronically similar materials lacking mobile electrons (dielectrics and semiconductors). In metals, free electrons serve as an additional reservoir of energy that can increase the rate of laser-induced demagnetization by another two orders of magnitude.61 

To further understand the nature and excitation mechanism of the light-driven coherent phonon mode that dominates the Δθ signal, we studied its dynamics as a function of both temperature and photon energy. Figure 4(a) shows the oscillations measured at various T. It is seen that the excitation is the most efficient below TN, where both the amplitude and the frequency of the oscillations are strongly dependent on T. Using the Fourier transform, we retrieved the frequency and amplitude of the oscillations and plotted them as functions of T in Figs. 4(b) and 4(c), respectively. As the temperature increases, the phonon frequency f softens until it reaches TN, where it stabilizes at f0 = 4.64 THz. Fitting the temperature evolution of the frequency shift Δf = ff0 to a critical law similar to the one used for αMLD, we find that their critical exponents closely match each other, thus indicating that Δf also follows L2. In accordance with Ref. 16, we assign the oscillations to the Bg phonon mode that involves antiphase motions of the magnetic Co2+ ions in the direction parallel to the orientation of their spins [Fig. 5(b)]. The comparison of the temperature behavior of the amplitude of the phonon mode in AFM and PM phases shows that, in PM (T > TN), there is no significant temperature dependence, but in AFM (TTN), the amplitude rises linearly as temperature decreases. The rise, concomitant with the onset of the AFM order, clearly indicates that the establishment of the Néel order facilitates the excitation of the lattice dynamics.

FIG. 4.

High frequency coherent spin-coupled THz phonon mode in CoPS3. (a) Time-resolved rotation of the probe polarization showing dynamics of the pump-excited coherent phonon plotted for various temperatures across TN. (b) Frequency f of the phonon mode as a function of temperature T. (c) Amplitude of the phonon mode as a function of temperature T. The solid lines are independent linear fits below and above TN.

FIG. 4.

High frequency coherent spin-coupled THz phonon mode in CoPS3. (a) Time-resolved rotation of the probe polarization showing dynamics of the pump-excited coherent phonon plotted for various temperatures across TN. (b) Frequency f of the phonon mode as a function of temperature T. (c) Amplitude of the phonon mode as a function of temperature T. The solid lines are independent linear fits below and above TN.

Close modal
FIG. 5.

(a) The amplitude of the Bg phonon mode as a function of the pump photon energy (left axis). Optical absorption of CoPS3 in the near IR to the visible range: the absorption line at around 0.9 eV corresponds to the 4T1g4T2g orbital transition (right axis). The data are taken from Ref. 44. The samples temperature is 77 K. Inset: Amplitude of the phonon as a function of the pump polarization angle γ with respect to the a-crystal axis. The signs “+” and “−” indicate the relative phase of the oscillations and correspond to 0° and 180°, respectively. (b) Schematics of the orbital excitation of the Bg phonon, where the blue and red ions are Co2+ ions with antiparallel spins.

FIG. 5.

(a) The amplitude of the Bg phonon mode as a function of the pump photon energy (left axis). Optical absorption of CoPS3 in the near IR to the visible range: the absorption line at around 0.9 eV corresponds to the 4T1g4T2g orbital transition (right axis). The data are taken from Ref. 44. The samples temperature is 77 K. Inset: Amplitude of the phonon as a function of the pump polarization angle γ with respect to the a-crystal axis. The signs “+” and “−” indicate the relative phase of the oscillations and correspond to 0° and 180°, respectively. (b) Schematics of the orbital excitation of the Bg phonon, where the blue and red ions are Co2+ ions with antiparallel spins.

Close modal

Figure 5(a) shows the amplitude of the Bg phonon mode as a function of the pump photon energy hv. Unlike the light-induced spin quenching, the laser-induced lattice dynamics does not correlate with the absorption coefficient but, instead, is most efficient if the photon energy is in resonance with the dd transition 4T1g4T2g in the Co2+ ions at 0.87 eV.45 We have also found that the efficiency of the excitation is strongly dependent on the incoming pump polarization. The inset in Fig. 5(a) shows how the amplitude of the phonon oscillations depends on the angle γ the pump polarization plane forms with the a-axis along which the spins chains are formed. It is seen that the excitation is most efficient if the pump is polarized at 45° with respect to the spins, while no oscillations are excited if the polarization is oriented either along or perpendicularly.

To explain the unusual temperature behavior of the Bg phonon mode and its excitation mechanism, we first refer to Ref. 62, which demonstrates that in FePS3, also characterized by unquenched angular momentum, the spin–lattice interaction leads to the hybridization of spin and lattice dynamics accompanied by renormalization of their frequencies. The energy of such coupling to the lowest order acquires the following form:
(2)
where U is a phenomenological coupling parameter, Q is the normal coordinate of the Bg mode, and l L is a component of the Néel vector along the b-axis, describing its deviation from the ground state L = Lâ. It is, thus, clear that to enable the spin–lattice coupling in CoPS3, the spins have to be brought out-of-equilibrium. Next, we consider the energy of the light–matter interaction enabling excitation of both the Néel vector l and the Bg phonon mode. The energy describing this interaction reads
(3)
where Ea,b are time-dependent electric field components of the pump pulse and α and β are phenomenological parameters that describe the strength of the light–matter coupling in CoPS3. It is seen that the pump can be coupled not only to the phonon mode but also to the Néel vector. Remarkably, the efficiency of this coupling is proportional to EaEb ∝ sin 2γ and, thus, perfectly agrees with the pump polarization dependence of the phonon amplitude; see the inset in Fig. 5(a). In the supplementary material we demonstrate that a sufficiently short pump pulse can impulsively excite coherent dynamics of both Q and l in accordance with a mechanism known as impulsive stimulated Raman scattering (ISRS).63,64 We also show that the inclusion of the spin–lattice coupling hybridizes their dynamics and may explain the enhancement of the phonon amplitude in the AFM phase. Indeed, the hybridization renormalizes the phonon coordinate Q̃=Q+ζl. The parameter ζ characterizes the degree of the hybridization and is proportional to the product UL. Therefore, it may naturally explain the enhancement of the phonon excitation in the AFM phase, where L ≠ 0. Moreover, the hybridization causes a renormalization of the phonon frequency Δf in the AFM phase,
(4)
which once again lines up with our experimental findings. We note that although the suggested theory explains experimental findings remarkably well, it still relies on the excitation of the coherent spin dynamics, which is not observed in our experiments. This is rather surprising, as the excitation and magneto-optical detection of the coherent spin dynamics with femtosecond optical pulses have been reported recently in both MnPS3 and NiPS3.19,20 The analogy with the latter is particularly striking as NiPS3 is magnetically isomorphous to the studied here CoPS3. Future probes of the laser-induced spin dynamics in CoPS3, e.g., time-resolved THz spectroscopy, may help to elucidate this issue.65 In accordance with these theoretical considerations, the resonance dependence of the amplitude of the phonon mode on the pump photon energy, shown in Fig. 5(a), implies that either both or one of the phenomenological parameters U and β is maximized at this photon energy. This implies that, among all possible dd electronic transitions in the range from 0.5 to 2.2 eV, the excitation of the Bg mode is most affected by the 4T1g to 4T2g transition, which can be seen as a change of the effective orbital momentum of the magnetic Co2+ ions; see Fig. 5(b). The corresponding changes to the interionic potential, mediated by the strong spin–lattice coupling, are likely due to the mechanical force that moves the ions in the direction parallel to the orientation of their spins.

To conclude, we have shown that in the vdW AFM CoPS3, characterized by a strong spin–lattice coupling, the AFM order can be effectively probed with MLD. Using MLD, we detected the ultrafast dynamics of spins and lattice induced by ultrashort pulses of light. We showed that light can suddenly heat the magnetic system, leading to a substantial loss (∼1%) of the spin ordering within nearly a single picosecond. Resonantly pumping dd transitions in magnetic Co2+ ions, we effectively change the orbital momentum of Co2+ ions and show that this excitation, mediated by the spin–lattice coupling, brings Co2+ ions in a coherent motion in the direction of the AFM Néel vector. Our experiments not only elucidate the nature of the ultrafast spin–lattice coupling in 2D vdW AFMs but also lay the ground for future ultrafast pump–probe experiments, particularly those aimed at resonant pumping of infrared-active structural phonon modes.7,66

The supplementary material includes five individual sections. Section I shows the spectral dependence of the MLD in CoPS3. Section II demonstrates the time-resolved traces of the pump-excited coherent phonon modes and their corresponding Fast Fourier Transform (FFT) spectra. Section III describes the fitting procedure for the time-resolved dynamics of the pump-induced quenching of the AFM order. Section IV shows the time-resolved traces of the pump-induced polarization rotation reflection in the reflection geometry. Section V demonstrates the temperature dependence of the spin–lattice coupling constant gsl. The phenomenological theory of the spin–phonon coupling mechanism is described in Sec. VI.

The authors are grateful to M. Matthiessen, J. R. Hortensius, A. D. Caviglia and M. I. Katsnelson for their fruitful discussions and to S. Semin, K. Saeedi, and C. Berkhout for their technical support. This work was funded by the Netherlands Organization for Scientific Research (NWO), the European Union Horizon 2020 and innovation program under the European Research Council ERC Grant Agreement No. 856538 (3D-MAGiC), the European Union Horizon 2020 innovation program under the Marie Skłodowska-Curie Grant Agreement No. 861300 (COMRAD), the National Research Fund of Ukraine within Project No. 2020.02/026, the Gravitation program of the Dutch Ministry of Education, Culture and Science (OCW) under the research program “Materials for the Quantum Age” (QuMat) registration number 024.005.006, and the ERC (Grant Nos. 1010 78206, ASTRAL). S.M.V. acknowledges the Generalitat Valenciana for the postdoctoral fellow APOSTD-CIAPOS2021/215, and S.A. was supported by the Computational Chemical Sciences program within the Office of Basic Energy Sciences, U.S. Department of Energy under Contract No. DE-AC36-08GO28308.

The authors have no conflicts to disclose.

D. Khusyainov: Data curation (lead); Formal analysis (lead); Investigation (lead); Methodology (lead); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). T. Gareev: Data curation (supporting); Formal analysis (supporting); Investigation (supporting); Methodology (supporting); Visualization (supporting). V. Radovskaia: Data curation (supporting); Formal analysis (supporting); Investigation (supporting); Methodology (supporting); Visualization (supporting). K. Sampathkumar: Data curation (supporting); Formal analysis (supporting); Investigation (supporting); Methodology (supporting); Writing – original draft (supporting); Writing – review & editing (supporting). S. Acharya: Conceptualization (supporting); Formal analysis (supporting); Writing – review & editing (supporting). M. Šiškins: Conceptualization (supporting); Resources (equal); Writing – review & editing (supporting). S. Mañas-Valero: Resources (equal); Writing – review & editing (equal). B. A. Ivanov: Conceptualization (equal); Investigation (equal); Methodology (equal); Writing – review & editing (equal). E. Coronado: Resources (equal); Supervision (equal). Th. Rasing: Funding acquisition (equal); Project administration (equal); Supervision (equal); Writing – original draft (equal); Writing – review & editing (equal). A. V. Kimel: Conceptualization (equal); Formal analysis (equal); Funding acquisition (equal); Project administration (equal); Supervision (equal); Validation (equal); Writing – review & editing (equal). D. Afanasiev: Conceptualization (lead); Data curation (equal); Formal analysis (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Project administration (lead); Resources (equal); Supervision (lead); Validation (equal); Visualization (equal); Writing – original draft (lead); Writing – review & editing (lead).

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

1.
E.
Beaurepaire
,
J.-C.
Merle
,
A.
Daunois
, and
J.-Y.
Bigot
,
Phys. Rev. Lett.
76
,
4250
(
1996
).
2.
A. V.
Kimel
,
A.
Kirilyuk
,
P. A.
Usachev
,
R. V.
Pisarev
,
A. M.
Balbashov
, and
T.
Rasing
,
Nature
435
,
655
(
2005
).
3.
D.
Bossini
and
T.
Rasing
,
Phys. Scr.
92
,
024002
(
2017
).
4.
A. M.
Kalashnikova
,
A. V.
Kimel
,
R. V.
Pisarev
,
V. N.
Gridnev
,
A.
Kirilyuk
, and
T.
Rasing
,
Phys. Rev. Lett.
99
,
167205
(
2007
).
5.
S.
Baierl
,
M.
Hohenleutner
,
T.
Kampfrath
,
A. K.
Zvezdin
,
A. V.
Kimel
,
R.
Huber
, and
R. V.
Mikhaylovskiy
,
Nat. Photonics
10
,
715
(
2016
).
6.
J. R.
Hortensius
,
D.
Afanasiev
,
M.
Matthiesen
,
R.
Leenders
,
R.
Citro
,
A. V.
Kimel
,
R. V.
Mikhaylovskiy
,
B. A.
Ivanov
, and
A. D.
Caviglia
,
Nat. Phys.
17
,
1001
(
2021
).
7.
D.
Afanasiev
,
J. R.
Hortensius
,
B. A.
Ivanov
,
A.
Sasani
,
E.
Bousquet
,
Y. M.
Blanter
,
R. V.
Mikhaylovskiy
,
A. V.
Kimel
, and
A. D.
Caviglia
,
Nat. Mater.
20
,
607
(
2021
).
8.
A. V.
Kimel
,
A.
Kirilyuk
,
A.
Tsvetkov
,
R. V.
Pisarev
, and
T.
Rasing
,
Nature
429
,
850
(
2004
).
9.
P.
Beaud
,
A.
Caviezel
,
S. O.
Mariager
,
L.
Rettig
,
G.
Ingold
,
C.
Dornes
,
S.-W.
Huang
,
J. A.
Johnson
,
M.
Radovic
,
T.
Huber
,
T.
Kubacka
,
A.
Ferrer
,
H. T.
Lemke
,
M.
Chollet
,
D.
Zhu
,
J. M.
Glownia
,
M.
Sikorski
,
A.
Robert
,
H.
Wadati
,
M.
Nakamura
,
M.
Kawasaki
,
Y.
Tokura
,
S. L.
Johnson
, and
U.
Staub
,
Nat. Mater.
13
,
923
(
2014
).
10.
C. D.
Stanciu
,
F.
Hansteen
,
A. V.
Kimel
,
A.
Kirilyuk
,
A.
Tsukamoto
,
A.
Itoh
, and
T.
Rasing
,
Phys. Rev. Lett.
99
,
047601
(
2007
).
11.
A.
Stupakiewicz
,
K.
Szerenos
,
D.
Afanasiev
,
A.
Kirilyuk
, and
A. V.
Kimel
,
Nature
542
,
71
(
2017
).
12.
A. V.
Kimel
,
A. M.
Kalashnikova
,
A.
Pogrebna
, and
A. K.
Zvezdin
,
Phys. Rep.
852
,
1
(
2020
).
13.
Q.
Zhang
,
K.
Hwangbo
,
C.
Wang
,
Q.
Jiang
,
J.-H.
Chu
,
H.
Wen
,
D.
Xiao
, and
X.
Xu
,
Nano Lett.
21
,
6938
(
2021
).
14.
B. H.
Zhang
,
Y. S.
Hou
,
Z.
Wang
, and
R. Q.
Wu
,
Phys. Rev. B
100
,
224427
(
2019
).
15.
A.
Ghosh
,
M.
Palit
,
S.
Maity
,
V.
Dwij
,
S.
Rana
, and
S.
Datta
,
Phys. Rev. B
103
,
064431
(
2021
).
16.
Q.
Liu
,
L.
Wang
,
Y.
Fu
,
X.
Zhang
,
L.
Huang
,
H.
Su
,
J.
Lin
,
X.
Chen
,
D.
Yu
,
X.
Cui
,
J.-W.
Mei
, and
J.-F.
Dai
,
Phys. Rev. B
103
,
235411
(
2021
).
17.
Y.-J.
Sun
,
J.-M.
Lai
,
S.-M.
Pang
,
X.-L.
Liu
,
P.-H.
Tan
, and
J.
Zhang
,
J. Phys. Chem. Lett.
13
,
1533
(
2022
).
18.
X.-X.
Zhang
,
S.
Jiang
,
J.
Lee
,
C.
Lee
,
K. F.
Mak
, and
J.
Shan
,
Nano Lett.
21
,
5045
(
2021
).
19.
M.
Matthiesen
,
J. R.
Hortensius
,
S.
Mañas-Valero
,
I.
Kapon
,
D.
Dumcenco
,
E.
Giannini
,
M.
Šiškins
,
B. A.
Ivanov
,
H. S. J.
van der Zant
,
E.
Coronado
,
A. B.
Kuzmenko
,
D.
Afanasiev
, and
A. D.
Caviglia
,
Phys. Rev. Lett.
130
,
076702
(
2023
).
20.
D.
Afanasiev
,
J. R.
Hortensius
,
M.
Matthiesen
,
S.
Mañas-Valero
,
M.
Šiškins
,
M.
Lee
,
E.
Lesne
,
H. S. J.
van der Zant
,
P. G.
Steeneken
,
B. A.
Ivanov
,
E.
Coronado
, and
A. D.
Caviglia
,
Sci. Adv.
7
,
eabf3096
(
2021
).
21.
P.
Zhang
,
T.-F.
Chung
,
Q.
Li
,
S.
Wang
,
Q.
Wang
,
W. L. B.
Huey
,
S.
Yang
,
J. E.
Goldberger
,
J.
Yao
, and
X.
Zhang
,
Nat. Mater.
21
,
1373
(
2022
).
22.
S.
Li
,
L.
Zhou
,
T.
Frauenheim
, and
J.
He
,
J. Phys. Chem. Lett.
13
,
6223
(
2022
).
23.
H.
Ling
and
A. R.
Davoyan
,
Nat. Photonics
16
,
259
(
2022
).
24.
B.
Liu
,
S.
Liu
,
L.
Yang
,
Z.
Chen
,
E.
Zhang
,
Z.
Li
,
J.
Wu
,
X.
Ruan
,
F.
Xiu
,
W.
Liu
,
L.
He
,
R.
Zhang
, and
Y.
Xu
,
Phys. Rev. Lett.
125
,
267205
(
2020
).
25.
A.
Barman
,
G.
Gubbiotti
,
S.
Ladak
,
A. O.
Adeyeye
,
M.
Krawczyk
,
J.
Gräfe
,
C.
Adelmann
,
S.
Cotofana
,
A.
Naeemi
,
V. I.
Vasyuchka
,
B.
Hillebrands
,
S. A.
Nikitov
,
H.
Yu
,
D.
Grundler
,
A. V.
Sadovnikov
,
A. A.
Grachev
,
S. E.
Sheshukova
,
J.-Y.
Duquesne
,
M.
Marangolo
,
G.
Csaba
,
W.
Porod
,
V. E.
Demidov
,
S.
Urazhdin
,
S. O.
Demokritov
,
E.
Albisetti
,
D.
Petti
,
R.
Bertacco
,
H.
Schultheiss
,
V. V.
Kruglyak
,
V. D.
Poimanov
,
S.
Sahoo
,
J.
Sinha
,
H.
Yang
,
M.
Münzenberg
,
T.
Moriyama
,
S.
Mizukami
,
P.
Landeros
,
R. A.
Gallardo
,
G.
Carlotti
,
J.-V.
Kim
,
R. L.
Stamps
,
R. E.
Camley
,
B.
Rana
,
Y.
Otani
,
W.
Yu
,
T.
Yu
,
G. E. W.
Bauer
,
C.
Back
,
G. S.
Uhrig
,
O. V.
Dobrovolskiy
,
B.
Budinska
,
H.
Qin
,
S.
van Dijken
,
A. V.
Chumak
,
A.
Khitun
,
D. E.
Nikonov
,
I. A.
Young
,
B. W.
Zingsem
, and
M.
Winklhofer
,
J. Phys.: Condens. Matter
33
,
413001
(
2021
).
26.
A.
Hirohata
,
H.
Sukegawa
,
H.
Yanagihara
,
I.
Zutic
,
T.
Seki
,
S.
Mizukami
, and
R.
Swaminathan
,
IEEE Trans. Magn.
51
,
1
(
2015
).
27.
S. Y.
Kim
,
T. Y.
Kim
,
L. J.
Sandilands
,
S.
Sinn
,
M.-C.
Lee
,
J.
Son
,
S.
Lee
,
K.-Y.
Choi
,
W.
Kim
,
B.-G.
Park
,
C.
Jeon
,
H.-D.
Kim
,
C.-H.
Park
,
J.-G.
Park
,
S. J.
Moon
, and
T. W.
Noh
,
Phys. Rev. Lett.
120
,
136402
(
2018
).
28.
R.
Galceran
,
B.
Tian
,
J.
Li
,
F.
Bonell
,
M.
Jamet
,
C.
Vergnaud
,
A.
Marty
,
J. H.
García
,
J. F.
Sierra
,
M. V.
Costache
,
S.
Roche
,
S. O.
Valenzuela
,
A.
Manchon
,
X.
Zhang
, and
U.
Schwingenschlögl
,
APL Mater.
9
,
100901
(
2021
).
29.
K.
Hwangbo
,
Q.
Zhang
,
Q.
Jiang
,
Y.
Wang
,
J.
Fonseca
,
C.
Wang
,
G. M.
Diederich
,
D. R.
Gamelin
,
D.
Xiao
,
J.-H.
Chu
,
W.
Yao
, and
X.
Xu
,
Nat. Nanotechnol.
16
,
655
(
2021
).
30.
E.
Ressouche
,
M.
Loire
,
V.
Simonet
,
R.
Ballou
,
A.
Stunault
, and
A.
Wildes
,
Phys. Rev. B
82
,
100408
(
2010
).
31.
H.
Chu
,
C. J.
Roh
,
J. O.
Island
,
C.
Li
,
S.
Lee
,
J.
Chen
,
J.-G.
Park
,
A. F.
Young
,
J. S.
Lee
, and
D.
Hsieh
,
Phys. Rev. Lett.
124
,
027601
(
2020
).
32.
U. F. P.
Seifert
,
M.
Ye
, and
L.
Balents
,
Phys. Rev. B
105
,
155138
(
2022
).
33.
F.
Wang
,
N.
Mathur
,
A. N.
Janes
,
H.
Sheng
,
P.
He
,
X.
Zheng
,
P.
Yu
,
A. J.
DeRuiter
,
J. R.
Schmidt
,
J.
He
, and
S.
Jin
,
Sci. Adv.
7
,
eabj4086
(
2021
).
34.
E.
Ergeçen
,
B.
Ilyas
,
D.
Mao
,
H. C.
Po
,
M. B.
Yilmaz
,
J.
Kim
,
J.-G.
Park
,
T.
Senthil
, and
N.
Gedik
,
Nat. Commun.
13
,
98
(
2022
).
35.
F.
Mertens
,
D.
Mönkebüscher
,
U.
Parlak
,
C.
Boix‐Constant
,
S.
Mañas‐Valero
,
M.
Matzer
,
R.
Adhikari
,
A.
Bonanni
,
E.
Coronado
,
A. M.
Kalashnikova
,
D.
Bossini
, and
M.
Cinchetti
,
Adv. Mater.
35
,
2208355
(
2023
).
36.
E. A.
Mashkovich
,
K. A.
Grishunin
,
R. M.
Dubrovin
,
A. K.
Zvezdin
,
R. V.
Pisarev
, and
A. V.
Kimel
,
Science
374
,
1608
(
2021
).
37.
T.
Satoh
,
R.
Iida
,
T.
Higuchi
,
Y.
Fujii
,
A.
Koreeda
,
H.
Ueda
,
T.
Shimura
,
K.
Kuroda
,
V. I.
Butrim
, and
B. A.
Ivanov
,
Nat. Commun.
8
,
638
(
2017
).
38.
Y.
Suzuki
,
G.
Hu
,
R. B.
van Dover
, and
R. J.
Cava
,
J. Magn. Magn. Mater.
191
,
1
(
1999
).
39.
F.
Formisano
,
R. M.
Dubrovin
,
R. V.
Pisarev
,
A. M.
Kalashnikova
, and
A. V.
Kimel
,
J. Phys.: Condens. Matter
34
,
225801
(
2022
).
40.
C.
Kim
,
J.
Jeong
,
P.
Park
,
T.
Masuda
,
S.
Asai
,
S.
Itoh
,
H.-S.
Kim
,
A.
Wildes
, and
J.-G.
Park
,
Phys. Rev. B
102
,
184429
(
2020
).
41.
C.
Kim
,
H.-S.
Kim
, and
J.-G.
Park
,
J. Phys.: Condens. Matter
34
,
023001
(
2022
).
42.
G.
Jackeli
and
G.
Khaliullin
,
Phys. Rev. Lett.
102
,
017205
(
2009
).
43.
A. R.
Wildes
,
V.
Simonet
,
E.
Ressouche
,
R.
Ballou
, and
G. J.
McIntyre
,
J. Phys.: Condens. Matter
29
,
455801
(
2017
).
44.
H.
Zhang
,
Z.
Ni
,
C. E.
Stevens
,
A.
Bai
,
F.
Peiris
,
J. R.
Hendrickson
,
L.
Wu
, and
D.
Jariwala
,
Nat. Photonics
16
,
311
(
2022
).
45.
E. J. K. B.
Banda
,
Phys. Status Solidi
135
,
K43
(
1986
).
46.
A. S.
Borovik-Romanov
,
N. M.
Kreĭnes
,
A. A.
Pankov
, and
M. A.
Talalaev
,
Sov. Phys.-JETP
37
,
890
(
1973
).
47.
T. W. J.
Metzger
,
K. A.
Grishunin
,
D.
Afanasiev
,
R. M.
Dubrovin
,
E. A.
Mashkovich
,
R. V.
Pisarev
, and
A. V.
Kimel
,
Appl. Phys. Lett.
121
,
252403
(
2022
).
48.
G. A.
Smolenskiĭ
,
R.
V Pisarev
, and
I. G.
Siniĭ
,
Sov. Phys. Uspekhi
18
,
410
(
1975
).
49.
J.
Wagner
,
A.
Sahasrabudhe
,
R. B.
Versteeg
,
L.
Wysocki
,
Z.
Wang
,
V.
Tsurkan
,
A.
Loidl
,
D. I.
Khomskii
,
H.
Hedayat
, and
P. H. M.
van Loosdrecht
,
Npj Quantum Mater.
7
,
28
(
2022
).
50.
J.
Stöhr
and
S.
Hans Christoph
,
Magnetism
(
Springer Berlin Heidelberg
,
Berlin, Heidelberg
,
2006
).
51.
M. F.
Collins
,
Magnetic Critical Scattering
(
Oxford University Press Inc.
,
New York
,
1989
).
52.
N.
Tesařová
,
T.
Ostatnický
,
V.
Novák
,
K.
Olejník
,
J.
Šubrt
,
H.
Reichlová
,
C. T.
Ellis
,
A.
Mukherjee
,
J.
Lee
,
G. M.
Sipahi
,
J.
Sinova
,
J.
Hamrle
,
T.
Jungwirth
,
P.
Němec
,
J.
Černe
, and
K.
Výborný
,
Phys. Rev. B
89
,
085203
(
2014
).
53.
A. V.
Kimel
,
G. V.
Astakhov
,
A.
Kirilyuk
,
G. M.
Schott
,
G.
Karczewski
,
W.
Ossau
,
G.
Schmidt
,
L. W.
Molenkamp
, and
T.
Rasing
,
Phys. Rev. Lett.
94
,
227203
(
2005
).
54.
A. J.
Kurtzig
,
R.
Wolfe
,
R. C.
LeCraw
, and
J. W.
Nielsen
,
Appl. Phys. Lett.
14
,
350
(
1969
).
55.
J. R.
Hortensius
,
D.
Afanasiev
,
A.
Sasani
,
E.
Bousquet
, and
A. D.
Caviglia
,
Npj Quantum Mater.
5
,
95
(
2020
).
56.
I. R.
Jahn
,
Phys. Status Solidi
57
,
681
(
1973
).
57.
A.
Kirilyuk
,
A. V.
Kimel
, and
T.
Rasing
,
Rev. Mod. Phys.
82
,
2731
(
2010
).
58.
A. V.
Kimel
,
R. V.
Pisarev
,
J.
Hohlfeld
, and
T.
Rasing
,
Phys. Rev. Lett.
89
,
287401
(
2002
).
59.
R. K.
Pathria
and
P. D.
Beale
,
Statistical Mechanics
(
Elsevier
,
2011
).
60.
Y.
Takano
,
N.
Arai
,
A.
Arai
,
Y.
Takahashi
,
K.
Takase
, and
K.
Sekizawa
,
J. Magn. Magn. Mater.
272–276
,
E593
(
2004
).
61.
M.
Pankratova
,
I. P.
Miranda
,
D.
Thonig
,
M.
Pereiro
,
E.
Sjöqvist
,
A.
Delin
,
O.
Eriksson
, and
A.
Bergman
,
Phys. Rev. B
106
,
174407
(
2022
).
62.
S.
Liu
,
A.
Granados del Águila
,
D.
Bhowmick
,
C. K.
Gan
,
T.
Thu Ha Do
,
M. A.
Prosnikov
,
D.
Sedmidubský
,
Z.
Sofer
,
P. C. M.
Christianen
,
P.
Sengupta
, and
Q.
Xiong
,
Phys. Rev. Lett.
127
,
097401
(
2021
).
63.
64.
D. M.
Juraschek
and
S. F.
Maehrlein
,
Phys. Rev. B
97
,
174302
(
2018
).
65.
C. A.
Belvin
,
E.
Baldini
,
I. O.
Ozel
,
D.
Mao
,
H. C.
Po
,
C. J.
Allington
,
S.
Son
,
B. H.
Kim
,
J.
Kim
,
I.
Hwang
,
J. H.
Kim
,
J.-G.
Park
,
T.
Senthil
, and
N.
Gedik
,
Nat. Commun.
12
,
4837
(
2021
).
66.
A. S.
Disa
,
M.
Fechner
,
T. F.
Nova
,
B.
Liu
,
M.
Först
,
D.
Prabhakaran
,
P. G.
Radaelli
, and
A.
Cavalleri
,
Nat. Phys.
16
,
937
(
2020
).

Supplementary Material