You are on page 1of 6

Theory of Out-of-Equilibrium Ultrafast Relaxation Dynamics in Metals

Pablo Maldonado,1 Karel Carva,2, 1 Martina Flammer,1 and Peter M. Oppeneer1


1
Department of Physics and Astronomy, Uppsala University, P. O. Box 516, S-75120 Uppsala, Sweden
2
Charles University, Faculty of Mathematics and Physics,
Department of Condensed Matter Physics, Ke Karlovu 5, CZ-12116 Prague 2, Czech Republic
(Dated: August 7, 2017)
Ultrafast laser excitation of a metal causes correlated, highly nonequilibrium dynamics of electronic
and ionic degrees of freedom, which are however only poorly captured by the widely-used two-
temperature model. Here we develop an out-of-equilibrium theory that captures the full dynamic
evolution of the electronic and phononic populations and provides a microscopic description of the
arXiv:1708.01470v1 [cond-mat.mtrl-sci] 4 Aug 2017

transfer of energy delivered optically into electrons to the lattice. All essential nonequilibrium
energy processes, such as electron-phonon and phonon-phonon interactions are taken into account.
Moreover, as all required quantities are obtained from first-principles calculations, the model gives
an exact description of the relaxation dynamics without the need for fitted parameters. We apply
the model to FePt and show that the detailed relaxation is out-of-equilibrium for picoseconds.

Excitation of a material with an intensive, ultra- electronic and phononic populations and provides the
short optical pulse brings the material’s electrons into full dynamic description of the nonequilibrium relax-
a strongly out-of-equilibrium state which is immediately ation processes. The decisive quantities that govern the
followed by intense, correlated dynamics between the dynamical relaxation are the phonon mode dependent
electrons and other degrees of freedom in the material, electron-phonon coupling constants and the mode depen-
such as the lattice or spin systems [1–4]. The abil- dent phonon-phonon scattering terms. It is important to
ity to measure the ultrafast relaxation dynamics of the stress that both quantities have an explicit dependence
involved subsystems using pump-probe techniques has on the electronic temperature, the phonon momentum
led to the discovery of many unexpected phenomena, and branch, and, moreover, that they are fully derived
such as ultrafast demagnetization [2], change of magnetic from microscopic ab initio theory. Thus, as the theory
anisotropy [5], or coherent generation of magnetic preces- only uses quantities obtained from ab initio calculations,
sion [6, 7]. More recently, ultrafast generation of lattice it provides a parameter free description of the relaxation
strain waves [8–10], coherent control of atomic and molec- dynamics and could hence become of great value for mod-
ular dynamics [11], coherent phonon generation [12, 13], eling and predicting nonequilibrium dynamics following
and laser-induced superconductivity at high temperature ultrafast laser excitation. As an example, the model is
[14] have been reported. In spite of the importance of used to describe the ultrafast dynamics in FePt after fem-
the material’s nonequilibrium state in these laser-induced tosecond laser irradiation, illustrating as well the limita-
phenomena, it is surprising that most theoretical de- tions of the 2TM.
scriptions of the ensuing out-of-equilibrium dynamics are To describe the nonequilibrium time evolution of the
based on the widely-used two-temperature model (2TM) electronic and phononic degrees of freedom of the laser-
[15, 16], which assumes that the electronic and phononic excited material, we divide the out-of-equilibrium metal-
subsystems are separately in thermodynamic equilibrium lic system in different, independent subsystems that in-
[17]. teract with one another, schematically shown in Fig. 1.
Research on nonequilibrium states of matter has Specifically, we describe the lattice by dividing it into
emerged recently as an important area in condensed mat- N independent phonon subsystems, each of them corre-
ter physics (see, e.g., [18, 19]), and consequently several sponding to a specific branch ν and momentum q. They
improved models have been developed which incorporate interact with one another through phonon-phonon scat-
aspects of out-of-equilibrium electronic dynamics [20–23]. tering and with the electrons via electron-phonon scat-
However, these still lack a complete out-of-equilibrium tering. These interactions are phonon momentum and
description of the full system and its time evolution, and branch dependent. Therefore these phonon subsystem
contain parameters that are either fitted experimentally populations nνq evolve separately during the nonequi-
or chosen from a macroscopic system at equilibrium. Ad- librium dynamics and we can define a “lattice temper-
ditionally, recent investigations [10, 24, 25] have empha- ature” T`Q (with Q ≡ νq) for each of them (the mean-
sized that the assumption of thermal phonons could lead ing of this definition is discussed further below). The
to marked disagreement with experimental observations. impulsive laser excitation brings a part of the electrons
Here we propose a general theory to describe the ul- into a nonthermal state (Fig. 1). For the electrons we
trafast dynamics triggered by ultrashort laser pulses in follow the description of Carpene [20], in which the elec-
metals. Contrarily to the two-temperature model, our tronic system is divided in a thermal bath that contains
model employs an out-of-equilibrium description of the the majority of thermal electrons with temperature Te
2

rium of the lattice. Hence, the rates of electron-phonon


and phonon-phonon scattering are the quantities that de-
termine the system’s temporal evolution to equilibrium.
To achieve a theoretical formulation we make use of
the conservation of total energy and classical P
kinetic the-
ory. The total lattice energy is given by E`P= Q ~ωQ nQ
and the total electronic energy by Ee = 2 k k fk , where
~ωQ is the phonon energy, k the electron Bloch energy
[26] (with k ≡ nk), and nQ and fk are phonon and elec-
tron populations, respectively. The latter are in local
equilibrium given by the Fermi-Dirac and Bose-Einstein
distributions
h k −F i−1 h ~ωQQ i−1
fk = e kB Te (t) + 1 , and nQ = e kB T` (t) − 1 . (1)
FIG. 1. Scheme of treating out-of-equilibrium dynamics of
electrons and lattice. The electronic subsystem is excited via
femtosecond laser pulses. The major part of the electronic sys- We can define the rates of energy exchange as
tem is considered in local thermal equilibrium whereas a small
∂E` X scatt. X scatt.
part that absorbs most of the laser energy is in a nonthermal = ~ωQ ṅQ e−ph + ~ωQ ṅQ ph−ph , (2)
state. The electronic subsystems relax energy by interaction ∂t
Q Q
with different phononic subsystems, which also exchange en-
∂Ee X scatt. X scatt.
ergy among themselves (depicted by arrows). =2 k f˙k e−ph = − ~ωQ ṅQ e−ph , (3)
∂t
k Q

and in a laser-excited nonthermal electron distribution, where the dot stands for the time derivative and the
which relaxes driving energy into the thermal distribu- subscripts denote the different scattering processes that
tion through the electron-electron and electron-phonon change the distribution. The equivalence in Eq. (3) stems
interactions. The latter interaction conducts the energy from the conservation of total energy. The time deriva-
from the laser-excited electrons to the different lattice tives of the distribution functions due to different scatter-
subsystems, bringing them to different temperatures, i.e., ing terms can be derived from the classical kinetic theory
into a nonequilibrium state. The phonon-phonon scatter- by using the well-known Fermi’s golden rule of scatter-
ing causes the transferred energy to be shared between ing theory. By doing so, we obtain an extended version
the phonon subsystems, guiding them toward a common of the Bloch-Boltzmann-Peierls equations (see [27], and
lattice temperature, and therefore to a thermal equilib- Supplemental Material (SM) [28]),

scatt. 2π X
f˙k e−ph = − |Mkk0 |2 {fk (1 − fk0 ) [(nQ + 1)δ(k − k0 − ~ωQ ) + nQ δ(k − k0 + ~ωQ )]
~
Q
−(1 − fk )fk0 [(nQ + 1)δ(k − k0 + ~ωQ ) + nQ δ(k − k0 − ~ωQ )]}, (4)
scatt. 4π X
ṅQ e−ph
=− |Mkk0 |2 fk (1 − fk0 ) [nQ δ(k − k0 + ~ωQ ) − (nQ + 1)δ(k − k0 − ~ωQ )] , (5)
~
Q
scatt. 2π X
ṅQ ph−ph = |Φ−Qk0 k00 |2 {(nQ + 1)(nk0 + 1)nk00 δ(ωQ + ωk0 − ωk00 )+
~ 0 00
k k
(nQ + 1)(nk00 + 1)nk0 δ(ωQ + ωk00 − ωk0 ) − nQ nk0 (nk00 + 1)δ(ωQ + ωk0 − ωk00 )
−nQ nk00 (nk0 + 1)δ(ωQ + ωk00 − ωk0 ) + (nQ + 1)nk0 nk00 δ(ωQ − ωk00 − ωk0 )
−nQ (nk0 + 1)(nk00 + 1)δ(ωQ − ωk00 − ωk0 )} . (6)

Here Mkk0 and Φ−Qk0 k00 are the electron-phonon and assumption on the short time scale of the typical out-of-
phonon-phonon matrix elements, respectively. equilibrium process. The laser driving field induces the
nonequilibrium electronic distribution and enters in the
Equations (2) and (3) describe the energy flow between
rate equations as source term. Substituting Eqs. (4)–(6)
electron and phonon subsystems under the assumption
in (2) and (3), and making an expansion of the distribu-
that the diffusion term can be neglected, which is a valid
3

tion functions to second order in the phonon mode and a set of coupled differential equations which connect the
electron temperature differences (see SM [28]), we obtain time evolutions of the electron temperature Te and the
temperatures T`Q of the different phonon modes,

∂T`Q h i 1X 0 1 ∂Ue−ph
CQ = −GQ (T`Q − Te ) 1 + J(ωQ , T`Q )(T`Q − Te ) − CQ ΓQk0 T`Q − T`k + , (7)
∂t 9 0 NQ ∂t
k
∂Te X h i ∂U
e−e
Ce = GQ (T`Q − Te ) 1 + J(ωQ , T`Q )(T`Q − Te ) + , Q = Q1 , . . . , QN . (8)
∂t ∂t
Q

Ce and CQ are the temperature-dependent electronic and come evident below, the out-of-equilibrium model [Eqs.
phonon mode dependent specific heats, respectively, and (7) and (8)] leads to results that are markedly different
GQ and ΓQk0 are the mode-dependent electron-phonon from those of the 2TM. Nonetheless, it can be recog-
coupling constants and the mode-dependent phonon nized that under simplifying assumptions the 2TM can
linewidths which are caused by the phonon-phonon inter- be obtained from Eqs. (7) and (8). To this end, it is im-
∂Ue−ph
actions. ∂U∂t
e−e
and ∂t describe the rate transferred portant to realize that GQ is the nonequilibrium equiva-
from the laser-induced nonequilibrium electron distribu- lent of the electron-phonon coupling constant Gep that is
tion into the thermal electronic bath via electron-electron commonly used in the 2TM. Assuming a single phonon
interaction and into the lattice through electron-phonon temperature T` for all phonon modes and wavevectors
interaction, respectively [20]. J(ωQ , T`Q ) is a function and neglecting quadratic terms in the temperature dif-
of the mode-dependent phonon frequencies and temper- ference Te −PT` , along with setting Ue−ph = 0, and using
atures, which accounts for the second-order term in tem- Gep = N1Q Q GQ (valid under the assumptions of the
perature differences; its full form is given in the SM [28]. 2TM, see SM), one recovers the common 2TM [16].
To obtain a full solution of the out-of-equilibrium To recognize the importance of the mode and wavevec-
model, defined by Eqs. (7) and (8), it is necessary to tor dependent electron-phonon coupling constant GQ we
compute the material specific quantities, the phonon perform ab initio calculations of this quantity for fer-
and electronic specific heats and phonon mode-dependent romagnetic FePt, which is the prime material for high-
electron-phonon and phonon-phonon linewidths. These density optic and magnetic recording [35, 36]. Bulk FePt
can be conveniently calculated using the spin-polarized orders in the L10 structure in which the (001) planes are
density functional theory (DFT) in the local-density ap- alternatively occupied by Fe and Pt atoms. Our ab ini-
proximation (LDA). Here we have employed the elec- tio calculation of the ground state magnetic properties of
tronic structure code ABINIT [29]. Specifically, the FePt is in good agreement with previous work [37].
mode-dependent electron-phonon linewidths were com- In Fig. 2 we show our ab initio calculated mode-
puted as response function within the density func- dependent electron-phonon coupling constants GQ of
tional perturbation theory whereas the mode-dependent FePt at 300 K together with the phonon dispersions along
phonon-phonon linewidths due to phonon-phonon inter- 9
action were determined using many-body perturbation 8
theory in a third-order anharmonic Hamiltonian which 7
includes up to three-phonon scattering [30]. The cou-
Frequency (THz)

6
pled equations (7) and (8) are then solved numerically,
5
with the ab initio quantities, and without any free fitting
4
parameters.
3
We emphasize that the phonon branch and wavevector
2
q dependent phonon temperatures act here as an auxil-
iary quantity that allows us to use for each phononic sub- 1
space a Bose-Einstein distribution with a local tempera- 0
ture T`Q . While this temperature might not be measur- Γ X R Γ M A Γ Z
able, recent electron diffraction experiments showed that
nonequilibrium phonon populations in reciprocal space FIG. 2. Calculated phonon dispersions of ferromagnetic FePt
can be measured [31]. The electron temperature Te is along high-symmetry lines in the simple tetragonal BZ. The
conversely a quantity that can be obtained from pump- symbol size is proportional to the magnitude of the mode-
probe photoemission measurements [32–34]. As will be- dependent electron-phonon coupling constant GQ at 300 K.
4

9 103
6
25 Ce(Te) with FePt DOS

Ce (Jmol-1K-1)
20 2 2
8
Ce(Te)=π kB g(EF) Te/3
5 15
Gep (1017 Wm-3K-1)

10 7
5
4 102

Frequency (THz)
0 6
0 1 2 3 4 5 6 7 8
5

τ (ps)
3
4
2
101
(a)
Geq 3

1 Calculated Geq(300 K) 2

1
0
0 1 2 3 4 5 6 7 8 0 10
0

Electron Temperature, Te (103 K) X Γ M A Γ Z

FIG. 3. Calculated total electron-phonon coupling as a func- FIG. 4. Calculated phonon dispersions of ferromagnetic FePt
tion of electronic temperature (red line), compared with the along high-symmetry lines. The symbol color is related to the
computed total electron-phonon coupling at 300 K (yellow phonon lifetimes at 300 K due to phonon-phonon interaction
line), and with a theoretical estimation (a : purple line) [38]. (ser color bar on the right).
The inset shows the temperature-dependent electronic specific
heat.
R +∞
tron temperature [39], Ce (Te ) = −∞ (∂fk /∂Te )g()d.
high-symmetry lines in the simple tetragonal Brillouin This is in contrast to the commonly used calculation
zone (BZ). We can readily see that the electron-phonon of Ce from the Sommerfeld expansion of the free en-
coupling constants are larger for optical phonon modes, ergy, which provides a linear temperature dependence,
reaching several orders of magnitude differences between i.e. Ce (Te ) = π 2 kB
2
g(F )Te /3. The difference between
some points of the BZ (see e.g. optical phonons at the these two different descriptions is shown in the inset of
X point compared to acoustic phonons at the Γ point). Fig. 3.
These findings demonstrate the limitations of consider- Another key quantity of our model is the explicit incor-
ing a single Gep with the lattice at one local thermal poration and calculation of the phonon anharmonicities,
equilibrium, since the range of values of GQ is several or- that enter into the model through the mode-dependent
ders of magnitude. On account of the different coupling phonon linewidths ΓQk . This quantity is related to
strengths the laser-excited electrons will couple mainly the mode-dependent phonon lifetime P τQ due to phonon-
to the optical phonon modes. phonon interaction, via τQ = [2 k ΓQk ]−1 . Notably,
To account for the scattering processes of electrons in the 2TM it is assumed that such interactions are very
away from the Fermi surface, we have additionally in- strong and lead to very short phonon lifetimes, and there-
cluded an electron temperature dependence of the mode- fore to an immediate equilibration of the lattice. In Fig.
dependent electron-phonon coupling constant, using [39] 4 we show the calculated τQ of FePt at 300 K along high-
symmetry lines in the BZ. It can be clearly seen how
+∞
∂fk g()2
Z
different phonon modes and branches have different life-
GQ (Te ) = GQ d , (9) times, as indicated in the color changing when moving
−∞ ∂ g 2 (F )
along the phonon dispersions. The phonon lifetimes for
with g() being the electron density of states at energy . the optical branches range from 2 to 10 ps, while the
In Fig. 3 we illustrate the relevance of the electron tem- acoustic branches have phonon lifetimes larger than 10
peratureP for GQ . The calculation shows a fast growth ps, diverging at the Γ point. These timescales are much
of N1Q Q GQ with Te (red line), as compared with the larger than the initial ultrafast dynamics of the system.
temperature independent value (yellow line) and an esti- Thus, our calculations show that the assumption of an
mated value of Geq used recently [38] to reproduce exper- immediately thermalized lattice, as made in the 2TM,
imental data with a 2TM (purple line). At high electronic is not tenable, since the phonon-phonon lifetimes are at
temperatures, Gep is an order of magnitude larger, which least one order of magnitude larger than the electron-
will clearly influence the early dynamics of the system af- phonon lifetimes.
ter laser irradiation. The combination of the main results shown in Figs.
Also, we would like to stress that to correctly assess the 2 and 4, namely strong mode-dependent excitation of
dependence of the electron heat capacity on the electronic phonon modes via electron-phonon coupling and slow
temperature, in our model we calculate it as the deriva- lattice thermalization through phonon-phonon interac-
tive of the total electron energy density against the elec- tion, suggests that the lattice remains out-of-equilibrium
5

not only at sub-ps timescales, but even on much larger 1800


timescales. This is a striking difference with respect to 1600 Te
the model by Waldecker et al. [24, 25], who predicted Range Tph

that phonons thermalize within a few picoseconds. The 1400 Av. Tph
440
2TM Te
disagreement with their results is a consequence of the 1200 2TM Tph 400
different theoretical description they proposed. They use

T (K)
360
a (non-microscopically) derived phonon branch depen- 1000
320
dence to account for the different strength in the electron-
800 280
phonon interaction, and obtain the phonon-phonon cou-
20 40 60 80 100
pling by fitting experimental data under the assumption 600
of equal phonon-phonon interaction strength between
400
branches. This assumption is however not justified as we
have seen in Fig. 4, where the phonon lifetime is strongly 200
mode dependent. Also, as we have seen in Fig. 2, the 0 2 4 time (ps) 6 8 10
electron-phonon coupling constants do in fact strongly
vary within the same phonon branch. FIG. 5. Ab initio calculated temporal evolution of the elec-
To provide the first ab initio description of laser- tronic temperature (blue line), average phonon temperature
induced nonequilibrium dynamics in a metal, and to (green line) and temperature range within which all the
answer the question for how long the lattice is out-of- phonon-mode temperatures are contained (red area). The
inset shows the temporal evolution from 20 to 100 ps. Dashed
equilibrium, we have numerically solved our nonequilib- lines show the results of the ab initio 2TM for FePt.
rium rate Eqs. (2) and (3), considering NQ = 163 phonon
modes (enough to achieve good numerical convergence).
In Fig. 5 we plot the temporal evolution of the elec-
tronic and phonon mode temperatures for times up to
100 ps. The electronic temperature (blue line) increases
in FePt [38]. The results, shown in the SM, are strikingly
rapidly, reaching up to 1575 K in 196 fs, followed by an
different. Not only does Te reach higher values (around
exponential decrease with a decay time of about 1225 fs
2500 K) in much shorter time (about 100 fs), also the fi-
(light shaded area). The final electronic temperature
nal electronic temperature is higher, about 550 K. Ad-
reached (after 100 ps) is 366 K. We also show the phonon
ditionally, the electron-lattice dynamics is much faster,
mode-dependent range of temperatures versus time (red
reaching a common thermal equilibrium after only 1.5 ps.
area). The individual evolution of each phonon mode is
In contrast, our results evidence that the lattice remains
left out for sake of simplicity. Initially, the maximum val-
out-of-equilibrium not only during short time scales after
ues in the range, which stem from optical phonon modes,
laser excitation but also on large timescales, and provide
increase, reaching a value of about 590 K at 1660 fs, fol-
a clear example that a complete nonthermal modeling is
lowed by a slow monotonic decrease. On the other hand,
needed to describe the out-of-equilibrium dynamics.
the minimum values, which stem from acoustic modes
close to the Γ point, keep increasing very slowly, reaching
a temperature of 313 K at 100 ps. Notably, the phonon In conclusion, we have developed a nonequilibrium the-
mode temperatures cover a range of hundreds of Kelvins ory to describe the out-of-equilibrium dynamics triggered
during several picoseconds, evidencing a nonequilibrium by ultrashort laser pulses. The model enables a fully ab
behavior. To estimate the weight of each phonon mode initio determination of the out-of-equilibrium dynamics.
in the thermalization process and to avoid the singular Our simulations for FePt unambiguously reveal that, in
behavior of some phonon modes, such as phonon modes contrast to previous understanding, the lattice is not in
close to Γ, we also show the average phonon tempera- equilibrium even after 20 ps. As ultrafast nonequilibrium
ture (green line), calculated as the sum of phonon mode dynamics of materials is a strongly emerging research
temperatures over NQ . Although the average phonon area and since our theory can provide a fully parameter-
temperature reaches an almost converged value very fast free description of the ensuing dynamics, we expect it to
(within the first picoseconds), it is stunning to observe become a valuable tool for future modeling of ultrafast
that this temperature still differs from the electronic tem- relaxation dynamics of laser-excited metals.
perature even at 100 ps, which confirms a continuing en-
ergy flow between the electronic and phononic systems. We thank H. Dürr and M. Bargheer for valuable discus-
Moreover, the phonon modes temperatures cover an in- sions. This work has been funded through the Swedish
terval of about 50 K at 100 ps. research Council (VR), the K. and A. Wallenberg Foun-
For comparison we have computed the electronic tem- dation (Grant No. 2015.0060) and the Röntgen-Ångström
perature evolution using the 2TM with the parameters Cluster. We also acknowledge support from the Swedish
obtained recently to simulate ultrafast demagnetization National Infrastructure for Computing (SNIC).
6

125102 (2014).
[24] L. Waldecker, R. Bertoni, R. Ernstorfer, and J. Vor-
berger, Phys. Rev. X 6, 021003 (2016).
[1] J. G. Fujimoto, J. M. Liu, E. P. Ippen, and N. Bloem- [25] L. Waldecker, T. Vasileiadis, R. Bertoni, R. Ernstorfer,
bergen, Phys. Rev. Lett. 53, 1837 (1984). T. Zier, F. H. Valencia, M. E. Garcia, and E. S. Zijlstra,
[2] E. Beaurepaire, J.-C. Merle, A. Daunois, and J.-Y. Phys. Rev. B 95, 054302 (2017).
Bigot, Phys. Rev. Lett. 76, 4250 (1996). [26] The factor 2 stems from the spin degeneracy, but a more
[3] B. Koopmans, G. Malinowski, L. F. Dalla, D. Steiauf, general form can be given.
M. Fähnle, T. Roth, M. Cinchetti, and M. Aeschlimann, [27] J. M. Ziman, Electrons and Phonons: The Theory
Nat. Mater. 9, 259 (2010). of Transport Phenomena in Solids (Oxford University
[4] B. Liao, A. A. Maznev, K. A. Nelson, and G. Chen, Nat. Press, Oxford, 1960).
Commun. 7, 13174 (2016). [28] Supplementary Material is available for this article.
[5] F. Hansteen, A. Kimel, A. Kirilyuk, and T. Rasing, [29] X. Gonze, B. Amadon, P.-M. Anglade, J.-M. Beuken,
Phys. Rev. Lett. 95, 047402 (2005). F. Bottin, P. Boulanger, F. Bruneval, D. Caliste, R. Cara-
[6] G. Ju, A. V. Nurmikko, R. F. C. Farrow, R. F. Marks, cas, M. Cote, T. Deutsch, L. Genovese, P. Ghosez, M. Gi-
M. J. Carey, and B. A. Gurney, Phys. Rev. Lett. 82, antomassi, S. Goedecker, D. R. Hamann, P. Hermet,
3705 (1999). F. Jollet, G. Jomard, S. Leroux, M. Mancini, S. Mazevet,
[7] M. van Kampen, C. Jozsa, J. T. Kohlhepp, P. LeClair, M. J. T. Oliveira, G. Onida, Y. Pouillon, T. Rangel, G.-
L. Lagae, W. J. M. de Jonge, and B. Koopmans, Phys. M. Rignanese, D. Sangalli, R. Shaltaf, M. Torrent, M. J.
Rev. Lett. 88, 227201 (2002). Verstraete, G. Zerah, and J. W. Zwanziger, Comp. Phys.
[8] J.-W. Kim, M. Vomir, and J.-Y. Bigot, Phys. Rev. Lett. Commun. 180, 2582 (2009).
109, 166601 (2012). [30] A. Togo, L. Chaput, and I. Tanaka, Phys. Rev. B 91,
[9] J.-W. Kim, M. Vomir, and J.-Y. Bigot, Sci. Rep. 5, 8511 094306 (2015).
(2015). [31] T. Chase, M. Trigo, A. H. Reid, R. Li, T. Vecchione,
[10] T. Henighan, M. Trigo, S. Bonetti, P. Granitzka, X. Shen, S. Weathersby, R. Coffee, N. Hartmann, D. A.
D. Higley, Z. Chen, M. P. Jiang, R. Kukreja, A. Gray, Reis, X. J. Wang, and H. A. Dürr, Appl. Phys. Lett.
A. H. Reid, E. Jal, M. C. Hoffmann, M. Kozina, S. Song, 108, 041909 (2016).
M. Chollet, D. Zhu, P. F. Xu, J. Jeong, K. Carva, P. Mal- [32] C.-K. Sun, F. Vallée, L. Acioli, E. P. Ippen, and J. G.
donado, P. M. Oppeneer, M. G. Samant, S. S. P. Parkin, Fujimoto, Phys. Rev. B 48, 12365 (1993).
D. A. Reis, and H. A. Dürr, Phys. Rev. B 93, 220301 [33] C. Guo, G. Rodriguez, and A. J. Taylor, Phys. Rev. Lett.
(2016). 86, 1638 (2001).
[11] H. Rabitz, R. Vivie-Riedle, M. Motzkus, and K. Kompa, [34] M. Lisowski, P. A. Loukakos, U. Bovensiepen, J. Stähler,
Science 288, 824 (2008). C. Gahl, and M. Wolf, Appl. Phys. A 78, 165 (2004).
[12] M. Harmand, R. Coffee, M. R. Bionta, M. Chol- [35] C.-H. Lambert, S. Mangin, B. S. D. C. S. Varaprasad,
let, D. French, D. Zhu, D. M. Fritz, H. T. Lemke, Y. K. Takahashi, M. Hehn, M. Cinchetti, G. Malinowski,
N. Medvedev, B. Ziaja, S. Toleikis, and M. Cammarata, K. Hono, Y. Fainman, M. Aeschlimann, and E. E. Fuller-
Nat. Photon. 7, 215 (2013). ton, Science 345, 1337 (2014).
[13] A. M. Lindenberg, S. L. Johnson, and D. A. Reiss, Annu. [36] R. John, M. Berritta, D. Hinzke, C. Müller, T. San-
Rev. Mater. Res. 47, 426 (2017). tos, H. Ulrichs, P. Nieves, J. Walowski, R. Mondal,
[14] M. Mitrano, A. Cantaluppi, D. Nicoletti, S. Kaiser, O. Chubykalo-Fesenko, J. McCord, P. M. Oppeneer,
A. Perucchi, S. Lupi, P. D. Pietro, D. Pontiroli, M. Ricco, U. Nowak, and M. Münzenberg, Sci. Rep. 7, 4114 (2017).
S. R. Clark, D. Jaksch, and A. Cavalleri, Nature 530, [37] Our ab initio calculated lattice parameters of ferromag-
461 (2016). netic FePt, a = 2.72 Å and c = 3.76 Å (c/a = 1.38),
[15] M. I. Kaganov, I. M. Lifshitz, and L. V. Tanatarov, Sov. are in agreement with experimental values (a = 2.73 Å,
Phys. JETP 4, 173 (1957). c = 3.72 Å, and c/a = 1.36). The computed ground state
[16] S. I. Anisimov, B. L. Kapeliovich, and T. L. Perel’man, magnetic moments M (Fe) = 2.94 µB , M (Pt) = 0.33 µB
Sov. Phys. JETP 39, 375 (1974). are also in good agreement with previous calculations and
[17] P. B. Allen, Phys. Rev. Lett. 59, 1460 (1987). experiments [40, 41].
[18] L. Stojchevska, I. Vaskivskyi, T. Mertelj, P. Kusar, [38] J. Mendil, P. Nieves, O. Chubykalo-Fesenko, J. Walowski,
D. Svetin, S. Brazovskii, and D. Mihailovic, Science 344, T. Santos, S. Pisana, and M. Münzenberg, Sci. Rep. 4,
177 (2014). 3980 (2014).
[19] H. Aoki, N. Tsuji, M. Eckstein, M. Kollar, T. Oka, and [39] Z. Lin, L. V. Zhigilei, and V. Celli, Phys. Rev. B 77,
P. Werner, Rev. Mod. Phys. 86, 779 (2014). 075133 (2008).
[20] E. Carpene, Phys. Rev. B 74, 024301 (2006). [40] P. M. Oppeneer, J. Magn. Magn. Mater. 188, 275 (1998).
[21] B. Y. Mueller and B. Rethfeld, Phys. Rev. B 87, 035139 [41] J. Lyubina, I. Opahle, M. Richter, O. Gutfleisch, K.-H.
(2013). Müller, L. Schultz, and O. Isnard, Appl. Phys. Lett. 89,
[22] K. Carva, M. Battiato, D. Legut, and P. M. Oppeneer, 032506 (2006).
Phys. Rev. B 87, 184425 (2013).
[23] V. V. Baranov and V. V. Kabanov, Phys. Rev. B 89,

You might also like