• Aucun résultat trouvé

Counter-rotating standing spin waves: A magneto-optical illusion

N/A
N/A
Protected

Academic year: 2021

Partager "Counter-rotating standing spin waves: A magneto-optical illusion"

Copied!
7
0
0

Texte intégral

(1)

HAL Id: hal-01490261

https://hal.archives-ouvertes.fr/hal-01490261

Preprint submitted on 15 Mar 2017

HAL is a multi-disciplinary open access

archive for the deposit and dissemination of

sci-entific research documents, whether they are

pub-lished or not. The documents may come from

teaching and research institutions in France or

abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est

destinée au dépôt et à la diffusion de documents

scientifiques de niveau recherche, publiés ou non,

émanant des établissements d’enseignement et de

recherche français ou étrangers, des laboratoires

publics ou privés.

Counter-rotating standing spin waves: A

magneto-optical illusion

S. Shihab, L. Thevenard, A. Lemaitre, C. Gourdon

To cite this version:

S. Shihab, L. Thevenard, A. Lemaitre, C. Gourdon. Counter-rotating standing spin waves: A

magneto-optical illusion. 2016. �hal-01490261�

(2)

S. Shihab,1 L. Thevenard,1 A. Lemaˆıtre,2 and C. Gourdon1, ∗

2

1Sorbonne Universit´es, UPMC Univ Paris 06, CNRS-UMR 7588, 3

Institut des NanoSciences de Paris, F-75005, Paris, France

4

2

Centre de Nanosciences et Nanotechnologies, CNRS, Univ. Paris-Sud,

5

Universit´e Paris-Saclay, 91460 Marcoussis, F-91460 France

6

(Dated: January 13, 2017)

7

We excite perpendicular standing spin waves by a laser pulse in a GaMnAsP ferromagnetic layer

8

and detect them using time-resolved magneto-optical effects. Quite counter-intuitively, we find the

9

first two excited modes to be of opposite chirality. We show that this can only be explained by

10

taking into account absorption and optical phase shift inside the layer. This optical illusion is

11

particularly strong in weakly absorbing layers. These results provide a correct identification of spin

12

waves modes, enabling a trustworthy estimation of their respective weight as well as an unambiguous

13

determination of the spin stiffness parameter.

14

PACS numbers: 75.78.Jp ,75.30.Ds,75.50.Pp

15

Since pioneering work on nickel [1], laser-induced

mag-16

netization dynamics has been widely used to investigate

17

ultrafast magnetic processes not only in magnetic metals

18

[2, 3], but also in magnetic semiconductors [4–9] and

in-19

sulators [10, 11], exploring the fundamentals of light-spin

20

interaction in view of a full and ultrafast optical control

21

of magnetic order.

22

Ultrashort pulses can trigger a wide variety of

pro-23

cesses, including ultrafast demagnetization [1], full

mag-24

netization reversal [12] as well as coherent precession

25

[2, 4, 10]. In magnetically ordered materials, ferro- and

26

ferrimagnets, as well as in antiferromagnets, the coherent

27

magnetization dynamics arises from collective spin

exci-28

tations, spin waves (SW) (or magnons, their quanta),

29

which attract a considerable interest motivated by their

30

possible use as information carriers in magnonics

appli-31

cations [13, 14]. Magnons are versatile excitations since

32

their dispersion curves, comprising magnetostatic and

ex-33

change modes [15], can be tuned by a magnetic field or

34

by micro- or nanostructuring the material in any of its

35

dimensions [14]. For instance, perpendicular standing

36

spin-wave (PSSW) modes in a single nanometric layer

37

of thickness L with free boundary conditions (no surface

38

anisotropy) will have their wave-vector quantized by an

39

integer p (k = pπ/L), and their energy by p2,

propor-40

tionally to the spin stiffness D.

41

SWs can be studied in the frequency domain by,

42

e.g., ferromagnetic resonance (FMR) experiments or

Bril-43

louin light scattering as well as in the time-domain by

44

time-resolved laser pump-probe (PP) experiments using

45

magneto-optical effects. The latter distinguish

them-46

selves in various ways from the former. The coherent

47

excitation of several PSSWs with different frequencies,

48

from the sub-GHz to the THz range, is made possible

49

by the wide frequency spectrum of the pulsed excitation

50

induced by the femtosecond pump laser pulses. In a

sin-51

gle time-scan PP experiments can detect several coherent

52

SW modes, and provide their time period, their

respec-53

tive phases and their damping. Coherent control

experi-54

ments using two pump beams can be performed [16, 17].

55

Furthermore the possibility to fully reconstruct the

mag-56

netization trajectory using two different magneto-optical

57

effects [18, 19] brings a deep insight into magnetization

58

dynamics.

59

Whereas the excitation mechanism of SWs by optical,

60

acoustical or magnetic field pulses [14, 20–22], has been

61

thoroughly discussed, their optical detection has been

62

much less addressed [4, 23]. In particular, in contrast to

63

the cavity FMR detection of SWs which was modeled a

64

long time ago [24], the respective amplitudes of optically

65

detected SWs remained unexplained [2, 9, 14, 25]. In this

66

Letter, we provide a comprehensive model to explain the

67

large amplitude of these non-uniform SWs that should

68

not be detectable in the framework of simple models [4].

69

We moreover present intriguing experimental results of

70

apparent different chiralities for SWs of different

pari-71

ties. We show that they can only be explained by this

72

theoretical description of the magneto-optical detection

73

through the Kerr and Voigt effects that takes into

ac-74

count the absorption depth and the optical phase shift

75

inside the layer. Important consequences are expected

76

when the SW wavelength (determined by the layer

thick-77

ness) becomes a few tenth of the wavelength of light in

78

the material, λ/η, where η is the refractive index. In

par-79

ticular, ignoring this effect can lead to an erroneous

de-80

termination of the SW stiffness, and of the relative mode

81

amplitude, a signature of the up-to-now elusive magnon

82

excitation mechanisms. We show that the optical phase

83

shift can lead to a striking and non-intuitive optical

ef-84

fect in the detection of SWs: an apparent reversal of the

85

magnetization rotation direction for SWs of odd parity

86

with respect to the layer mid-plane. A key result of this

87

paper is that the optical phase shift provides a unique

88

tool for the determination of the SW mode number, or

89

in other words its parity.

90

We study thin layers of the ferromagnetic

(3)

2 tor alloy GaMnAsP. Most samples show only one PSSW

92

mode in the FMR spectra while one or two modes are

op-93

tically detected in the TRMO signal [8, 29]. The results

94

presented here are obtained in an in-plane magnetized

95

GaMnAsP layer with thickness L=50 nm and

Phospho-96

rus concentration 4.3 % grown on a (001) GaAs

sub-97

strate by molecular beam epitaxy and annealed 1 hour

98

at 250◦C. The effective Mn concentration is 4 % and

99

the Curie temperature is 85 K. The anisotropy constants

100

were determined by FMR. PP experiments are carried

101

out at T =12 K in zero external magnetic field after a

102

60 mT in-plane initialization of the magnetization

direc-103

tion. The laser source is a 76 MHz Ti:Sapphire laser

104

at a wavelength λ=700 nm. To limit thermal effects,

105

low pump and probe fluence are used (1 µJ cm−2 and

106

0.4 µJ cm−2, respectively) [30]. The pump beam is

mod-107

ulated at 50 kHz. The pump-induced magnetization

dy-108

namics is detected as a function of the pump-probe delay

109

through the rotation of the probe beam linear

polariza-110

tion detected by a balanced optical diode bridge and a

111

lock-in amplifier. The static rotation and ellipticity

sig-112

nals are obtained with the probe beam only.

113

The existence of circular and linear magnetic birefrin-gence/dichroism [31] makes the TRMO signal sensitive to both the out-of-plane δθ and the in-plane δφ components of the transient magnetization (Fig. 1(a)). This allows for the reconstruction of the magnetization trajectory using the expected dependency of the rotation angle δβron the

probe polarization angle β (Ref. 19 and Suppl-info): δβrexp(t) =Krδθexp(t) + 2Vrδφexp(t) cos 2(β − φ0)

− 2Vr

δM (t)

M sin 2(β − φ0) , (1) where Kr and Vr are the static Kerr and Voigt rotation

114

coefficients, M is the magnetization vector modulus, and

115

φ0is the in-plane equilibrium angle of the magnetization.

116

Figure 1(b) shows the dependence of the TRMO signal on

117 118

the incident probe polarization. The signal is fitted with

119

u(t) + v(t) sin 2β + w(t) cos 2β from which we obtain the

120

δθexp(t) and δφexp(t) functions taking into account the

121

magnetization equilibrium angle φ0(M close to [100]) as

122

shown in Fig. 1(c) (see Suppl-info). The plot of the

tra-123

jectory, δθexp(t) versus δφexp(t) shown in Fig. 1(d) reveals

124

a very complex dynamics that actually results from the

125

contributions of two SWs that clearly appear in δθexp(t)

126

and δφexp(t) and their Fourier transform (Fig. 1(c) and

127

inset). Let us note the large ratio (≈ 0.6) of the

am-128

plitude of the second SW with respect to the first one.

129

δθexp(t) and δφexp(t) are fitted with two damped

oscil-130

lating signals and a sum of two exponentials that reflects

131

the shape of the laser-induced pulsed excitation (τ1=0.03

132

ns, τ2=1 ns). Plotting separately the trajectory for each

133

SW (f0=2.36 GHz and f1=3.90 GHz) in Fig. 1(e, f)

re-134

sults in a much clearer picture of the magnetization

pre-135

cession. Surprisingly, the magnetization seems to rotate

136

in opposite directions for the two SWs. As we shall see

137

FIG. 1. (a) Reference frame. (b) Dependence of the TRMO signal δβ on the probe beam linear polarization with respect to the sample crystallographic axes. (c) Time dependence of δθexpand δφexp. Inset: Fourier transform amplitude of δθexp.

(d) Optically detected magnetization trajectory. (e), (f), (g) Decomposition of the experimental magnetization trajectory (d) into two oscillating signals at frequencies f0=2.36 GHz

and f1=3.90 GHz and a non-oscillating signal, respectively.

below, this is an “optical illusion” resulting from an

opti-138

cal phase shift inside the layer. To demonstrate this, we

139

first describe the SW excitation by a laser pulse using the

140

Landau-Lifshitz-Gilbert (LLG) equation and appropriate

141

boundary conditions at the top and bottom interfaces.

142

We then calculate the detected magneto-optical signal

143

using a multi-layer and transfer matrix model, which we

144

show to be indispensable to recover the observed SWs

145

chiralities.

146

The SW space and time profiles are obtained by solv-ing the LLG equation within the small precession angle approximation for in-plane static magnetization:

˙ δφ = γFθθδθ − D ∂2δθ ∂z2  + αGδ ˙θ ˙ δθ = −γFφφδφ − D ∂2δφ ∂z2 + δBexc(z, t)  − αGδ ˙φ . (2) F is equal to E/M where E is the magnetic anisotropy

147

energy [28]. αG is the Gilbert damping. Fi,j = ∂

2F

∂i∂j are

148

the second derivatives of F with respect to the angles

149

(i, j). Since it was shown that in thin (Ga,Mn)(As,P)

lay-150

ers the lowest frequency PSSW is a nearly uniform mode,

151

independent of the layer thickness [8, 9], the boundary

152

conditions at z=0 and z = L were chosen to ensure very

153

weak surface pinning, giving nearly flat δθ(z) and δφ(z)

(4)

profiles at any t for this mode (see Suppl-info). Under

155

symmetric conditions the p-PSSW eigenmodes (p=0, 1,

156

2...) are even (odd) with respect to the layer mid-plane

157

for even (odd) p. δBexc(z, t) is the optically induced

ef-158 159

fective field that launches the magnetization precession.

160

It arises from the in-plane rotation of the magnetic easy

161

axis induced by transient thermal effects or by the optical

162

spin-orbit torque [32]. δBexc(z, t) is taken as a product

163

of time and space functions f (t)g(z). f (t) is chosen so

164

that the calculated δθ(t) and δφ(t) match the

experimen-165

tal ones. g(z) is taken as a Fourier series over the PSSW

166

eigenmodes (see Suppl-info). The depth dependence of

167

the magnetization trajectory for the p=0, 1, 2 SW modes

168

(with time dependence cos(2πfpt+φp)) is shown in Fig. 2

169

(a-c), respectively. It is seen that inside the layer the

di-170

rection of rotation is the same for the three modes. For

171

modes 1 and 2 the magnetization vector experiences a

172

π-shift at each node, but its direction of rotation does

173

not change with z (Fig. 2(b,c)).

174

The TRMO signal is then calculated using a multi-layer and transfer matrix model to obtain the Kerr and Voigt rotation angle and ellipticity. The magnetic layer of thickness L is divided in N sub-layers with magnetization components M (mx,y+ δmx,y,z(zi, t)) (Fig. S1 of

Suppl-info). The calculation is performed for normal incidence of light along the z-direction and linear polarization mak-ing an angle β with the x-axis. The theoretical dynam-ical polarization rotation δβth

r (t) is obtained by taking

the limit of an infinite number of sub-layers (N → ∞). δβthr (t) is the sum of the Kerr and Voigt rotation angles, δβthr (t) = δβKthr(t) + δβ th Vr(t) cos 2(β − φ0) with: δβKthr(t) = −4π λ Re " n2Q n2− 1 Z L 0 ei4πnzλ δθ(z, t)dz # δβVthr(t) = 4π λIm " n2B (n2− 1) Z L 0 ei4πnzλ δφ(z, t)dz # , (3)

where B = B1+ Q2. Q (∝ M ) and B1 (∝ M2) are the

175

elements of the dielectric permittivity tensor describing

176

the Kerr and Voigt effects, respectively (see Suppl-info).

177

n = η + iκ is the layer mean complex refractive index

178 and ei4πnz λ = e−αzei 4πηz λ with α = 4πκ λ the absorption 179 coefficient. 180

The important result is the modulation of the

spa-181

tial dependence of the δθ(z, t) and δφ(z, t)

magnetiza-182

tion components by the optical phase factor ei4πηzλ that 183

reflects the propagation of light from the surface to the

184

depth z and back. The phase factor is damped by the

185

e−αz absorption factor. Therefore, in the case of strong

186

absorption as in metallic layers, the TRMO signal will be

187

sensitive only to the SW amplitude very close the surface

188

within the absorption depth. In the case of weak

absorp-189

tion and layer thickness L comparable to a fraction of

190

the light wavelength inside the material λ/η, the optical

191

phase shift plays a crucial role. It is precisely the case of

192

the sample studied here where L ≈ λ/4η.

193

Actually, for static magnetization, the importance of a

194

phase shift factor that makes the magneto-optical effects

195

sensitive to the magnetization at a specific depth inside

196

single or multiple ferromagnetic layers was theoretically

197

pointed out [33, 34] and evidenced experimentally in the

198

90s [35] and recently in GaMnAs layers [36]. Similar

199

ideas were at play when conceiving magneto-optical

sen-200

sors using magnetic quantum wells in optical cavities [37]

201

or magneto-photonic crystals with enhanced Faraday

ro-202

tation [38, 39]. However, these ideas had so far not been

203

applied to the time-resolved optical detection of PSSWs

204

in ferromagnetic layers.

205

The dynamical rotation angles δβKth

r(t) and δβ

th Vr(t)

are calculated according to Eq. 3 with the real and imaginary parts of Q and B extracted from the static rotation and ellipticity using Eqs. S12 and S13 of Suppl-info. In order to compare the opti-cally detected magnetization trajectory and the theo-retical one we define δθopt(z, t) and δφopt(z, t) so that

δβth

r (t) = Krhδθoptiz(t) + 2Vrhδφoptiz(t) cos 2(β − φ0)

where h· · · iz = (1/L)RL

0 · · · (z)dz denotes the average

value over the layer thickness, Kr and Vr are given by

Eqs. S11 and S18 of Suppl-info, respectively and δθopt(z, t) = − 1 Kr Re φoptnQ (n2− 1)e i4πnz λ  δθ(z, t) , (4) δφopt(z, t) = 1 2Vr Im φoptnB (n2− 1)e i4πnz λ  δφ(z, t) , (5) with φopt = 4πnL/λ the complex optical phase.

206

Figures 2(d-f) show the depth dependence of the

207

(δφopt, δθopt) parametric plot. The effect of the phase

208

factor ei4πnz

λ is clearly observed when compared to the 209

(δφ, δθ) trajectory shown in Figs 2 (a-c). Despite the

210

difference between Figs 2 (a) and (d), for a uniform SW

211

mode the optically detected trajectory is the same as

212

the simple average of the magnetization dynamics over

213

the layer as expected from the expression of (Kr, V r)

214

(Eqs. S11, S18) and the definition of (δθopt, δφopt). This

215

is indeed verified on Figs 2 (g,j). The quasi-uniform mode

216

0 is detected as rotating clockwise (CW) with time as

dic-217

tated by the sign of the gyromagnetic factor. An opposite

218

rotation direction can be expected for a non-uniform SW

219

mode if the sign of either hδθopti or hδφopti is changed

220

with respect to that of hδθi or hδφi. This is indeed what

221

is found for the p=1 odd mode, which rotates

counter-222

clock-wise (CCW) as shown in Fig. 2(k). The p=2 mode

223

rotates CW like the p = 0 mode (Fig. 2(l)).

224

In order to explain why odd modes may exhibit an apparently inverted direction of rotation when de-tected optically, we take a simplified model and calcu-late the sign of the amplitude ratio δθopt(t)/δφopt(t).

The damped p-PSSW mode at frequency fpis expressed

as δθp(z, t) = aθpexp (−χpt) cos (2πfpt) cos (pπz/L),

(5)

Ne-4

FIG. 2. (left) Magnetization trajectory in the depth of the ferromagnetic layer for the p=0 (a), p=1 (b), p=2 (c) PSSW modes and corresponding detected trajectory in (g), (h), (i) assuming that the optical signal would result from a depth-averaged amplitude. (right) (d), (e), (f) Theoretical effective magnetization trajectory in the depth of the ferromagnetic layer taking into account the optical phase factor and corresponding optically detected trajectory in (j), (k), (l). L=50nm, λ=700 nm, the other parameters are given in Suppl-info.

glecting absorption, the optical precession amplitudes normalized to the excitation amplitudes aθp and aφp are:

δθpopt= −4π Kr η η2− 1 Z ` 0

ReQ ei4πu cos (pπu/`) du

δφoptp = 4π 2Vr η η2− 1 Z ` 0

ImB ei4πu cos (pπu/`) du , (6) where ` = L/(λ/η). It is straightforward to show that

225

for even p (even modes) the ratio rp = δθoptp /δφoptp is

226

positive and equal to 1. For odd modes, rp can on

227

the contrary be positive or negative depending on the

228

layer thickness and the ratios Bi/Br and Qi/Qr of the

229

imaginary and real parts of B and Q, respectively. It

230

is moreover independent of p and is given by rodd p = 231 −(Bi BrCl+ Sl)( Qi QrCl+ Sl)(Cl − Bi BrSl) −1(C l− QQi rSl) −1 232

with Cl = cos(2π`) and Sl = sin(2π`). Therefore the

233

possibility to change the sign of only one of the δθ and

234

δφ components (rp <0) and hence to observe a change

235

of the direction of rotation is achieved exclusively for the

236

odd SW modes. This result is not fully conserved when

237

taking into account absorption as can be seen in Fig. 3

238

where the direction of rotation (CW, CCW) given from

239

the sign of rp is plotted in (dark/light) gray scale.

How-240

ever, given our parameters, r1is always negative for L in

241

the range 26-72 nm encompassing the layer thickness of

242

our sample (50 nm) while r0and r2 are positive. This is

243

an important result of this paper as it provides a tool to

244

identify PSSW modes.

245 246

This model also accounts very well for the large

am-247

plitude ratio of the high/low frequency modes observed

248 0 20 40 60 80 100 0 1 2 M o d e # Layer thickness (nm) r p >0 apparent CW rotation r p <0 apparent CCW rotation 0.0 0.1 0.2 0.3 0.4 0.5 Layer thickness (

FIG. 3. Theoretical optically detected direction of rotation of the magnetization vector for p=0, 1, and 2 PSSWs modes from the amplitude ratio rpof δθoptp and δφ

opt

p (see text). The

dashed line indicates the layer thickness. λ=700 nm, η=3.67, κ=0.1.

experimentally. If the optically detected signal were

pro-249

portional to the average of δθ or δφ over the layer

thick-250

ness [4], hδθiz and hδφiz, only the uniform PSSW mode

251

should be detected in the case of free boundary

condi-252

tions, all the higher ones having zero integral. This is

253

illustrated in Fig. 2(h,i) where the calculated (hδθi , hδθi)

254

signal is zero for the odd p=1 mode and very small for

255

the p = 2 mode (it would be strictly zero for zero surface

256

anisotropy). In order to observe higher modes a strong

257

surface pinning would be necessary to give them a

(6)

zero integral. Even then, only the even modes would be

259

detectable since the odd ones would keep a zero integral

260

for symmetric boundary conditions [4] . Our results

def-261

initely prove that the high-frequency mode is the first

262

odd mode. This reconciles results obtained by different

263

groups on the determination of the spin stiffness constant

264

that differ by a factor of 4 depending on whether the

high-265

frequency PSSW mode is identified as the p = 1 or p = 2

266

mode [8, 29]. Furthermore it explains why the TRMO

267

signal can show PSSW modes that are not observed in

268

FMR spectra.

269

In this Letter we have highlighted the role of the

op-270

tical phase shift in the amplitude of optically detected

271

SW modes. This solves the mystery of counter-rotating

272

SWs but, more importantly, provides a definite

assign-273

ment of SW mode number, thereby enabling a reliable

274

determination of the spin stiffness constant with only

275

two optically detected modes. The comprehensive model

276

developed here, which comprises both Kerr and Voigt

277

effects, provides useful guidelines (through Eq. 3) for

278

optimizing the optical detection of SWs. It may also

279

explain varying SW amplitude ratios observed in

differ-280

ent layers/materials [2, 29]. It can be straightforwardly

281

extended to longitudinal Kerr and Faraday effects, for

282

which similar effect of the complex optical phase are

283

expected, and therefore be applied to any kind of

ex-284

perimental geometry and magnetic layer, whether ferro-,

285

ferri-, or antiferromagnetic, from metals to insulators.

286

We thank F. Perez and B. Jusserand for fruitful

dis-287

cussions, B. Eble for cryogenic data, B. Gallas for

ellip-288

sometric data, and M. Bernard, F. Margaillan, F.

Bre-289

ton, S. Majrab, C. Lelong for technical assistance. This

290

work has been supported by UPMC (Emergence 2012),

291

Region Ile-de-France (DIM Nano-K MURAS2012), and

292

French ANR (ANR13-JS04-0001-01).

293

e-mail: gourdon@insp.jussieu.fr

294

[1] E. Beaurepaire, J. Merle, a. Daunois, and J. Bigot,

295

Phys. Rev. Lett. 76, 4250 (1996).

296

[2] M. van Kampen, C. Jozsa, J. Kohlhepp, P. LeClair,

297

L. Lagae, W. de Jonge, and B. Koopmans, Phys. Rev.

298

Lett. 88, 227201 (2002).

299

[3] J. Kisielewski, A. Kirilyuk, A. Stupakiewicz,

300

A. Maziewski, A. Kimel, T. Rasing, L. T.

301

Baczewski, and A. Wawro, Phys. Rev. B 85, 184429

302

(2012).

303

[4] D. Wang, Y. Ren, X. Liu, J. Furdyna, M.

Grims-304

ditch, and R. Merlin, Phys. Rev. B 75, 233308 (2007).

305

[5] Y. Hashimoto, S. Kobayashi, and H. Munekata,

306

Phys. Rev. Lett. 100, 067202 (2008).

307

[6] J. Qi, Y. Xu, a. Steigerwald, X. Liu, J. Furdyna,

308

I. Perakis, and N. Tolk, Phys. Rev. B 79, 085304

309

(2009).

310

[7] E. Rozkotova, P. Nemec, P. Horodyska, D. Sprinzl,

311

F. Trojanek, P. Maly, V. Novak, K. Olejnik,

312

M. Cukr, and T. Jungwirth, Appl. Phys. Lett. 92,

313

122507 (2008).

314

[8] S. Shihab, H. Riahi, L. Thevenard, H. J. von

315

Bardeleben, A. Lemaˆıtre, and C. Gourdon, Appl.

316

Phys. Lett. 106, 142408 (2015).

317

[9] P. Nˇemec, V. Nov´ak, N. Tesaˇrov´a, E.

Rozko-318

tov´a, H. Reichlov´a, D. Butkoviˇcov´a, F. Troj´anek,

319

K. Olejn´ık, P. Mal´y, R. P. Campion, B. L.

Gal-320

lagher, J. Sinova, and T. Jungwirth, Nat. commun.

321

4, 1422 (2013).

322

[10] F. Hansteen, A. Kimel, A. Kirilyuk, and T. Rasing,

323

Phys. Rev. Lett. 95, 1 (2005).

324

[11] A. Kirilyuk, A. V. Kimel, and T. Rasing, Rep. Prog.

325

Phys. 76, 026501 (2013).

326

[12] S. Mangin, M. Gottwald, C.-H. Lambert, D. Steil,

327

V. Uhl´ı, L. Pang, M. Hehn, S. Alebrand,

328

M. Cinchetti, G. Malinowski, Y. Fainman,

329

M. Aeschlimann, and E. E. Fullerton, Nat. Mat.

330

13, 286 (2014).

331

[13] Y. Kajiwara, K. Harii, S. Takahashi, J. Ohe,

332

K. Uchida, M. Mizuguchi, H. Umezawa, H. Kawai,

333

K. Ando, K. Takanashi, S. Maekawa, and E. Saitoh,

334

Nature 464, 262 (2010).

335

[14] B. Lenk, H. Ulrichs, F. Garbs, and M. Mnzenberg,

336

Phys. Rep. 507, 107 (2011).

337

[15] S. O. Demokritov, B. Hillebrands, and A. N.

338

Slavin, Phys. Rep. 348, 441 (2001).

339

[16] E. Rozkotova, P. Nemec, N. Tesarova, P. Maly,

340

V. Novak, K. Olejnik, M. Cukr, and T. Jungwirth,

341

Appl. Phys. Lett. 93, 232505 (2008).

342

[17] Y. Hashimoto and H. Munekata, Appl. Phys. Lett.

343

93, 202506 (2008).

344

[18] F. Teppe, M. Vladimirova, D. Scalbert,

345

M. Nawrocki, and J. Cibert, Sol. St. Com. 128, 403

346

(2003).

347

[19] N. Tesaˇrov´a, P. Nˆemec, E. Rozkotov´a, J. ˇSubrt,

348

H. Reichlov´a, D. Butkoviˆcov´a, F. Troj´anek,

349

P. Mal´y, V. Nov´ak, and T. Jungwirth, Appl. Phys.

350

Lett. 100, 102403 (2012).

351

[20] A. Kalashnikova, A. Kimel, R. Pisarev, V.

Grid-352

nev, P. Usachev, A. Kirilyuk, and T. Rasing, Phys.

353

Rev. B 78, 104301 (2008).

354

[21] Z. Liu, F. Giesen, X. Zhu, R. D. Sydora, and M. R.

355

Freeman, Phys. Rev. Lett. 98, 87201 (2007).

356

[22] M. Bombeck, A. S. Salasyuk, B. A. Glavin, A. V.

357

Scherbakov, C. Br¨uggemann, D. R. Yakovlev,V. F.

358

Sapega, X. Liu, J. K. Furdyna, A. V. Akimov,and M.

359

Bayer, Phys. Rev. B 85, 195324 (2012).

360

[23] J. Hamrle, J. Pi˘stora, B. Hillebrands, B. Lenk, and

361

M. M¨unzenberg, J. Phys. D: Appl. Phys. 43, 325004

362

(2010).

363

[24] C. Kittel, Phys. Rev. 110, 2139 (1958).

364

[25] J. Wu,N. D. Hughes,J. R. Moore,and R. J. Hicken

365

J. Magn. Magn. Mater. 241, 96 (2002).

366

[26] X. Liu and J. K. Furdyna, J.Phys.: Cond. Mat. 18,

367

R245 (2006).

368

[27] J. Zemen, J. Kuˇcera, K. Olejn´ık, and T. Jungwirth,

369

Phys. Rev. B 80, 1 (2009).

370

[28] M. Cubukcu, H. J. von Bardeleben, J. L. Cantin,

371

and A. Lemaˆıtre, Appl. Phys. Lett. 96, 102502 (2010).

372

[29] N. Tesaˇrov´a, D. Butkoviˇcov´a, R. P. Campion,

373

A. W. Rushforth, K. W. Edmonds, P. Wadley, B. L.

374

Gallagher, E. Schmoranzerov´a, F. Troj´anek,

375

P. Mal´y, P. Motloch, V. Nov´ak, T. Jungwirth, and

(7)

6

P. Nˇemec, Phys. Rev. B 90, 155203 (2014).

377

[30] S. Shihab, L. Thevenard, A. Lemaˆıtre, J.-Y.

378

Duquesne, and C. Gourdon, J. Appl. Phys. 119,

379

153904 (2016).

380

[31] A. V. Kimel, G. V. Astakhov, a. Kirilyuk, G. M.

381

Schott, G. Karczewski, W. Ossau, G. Schmidt,

382

L. W. Molenkamp, and T. Rasing, Phys. Rev. Lett.

383

94, 227203 (2005).

384

[32] N. Tesaˇrov´a, P. Nˇemec, E. Rozkotov´a, J.

Ze-385

men, T. Janda, D. Butkoviˇcov´a, F. Troj´anek,

386

K. Olejn´ık, V. Nov´ak, P. Mal´y, and T. Jungwirth,

387

Nat. Phot. 7, 492 (2013).

388

[33] G. Traeger, L. Wenzel, and a. Hubert, Phys. Stat.

389

Sol. (a) 131, 201 (1992).

390

[34] A. Hubert and G. Traeger, J. Mag. Magn. Mat. 124,

391

185 (1993).

392

[35] G. P´enissard, P. Meyer, J. Ferr´e, and D. Renard,

393

J. Mag. Magn. Mat. 146, 55 (1995).

394

[36] H. Terada, S. Ohya, and M. Tanaka, Appl. Phys.

395

Lett. 106, 222406 (2015).

396

[37] C. Gourdon, V. Jeudy, M. Menant, D. Roditchev,

397

L. A. Tu, E. L. Ivchenko, and G. Karczewski, Appl.

398

Phys. Lett. 82, 230 (2003).

399

[38] M. Inoue, K. Arai, T. Fujii, and M. Abe, J. Appl.

400

Phys. 83, 6768 (1998).

401

[39] I. L. Lyubchanskii, N. N. Dadoenkova, M. I.

402

Lyubchanskii, E. a. Shapovalov, and T. Rasing, J.

403

Phys. D: Appl. Phys. 36, R277 (2003).

Figure

FIG. 1. (a) Reference frame. (b) Dependence of the TRMO signal δβ on the probe beam linear polarization with respect to the sample crystallographic axes
FIG. 2. (left) Magnetization trajectory in the depth of the ferromagnetic layer for the p=0 (a), p=1 (b), p=2 (c) PSSW modes and corresponding detected trajectory in (g), (h), (i) assuming that the optical signal would result from a depth-averaged amplitud

Références

Documents relatifs

Admettons donc, comme cela est on ne peut plus logique, que les valeurs des or- ganes ont été déterminées avec discerne- ment et qu'elles sont bien adaptées à la rapidité

• Create a Crisis Prevention Navigator position to help Islanders identify and access government services and resources. • Increase the number of available

Le développement du modèle industriel de Capgemini dit de Rightshore ® , consistant à délocaliser une partie de la production d’une partie de ses services dans des centres

In February 2010, Abbott acquired Solvay’s pharmaceuticals business Solvay Pharmaceuticals for approximately $6.1 billion, in cash, plus additional payments of up to EUR 100 million

The fair value of the consolidated real estate portfolio, project developments included, amounted to 494 million euros at the end of 2010 compared to 538 million euros as at 31

Commentaire sur les participations en private equity à leur juste valeur En 2010, le secteur private equity a contribué au résultat du groupe à concurrence de 13,0 millions d’euros

Conclusion Nous vous demanderons de donner quitus entier et définitif à votre Conseil d’Administration de sa gestion pour l’exercice clos le 31 décembre 2011, ainsi

Chers Actionnaires, Nous vous avons réunis en Assemblée Générale, conformément à la loi et aux dispositions de nos statuts, à l’effet notamment de vous rendre compte des