• Aucun résultat trouvé

Resonant X-ray second harmonic generation in atomic gases

N/A
N/A
Protected

Academic year: 2021

Partager "Resonant X-ray second harmonic generation in atomic gases"

Copied!
26
0
0

Texte intégral

(1)

HAL Id: hal-02442214

https://hal.archives-ouvertes.fr/hal-02442214

Submitted on 16 Jan 2020

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.

Resonant X-ray second harmonic generation in atomic

gases

Ji-Cai Liu, Catalin Miron, Hans Ågren, Sergey Polyutov, Faris Gel’Mukhanov

To cite this version:

Ji-Cai Liu, Catalin Miron, Hans Ågren, Sergey Polyutov, Faris Gel’Mukhanov. Resonant X-ray second harmonic generation in atomic gases. Physical Review A, American Physical Society 2019, 100 (6), �10.1103/PhysRevA.100.063403�. �hal-02442214�

(2)

Resonant X-ray second harmonic generation in atomic gases

Ji-Cai Liu1,2∗, Catalin Miron3,4,5, Hans ˚Agren6,7,

Sergei Polyutov8,9, and Faris Gel’mukhanov6,8,9,10

1Department of Mathematics and Physics,

North China Electric Power University, 102206 Beijing, China

2Collaborative Innovation Center of Light Manipulations and Applications,

Shandong Normal University, Jinan 250358, China

3Synchrotron SOLEIL,

l’Orme des Merisiers, Saint-Aubin, BP 48,

91192 Gif-sur-Yvette Cedex, France

4Extreme Light Infrastructure Nuclear Physics (ELI-NP),

‘Horia Hulubei’ National Institute for Physics and Nuclear Engineering,

30 Reactorului Street, RO-077125 Mˇagurele,

Judet Ilfov, Romania

5LIDYL, CEA, CNRS,

Universit´e Paris-Saclay, CEA Saclay,

91191 Gif-sur-Yvette, France

6Theoretical Chemistry & Biology,

School of Biotechnology, Royal Institute of Technology,

S-106 91 Stockholm, Sweden

7Department of Physics and Astronomy,

UppsalaUniversity, SE-75120Uppsala, Sweden

8Siberian Federal University,

(3)

660041 Krasnoyarsk, Russia

9 Kirensky Institute of Physics,

Federal Research Center KSC SB RAS, 660036 Krasnoyarsk, Russia

10Institute for Methods and Instrumentation

in Synchrotron Radiation Research FG-ISRR,

Helmholtz-Zentrum Berlin f¨ur Materialien und Energie Albert-Einstein-Strasse 15,

12489 Berlin, Germany

Abstract

We explore the X-ray second harmonic generation process induced by resonant two-photon ab-sorption in systems with inversion symmetry. We show that this process becomes allowed in the X-ray region due to non-dipole contributions. It is found that while a plane wave pump field generates only a longitudinal second harmonic field, a Gaussian pump beam creates also a radially polarized transverse second harmonic field which is stronger than the longitudinal one. Contrary to the longitudinal component, the transverse second harmonic field with zero intensity on the axis of the pump beam can run in free space. Our theory is applied to Ar and Ne atomic vapours and predicts an energy conversion efficiency of X-ray SH generation of 3.2 × 10−11 and 1.3 × 10−12, respectively.

(4)

I. INTRODUCTION

Modern X-ray Free Electron Laser (XFEL) facilities can deliver high intensities 1015−1019

W/cm2 making it possible to significantly populate a core-excited state and even create

population inversion and X-ray lasing [1–7]. At these intensities, X-ray matter interaction becomes nonlinear creating room for studies of nonlinear effects like stimulated X-ray Raman scattering [3–5, 8–10], pulse compression [3–5], X-ray four-wave mixing [3–5, 11, 12], and nonlinear wave mixing of X-ray and near-infrared beams [13]. Special attention was paid to the competition between stimulated X-ray emission and Auger decay [14–16]. Second harmonic generation (SHG) is a nonlinear optical process of sum frequency generation which produces new photons with twice the frequency. SHG has traditionally been studied as an even-order nonlinear optical effect allowed in media without inversion symmetry[17] and is one of the best-understood nonlinear effects in optics[18]. In light of the XFEL development its study in the Angstrom regime, e.g., on the natural scale of atomic and molecular structure of matter, has become of great interest both from a fundamental and practical viewpoint. A pioneering study by Shwartz, Yudovich et al[19, 20] gave recently experimental evidence for ”off-resonant” SHG in diamond in the hard X-ray region with a pump frequency of 7.3 keV.

In the present work we show that due to the large momentum of the photon k the nonlinearity in the X-ray region is different from conventional nonlinearities in the visible regime and that SHG is generally possible to observe for centro-symmetric systems even when phase matching conditions do not prevail. We present a theoretical study of X-ray SHG in atomic gases induced by resonant two-photon absorption (TPA). We show that the plane wave pump field can create only longitudinally polarised X-ray second harmonic (SH) fields which can not propagate in free space (see however ref.[21]), but also that a Gaussian pump pulse induces in addition transverse second harmonic fields which contrary to the longitudinal component can run in free space. Our idea is, in a certain sense, inverse to the use of the longitudinal component of focused light beams in laser particle accelerators[22]. Another important feature of the SHG problem studied here is that the transverse field, being strictly equal to zero on the beam axis, has an unusual radial polarization.

Our work is organized as follows. We outline in Sec. II A the basic theory of SHG

(5)

the following Sec. II B we show that a Gaussian pump field creates also the transversely polarized SH field. Sec. II C is devoted to the analysis of the longitudinally and radially polarized SH fields. We shed light on the role of the absorption of an SH field in Sec. II D.

Some theoretical details can be found in Appendixes A, B and C. We discuss our results

further in Sec. III where we numerically analyze the efficiency of SHG in Ne and Ar vapours. Finally, in Sec. IV, we come to the conclusions.

II. THEORY

Quantum mechanically, the second order nonlinearity in the optical susceptibility

origi-nates from a perturbational solution of the Schr¨odinger’s equation. To get insight into the

physics of the SHG process in the X-ray region, we consider the propagation of an X-ray

pump field Ep in an atomic gas far away from the absorption edge. To induce the SHG

we choose twice the frequency of the pump field to be resonant with the frequency of a

two-photon transition 2ω ≈ ω10. The scheme of SHG is shown in Fig.1 where the pump

field resonantly promotes the 1s electron of atom to the np unoccupied level. The resonant population of the state |1i in the course of TPA is followed by the emission of the SH field E. Let us start from the atom-field interaction which reads as (we use SI units)

V = V(1)+ V(2) = − e 2mc(p · Ap+ Ap · p) + e2 2mc2A 2 p, (1)

where m and e are the mass and charge of the electron, respectively. c is the speed of light and p is the operator of electronic momentum. Below we will use more frequently the electric field instead of the vector potential, E = −∂A/∂t. The square of the vector potential of the

pump field A2p describes the TPA process in the first order of perturbation theory while the

scalar product p · Ap contributes to the TPA in second order of perturbation theory.

A. Plane wave pump field

It is instructive to consider first the interaction with the simplest and most fundamental

electromagnetic wave, the transverse plane wave Ap = (Ap/2) exp(−ı(ωt−k·r−k·r(e)))+c.c

Ep = 1 2Epe −ı(ωt−k·r−k·r(e)) + c.c., Ep = eEp(0) = − ∂Ap ∂t = ıωAp (2)

(6)

with the polarization e being orthogonal to the photon momentum k. Here r(e) is the

coordinate of the electron with respect to the atom and r is the radius vector of atom in the laboratory frame. To avoid unnecessary complexity (see also below) we will focus only on

the A2p term assuming that the wave length of the photon is longer than the size of the core

orbital ka  1 V(2) = e 2 8mc2A 2e−ı2(ωt−k·r(e)−k·r) + c.c. + const (3) ≈ e 2 8mc2A (0)2 p e −ı2(ωt−k·r) (1 + ı2k · r(e)) + c.c. + const.

The term A(0)2p being independent on the electron radius vector r(e)can not induce transitions

between electronic states. Thus the transition between the ground (s) and core-excited (p) states is induced solely by the the matrix element

V10(2) ≈ −ı eE

(0)2 p

4mc2ω2(k · d10)e

−ı[(2ω−ω10)t−2k·r] (4)

of the second term k · r(e)on the right-hand side of eq.(3). The rotating-wave approximation

is used here by keeping only the near resonant term. This pure non-dipole process opens

the s → p TPA channel with the transition dipole moment d01 = er01 = eh0|r(e)|1i (Fig.1).

We chose the axis z of quantization to lie along the photon momentum k. In this frame

the pump field populates only the npz level (see Fig.1) and the problem is reduced to the

interaction with a two-level atom with the transition dipole moment parallel to the photon momentum

d01 k k. (5)

The resonant TPA population of the core-excited state of p-symmetry is followed by the dipole allowed one-photon transition p → s which creates the SHG field with the double frequency 2ω. This explains why the SHG is possible in systems with inversion symmetry in the X-ray region.

To quantify the studied process one should compute the polarization P. The induced macroscopic polarisation of the medium being the expectation value of the dipole moment d is specified in terms of the density matrix ρ(t)

P = N Tr(dρ) = N (d01(t)ρ10(t) + d10(t)ρ01(t)) , (6)

where N is the concentration of atoms and d01(t) = d01exp(ıω10t) is the dipole moment

(7)

ρ10(t) = %10exp(−ı(νt − 2k · r)) satisfies the following kinetic equation in the interaction picture[23]  ∂ ∂t + Γ − ıν  %10= 1 ~m eEp(0) 2ω 2 (k · r10) (ρ11− ρ00) , (7)

where ν = 2ω − ω10 is the detuning from the two-photon resonance and Γ is the lifetime

broadening of core-excietd state |1i. We neglect the very weak depopulation of the ground

state in the course of the two-photon absorption, (ρ00 ≈ 1, ρ11  1) and assume that the

duration τ of the pump pulse is longer than the lifetime of the core excited state 1/Γ. In this case one can use the stationary solution of eq.(7)

%10= %∗01 = − 1 ~m eEp(0) 2ω !2 (k · r10) Γ − ıν (8)

to find the induced macroscopic polarization taking into account eq.(5)

P = P e−ı2(ωt−k·r)+ c.c., (9) P = ˆkp, p = − eE (0) p 2ω !2 N ker201 m~(Γ − ıν).

Here ˆk = k/k is the unit vector along k. One can see that the pump radiation creates a

macroscopic polarization P oriented along the direction of propagation of the pump field k, and hence the SH field E, which is created in the course of the spontaneous transition |1i → |0i, is also parallel to k. This longitudinal field exists everywhere where there is pump field and the medium and this field copies exactly the polarization according to Maxwell’s

equation for the induction, ∇ · D = ∂(0E + P)/∂z = 0:

E = −2

0

P 6= 0, D = 0, H = 0. (10)

This does not contradict the well known fact that the plane wave longitudinal field does not exist in free space[21]. This statement means that the longitudinal field can not propagate in free space. The longitudinal field E exists only in the region where the pump field creates longitudinal polarization P ∝ k. This longitudinal field oscillating in time and space is a pure electric field, H = 0.

As we have already noticed above the TPA process is a first-order process with respect

to A2p and a second-order process with respect to p · Ap. Here we study the two-photon

(8)

result in the same orientation of the TPA induced polarization we consider here only the

A2

p contribution. The taking into account of the p · Ap TPA process will result only in a

rescaling of the SHG efficiency.

B. Gaussian pump beam and paraxial equation

In this section we will show that a Gaussian pump beam

Ep = 1 2Epe −ı(ωt−kz)+ c.c., (11) Ep = ˆxEp(0)g  t − z c  w0 w(z)exp  − % 2 w2(z)  exp  ı  k % 2 2R(z) − ψ(z) 

makes it possible to transform the longitudinal SHG X-ray field into a transfer field which

can propagate in free space in contrast to the pure longitudinal field. Eq. (11) for pulsed

Gaussian beam is obtained in Appendix A by convoluting of the fundamental Gaussian mode with the Gaussian distribution of spectral components. As shown in Appendix A, Ep satisfies

paraxial equation. Eq.(11) identifies R(z) = z (1 + (zR/z)2) as the radius of curvature of the

wavefront of the beam at z, w0 as the beam waist and g(t) = exp(−t2/2τ2) as the temporal

shape of the pulse with duration τ . Here w(z) = w0p1 + (z/zR)2, ψ(z) = arctan(z/zR),

ρ =px2+ y2, w

0/zR∼ 1/kw0  1. The Gaussian beam remains well collimated up to the

Rayleigh range zR = kw20/2 (Fig.2).

Since the wavefront is not orthogonal to z, as one can see from the phase φ = 2k(z+ρ2/2R)

of Ep2 ∝ exp(ıφ), the polarization P is slightly tilted from the z-axis. To find the matrix

element V01(2) of the interaction with the Gaussian pump beam (11) we need the value of this

interaction at the point of the electron r(e) with respect to the atom r = (ρ, z), namely at

r + r(e),

h0|eı(φ+δφ)|1i ≈ eıφh0|1 + ıδφ|1i = κ · h0|r(e)|1ieıφ ı2k, (12)

where we used the Taylor expansion φ(r+r(e)) = φ(r)+δφ with δφ = ∇φ·r(e). Similar to the

derivation of eq.(9) one obtains a polarisation that is oriented along ∇φ ≡ (∂zφ, ∂ρφ) = 2kκ

P = −κ eEp 2ω 2 N ker2 01 m~(Γ − ıν), κ = ˆz + ˆρ ρ R (13)

instead of the beam axis ˆz k k.

Let us write the optical wave equation for the SHG field E

∇(∇ · E) − ∇2E + 1 c2 ∂2E ∂t2 = −µ0 ∂2P ∂t2 (14)

(9)

in the usual manner[18] starting from the couple of Maxwell’s equations (in SI units) for nonmagnetic materials (µ=1) ∇ × E = −µ0 ∂H ∂t , ∇ × H = ∂D ∂t . (15)

Contrary to conventional theories[18] where ∇ · E ∼ (k · e) = 0 for the transverse electro-magnetic field (k ⊥ e), we can not ignore ∇ · E here. This is because the polarization P is essentially a longitudinal one (see eq.(13)): ∇ · P 6= 0. To resolve this problem we use the

Maxwell’s equation for the induction D = ε0E + P

∇ · D = 0, ∇ · E = −1

0

∇ · P, (16)

which makes it possible to rewrite the wave equation (14) as follows

−∇2E + 1 c2 ∂2E ∂t2 = 1 0  −1 c2 ∂2P ∂t2 + ∇(∇ · P)  , (17) E = 1 2Ee −ı2(ωt−kz) + c.c., P = Pe−ı2(ωt−kz)+ c.c..

This wave equation differs from the conventional one [18] by the extra term ∇(∇ · P) 6= 0 which is not equal to zero because of the longitudinal contribution in P. We would like to point out that when the pump field is a plane wave there is only a longitudinal SH field

E k z (see Sec.II A). In this case the wave equation (14) becomes ∂2(

0E + P)/∂t2 = 0,

which is very different from eq.(17) because ∇(∇ · E) − ∇2E = (∂2/∂z2− ∂2/∂z2)E ≡ 0.

Now we are at the stage to simplify the wave equation (17). In our case the wave

propagates primarily along the z-axis with a small divergence angle (Fig. 2)

θ0 ≈ 1 2kw0 = w0 zR ∼ λ w0  1. (18)

Here λ is the wavelength of the pump field. We assume also that the pulse duration τ is

much longer than the period of field oscillations 2π/ω. This makes it possible to neglect

∂2E/∂z2 and ∂2E/∂t2 in eq. (17) (see Ref. [24]) and to get the following paraxial equation

for the SHG field  ∂ ∂z + 1 c ∂ ∂t− ı 4k∆⊥  E = ı 2kf , kw0  1, τ ω  1, (19)

where ∆⊥ = ∇2ρ= ∂2/∂x2+ ∂2/∂y2 is the Laplacian operator over the transverse cartesian

(10)

longitudinal (fz) and transverse components (fρ) ˜ f = ˆz ˜fz+ ˆρ ˜fρ, (20) ˜ f = 1 0 −1 c2 ∂2P˜ ∂t2 + ∇(∇ · ˜P) ! ≈ 1 0  (2k)2P + ∇(∇ · ˜˜ P)  , ˜ P = Pe−ı2(ωt−kz), ˜f = f e−ı2(ωt−kz)

Taking into account eqs.(11), (13) and (18) one can get the following expression for the transverse and longitudinal components of f

fρ = − ı8kρ 0w2 P, (21) fz = ı2kP 0(z2+ zR2) h2kρ2zR(ızR+ z) z2+ z2 R − ı4zR− z i .

One should point out that the origin of the fρ is the term ∇(∇ · ˜P) = ˆρ∂ρ(∂zP ) + · · · . A˜

simple estimation shows that the transverse contribution dominates: |fρ/fz| ∼ kw0  1.

As one can see from the paraxial equation (19) the transverse and longitudinal components of f generate, respectively, the transverse and longitudinal components of the SH field E.

C. Spatial distribution of the transverse and longitudinal SH fields. Radial polar-ization

It is convenient to write the solution of the paraxial equation (19) in terms of the retarded Green’s function (see Appendix B)

E(z, ρ, t) = 1 2π Z G(z − z0, ρ − ρ0, t − t0)f (z0, ρ0, t0)dz0dρ0dt0, (22)  ∂ ∂z + 1 c ∂ ∂t − ı 4k∆⊥  G(z − z0, ρ − ρ0, t − t0) = δ(z − z0)δ(ρ − ρ0)δ(t − t0)Θ(t − t0). G(z − z0, ρ − ρ0, t − t0) = −ıδ  t0− t − z 0− z c )  Θ(t − t0) k π(z − z0)exp  ık|ρ − ρ 0)2 z − z0  , which guarantees that no contribution at remotely early times, t, before the source ˜

f (z0, ρ0, t0) = f (z0, ρ0, t0) exp(−ı2(ωt0 − kz0)) has been activated. Taking into account that

ˆ

(11)

the plane (ˆρ, ˆy) orthogonal to the z-axis using eq.(B5) 2π Z 0 dϕ[ˆρ0fρ(z0, ρ0, t0) + ˆzfz(z0, ρ0, t0)] exp  ık|ρ − ρ 0)2 z − z0  (23) = 2π exp  ık(ρ 2+ ρ02) z − z0  h − ˆρıfρ(z0, ρ0, t0)J1  2kρρ0 z − z0  + ˆzfz(z0, ρ0, t0)J0  2kρρ0 z − z0  i ,

where Jn(x) is a Bessel function. One can obtain the remaining integral over ρ0 using eq.(B5)

and get the following expressions for transverse and longitudinal contributions

E(z, ρ, t) = ˆρEρ(z, ρ, t) + ˆzEz(z, ρ, t), (24) Ei(z, ρ, t) = Ep(0)g 2 t − z c  Ji(z, ρ), i = (ρ, z), where Jρ(z, ρ) = −2ρs0 z Z −∞ dz0 e Φ w4(z02(z0), (25) Jz(z, ρ) = ı4πw0s0 (kw0)3 z Z −∞ dz0 e Φ w4(z0)α(z0)  2(ızR+ z0) w2(z0)α(z0)  z − z0− 2k 2ρ2 α(z0)  − (ı4zR+ z0)  , Φ = ıkρ 2 z − z0 − ı2ψ(z 0 ) − k 2ρ2 (z − z0)α(z0), α(z 0 ) = 2(z − z 0) w2(z0) − ık  z − z0 R(z0) + 1  , s0 = 8π G Γ − ıνN zRr01re

Here re = e2/(4π0mc2) = 2.82 × 10−13cm is the classical electron radius and G = E

(0)

p d01/~

is the Rabi frequency. It is important to notice that there is no transverse field on the beam axis

Eρ(z, ρ = 0, t) = 0. (26)

Eq.(24) indicates that the transverse SH field ˆρEρ is oriented along the radius ρ

perpendic-ular to the beam axis (Fig. 3). This means that the transverse field has radial polarization (see also Sec. III).

D. Role of photoabsorption

In the equations above the photoabsorpion of X-rays is ignored. This approximation is valid for the pump beam whose frequency is far from any resonance. In contrast, the SHG

(12)

field is in strict resonance with the dipole allowed transition |0i → |1i (1s − 3p for Ne and 1s − 4p for Ar). Therefore, this absorption channel should be taken into account. With the solution (24) at hand we are almost prepared to include the photoabsorption in the SH field. As shown in Appendix C the photoabsorption of the SHG field modifies only the integrands

at the right-hand side of equations (25) for Jρ(z, ρ) and Jz(z, ρ). Namely, these integrands

should be multiplied by the factor

exp z

0− z

2` 

, (27)

where ` = 1/N σabs is the photoabsorption length while σabs is the resonant photoabsorption

cross section. According to simulations the photoabsorption length should be larger or comparable with the Rayleigh range

` ∼ zR (28)

to make it possible for the SHG field to reach the optimal value.

III. RESULTS OF SIMULATIONS AND DISCUSSION

We applied the theory outlined above to two atomic systems, Ne and Ar, under the strict

two-photon resonance (2ω = ω10) with 1s → 3p transitions for Ne and 1s → 4p transitions

for Ar. In the simulations, the peak pump intensity used was Ip(0) = c0|E

(0)

p |2/2 = 1016

W/cm2 (Fig. 2), and the following parameters were adopted for Ne: ~ω

1s−3p = 867.4 eV,

σabs(1s − 3p) = 1.5 × 10−18 cm2[25], 2~Γ = 0.27 eV [26], and for Ar: ~ω1s−4p= 3203.42 eV,

σabs(1s − 4p) = 0.12 × 10−18 cm2[27], 2~Γ = 0.66 eV[28]. The concentration of the atoms and

the beam waist were equal to N = 1019 cm−3 and w

0 = 1.0µm, respectively. The Rayleigh

range was zR≈ 103µm and zR ≈ 4 × 103µm for Ne and Ar, respectively. The corresponding

values of the photoabsorption lengths ` ≈ 0.67 × 103µm and 8 × 103µm satisfy the condition

(28).

We solved the paraxial equation with homogeneous distribution of the concentration. The

SHG radiation is characterized by the intensity distributions of the transverse (Iρ(z, ρ, t))

and longitudinal (Iz(z, ρ, t)) components of the SH field (24)

Ii(z, ρ, t) =

1

2c0|Ei(z, ρ, t)|

(13)

and by the energy conversion efficiency βi = Wi(z) Wp , Wi(z) = 2π Z ∞ 0 dt Z 2π 0 dρIi(z, ρ, t). (30)

First we studied the SH field neglecting the photoabsorption. In Fig. 4 and Fig. 5 we

display the spatial distribution of the SH intensities Iρ(z, ρ, t) and Iz(z, ρ, t) for Ne and Ar,

respectively. One can see that the transverse and longitudinal SH fields show very different

radial structures with Iρ= 0 on the axis of the beam ρ = 0, and the transverse field ˆρEρ(24)

has an unusual radial polarization as shown in Fig. 3. The transverse SH field Iρ is much

stronger than the longitudinal one Iz, and the energy conversion efficiency of the transverse

SH field is about four to five orders of magnitude larger than that of the longitudinal SH field as shown in Fig. 6.

Because of the deeper ionization potential and smaller core-electron transition dipole moment, the conversion efficiencies of the X-ray SH fields from Ar atomic vapour are much smaller than those from Ne (Fig. 6). However, when the photoabsorption of the generated SH fields is considered, the final conversion efficiencies from Ar and Ne become comparable. Below we investigated the conditions for the experimental observation of the SHG process with X-rays in atomic Ne and Ar vapours by taking into account the resonant one-photon absorption of the SH field during propagation. The photoabsorption changes the spatial distribution of the SH field (Fig. 7) and reduces the energy conversion efficiency in one order of magnitude for Ne and in four times for Ar, as one can see from Fig. 6 and Fig.8. As it is expected the SH field is confined in the focal region in the range limited by the photoabsorption length ` (see Fig. 7). Due to this circumstance the energy conversion

efficiency becomes maximal at z = zmax = 0.7 mm for Ne and zmax = 0.5 cm for Ar. This

range defines the size of the gas cell which should be around zmax.

IV. SUMMARY

In this paper, we investigated the second harmonic generation in systems with inversion symmetry in the X-ray region. Our theory is applied to SHG in neon and argon pumped by a strong X-ray field tuned in resonance with the two-photon transition 1s → 3p in Ne and 1s → 4p in Ar. The non-dipole population of these core-excited states is followed by the emission of the SH field. We describe the SHG in atoms in terms of a density

(14)

matrix formalism and paraxial equation taking into account the resonant photoabsorption of the SH radiation. In contrast to the plane wave pump field, the Gaussian pump beam generates transverse SH photons with radial polarization. By taking into account the X-ray photoabsorption effect, the energy conversion efficiencies to the transverse SH fields are

expected to be orders of 10−11and 10−12in Ne (867.4 eV) and Ar (3203.4 eV) atomic vapours

for the pump 1016 W/cm2, respectively.

Acknowledgements. The research leading to these results has received funding from a public grant from the Laboratoire d’Excellence Physics Atoms Light Matter (LabEx PALM), the Laboratoire d’Excellence MiChEm, and the Swedish Research Council (VR). JL thanks the support by the National Science Foundation of China under grant Nos. 11574082 and 11974108, and the Fundamental Research Funds for the Central Universities under grant No. 2018MS050. FG and SP acknowledge support within the State contract of the Russian Federation Ministry of Education and Science for Siberian Federal University for scientific

research in 2017-2019 (Project 3.2662.2017). H˚A acknowledges the Knut and Alice

Wallen-berg foundation for financial support (Grant No. KAW-2013.0020).

Appendix A: The pulsed Gaussian beam (11) is eigen function of paraxial operator

To get pulsed Gaussian beam (11) with the carrier frequency ω we need to convolute fundamental Gaussian mode

E(ω, t) = 1 2E(ω)e −ı(ωt−kz) + c.c., (A1) E(ω) = ˆxEp(0) w0 w(z) exp  − % 2 w2(z)  exp  ı  k % 2 2R(z) − ψ(z)  ,

with the spectral distribution g(ω0− ω) = exp(−(ω0− ω)2τ2/2)/τ2π centered at the carrier

frequency ω Ep(ω, t) = ∞ Z −∞ g(ω0 − ω)E(ω0, t)dω0. (A2)

The mode E(ω) is the eigen function of the stationary paraxial or Helmholtz equation  ∂ ∂z − ı 2k∆⊥  E(ω) = 0. (A3)

(15)

The substitution of the fundamental mode (A1) in the convolution (A2) results in an ex-pression Ep(ω, t) = ˆ x 2E (0) p ∞ Z −∞ dω0g(ω0− ω) × exp  −ıω0  t − z c − ρ2 2cR(z)  w0 w(z)exp  − % 2 w2(z)  e−ıψ(z)+ c.c. (A4)

We neglect the ω0-dependence of the Rayleigh range zR = k0w02/2 ≈ kw02/2 because the

variation of the frequency ∆ω = |ω0− ω| ∼ 1/τ in the Fourier transform (A4) is negligibly small in comparison with the carrier frequency of X-ray pulse: ∆ω/ω ∼ 1/τ ω  1. Thus

Ep(ω, t) ≈ ˆ x 2E (0) p exp  −ıω  t − z c − ρ2 2cR(z)  w0 w(z)exp  − % 2 w2(z)  e−ıψ(z) ×g  t − z c − ρ2 2cR(z)  + c.c., g(t) = exp  − t 2 2τ2  (A5)

Within paraxial approximation [24] (kw0  1) we can neglect ρ2/2cR(z) in the Gaussian

g (t − z/c − ρ2/2cR(z))) because z ∼ zR∼ kw02  ρ2 2R(z) ∼ w2 0 zR ∼ 1 k, ρ2 2cR(z)τ ∼ 1 τ ω  1. (A6) However, we should keep ρ2/2cR(z) in the oscillatory term exp[−ıω(t − z/c − ρ2/2cR(z))]

because ω ρ 2 2cR(z) = kρ2 2R(z) ≈ kw2 0 2zR ∼ 1. (A7)

Finally we get eq.(11) for Ep(ω, t).

Now we are in stage to show that Ep from eq.(11) satisfies the paraxial equation

 ∂ ∂z + 1 c ∂ ∂t − ı 2k∆⊥  Ep= 0. (A8)

Let us apply the operator  = −∇2+ ∂2/c22t to both sides of eq.(A2)

Ep(ω, t) = ∞

Z

−∞

g(ω0 − ω)E(ω0, t)dω0. (A9) Using the paraxial approximation, kw0  1, τ ω  1 and eq.(A3) we get eq.(A8)

 ∂ ∂z + 1 c ∂ ∂t − ı 2k∆⊥  Ep = eı(ωt−kz) (A10) × ∞ Z −∞ dω0g(ω0− ω)k 0 ke −ı(ω0t−k0z) ∂ ∂z − ı 2k0∆⊥  E(ω0) = 0. (A11)

(16)

The paraxial approximation is broken when kw0 <∼ 1, τ ω ∼ 1. In this case one should<

restore in eq. (19) the second derivatives over z and time: ∂/∂z → ∂/∂z − (ı/4k)∂2/∂z2,

∆⊥ → ∆⊥ − ∂2/c2∂t2. However, such conditions are difficult to reach in X-ray region.

For example, the condition τ ω ∼ 1 corresponds to few cycle X-ray pulse, where the pulse duration is comparable with the period of the field oscillations.

Appendix B: Green’s function for the time-dependent paraxial equation

Let us find the Green’s function of non-stationary paraxial equation  ∂ ∂z + 1 c ∂ ∂t− ı 2K∆⊥  G(z, ρ, t) = Θ(t)δ(t)δ(z)δ(ρ), (B1)

where ∆⊥ = ∂2/∂x2 + ∂2/∂y2, ρ = (x, y), δ(ρ) = δ(x)δ(y) and Θ(t) is the step function

which is equal to zero when t < 0. Taking the Fourier transform of the Green’s function and of the Dirac δ-functions we get

G(z, ρ, t) = Θ(t) (2π)4 ∞ Z −∞ dµ ∞ Z −∞ dν ∞ Z −∞ dp ∞ Z −∞ dqGµ,ν,p,qeıµt+ıνz+ıpx+ıqy, Gµ,ν,p,q= − 2ı µ c + ν + p2+q2 2K . (B2)

Keeping in mind that t ≥ 0 and taking the integral along half circle in upper half plane

around the pole µ = −cν + p22K+q2

∞ Z −∞ eıµt µ c + ν + p2+q2 2K dµ = −ıcπ exp  −ıct  ν + p 2+ q2 2K  , (B3)

we obtain the following expression for the Green’s function G(z, ρ, t) = −ıδ  t − z c  Θ(t) K 2πzexp  ıKρ 2 2z  (B4) which allows to find the SHG field (22) with help of the following integrals[29]

J0(a) = 1 2π Z 2π 0 eıa cos θdθ, 1 2π Z 2π 0

cos θe−ıa cos θdθ = ı d

daJ0(a) = −ıJ1(a), (B5) ∞ Z 0 e−a2ρ2ρn+1Jn(bρ)dρ = bn (2a2)n+1e −b2 4a2, Re(a2) > 0.

(17)

Appendix C: Photoabsorption of SHG field

The strongest absorption channel is the absorption of SH field which is in resonance with |0i → |1i transition. To take into account this photoabsorption we need to add −E/2` at the right-hand side of paraxial equation (19)

 ∂ ∂z + 1 c ∂ ∂t− ı 4k∆⊥  E = −1 2`E + ı 2kf , (C1)

where ` = 1/σN is the length of resonant absorption of the SHG field with the

photoabsorp-tion cross secphotoabsorp-tion σ. Using the substituphotoabsorp-tion E = ˜E exp(−z/2`) one can see that ˜E satisfies

paraxial equation (19)  ∂ ∂z + 1 c ∂ ∂t− ı 4k∆⊥  ˜ E = ı 2kf e z/2` (C2)

with modified source term. This equation has the solution given by eq.(22) with f replaced by

f exp(z0/2`). Taking into account this we get immediately the solution of paraxial equation

with photoabsorption (C1) E(z, ρ, t) = ˜Ee−z/2` = e −z/2` 2π Z G(z − z0, ρ − ρ0, t − t0)f (z0, ρ0, t0)ez0/2`dz0dρ0dt0. (C3)

This means that to include the photoabsorption we should multiply by exp((z0− z)/2`) the

integrand at the right-hand side of equations (25) for Jρ(z, ρ) and Jz(z, ρ).

[1] N. Rohringer and R. London, Atomic inner-shell x-ray laser pumped by an x-ray free-electron laser, Phys. Rev. A 80, 013809 (2009).

[2] N. Rohringer, D. Ryan, R. London, M. Purvis, F. Albert, J. Dunn, J. D. Bozek, C. Bostedt, A. Graf, R. Hill, S. P. Hau-Riege, and J. J. Rocca, Atomic inner-shell X-ray laser at 1.46 nanometres pumped by an X-ray free-electron laser, Nature 481, 488-491 (2012).

[3] Y.-P. Sun, J.-C. Liu, and F. Gel’mukhanov, Slowdown and compression of a strong X-ray free-electron pulse propagating through the Mg vapors, EuroPhys. Lett. 87, 64002 (2009). [4] Y.-P. Sun, J.-C. Liu, and F. Gel’mukhanov, The propagation of a strong x-ray pulse followed by

pulse slowdown and compression, amplified spontaneous emission and lasing without inversion, J. Phys. B: At. Mol. Opt. Phys. 42, 201001 (2009).

(18)

[5] Y.-P. Sun, J.-C. Liu, C.-K. Wang, and F. Gel’mukhanov, Propagation of a strong x-ray pulse: Pulse compression, stimulated Raman scattering, amplified spontaneous emission, lasing with-out inversion, and four-wave mixing, Phys. Rev. A 81, 013812 (2010).

[6] Q. Miao, J.-C. Liu, H. ˚Agren, J-E. Rubensson, and F. Gel’mukhanov, Dissociative X-ray Lasing, Phys. Rev. Lett. 109, 233905 (2012).

[7] V. Kimberg and N. Rohringer, Amplified X-ray emission from core-ionized diatomic molecules, Phys. Rev. Lett. 110, 043901 (2013).

[8] V. Kimberg and N. Rohringer, Stochastic stimulated electronic X-ray Raman spectroscopy, Struct. Dynam. 3, 034101 (2016).

[9] V. Kimberg, A. Sanchez-Gonzalez, L. Mercadier, C. Weninger, A. Lutman, D. Ratner, R. Coffee, M. Bucher, M. Mucke, M. Ag˚aker, C. S˚athe, C. Bostedt, J. Nordgren, J.-E. Rubensson, and N. Rohringer, Stimulated X-ray Raman scattering: a critical assessment of the building block of nonlinear X-ray spectroscopy, Faraday Discuss. 194, 305-324, (2016).

[10] M. Beye, S. Schreck, F. Sorgenfrei, C. Trabant, N. Pontius, S. Sch¨ußler-Langeheine, W. Wurth, and A. F¨ohlisch, Stimulated X-ray emission for materials science, Nature 501, 191-194 (2013) [11] S. Tanaka and S. Mukamel, Coherent X-Ray Raman spectroscopy: A nonlinear local probe

for electronic excitations, Phys. Rev. Lett. 89, 043001 (2002).

[12] S. Mukamel, Multiple core-hole coherence in x-ray four-wave-mixing spectroscopies, Phys. Rev. B 72, 235110 (2005).

[13] T. E. Glover, D. M. Fritz, M. Cammarata, T. K. Allison, S. Coh, J. M. Feldkamp, H. Lemke, D. Zhu, Y. Feng, R. N. Coffee, M. Fuchs, S. Ghimire, J. Chen, S. Shwartz, D. A. Reis, S. E. Harris, and J. B. Hastings, X-ray and optical wave mixing, Nature 488, 603 (2012).

[14] N. Rohringer and R. Santra, Resonant Auger effect at high x-ray intensity, Phys. Rev. A 77, 053404 (2008).

[15] J.-C. Liu, Y.-P. Sun, C.-K. Wang, H. ˚Agren, and F. Gel’mukhanov, Auger effect in the presence of strong x-ray pulses, Phys. Rev. A 81, 043412 (2010).

[16] Y.-P. Sun, Q. Miao, A.-P. Zhou, R.-J. Liu, B. Liu, and F. Gel’mukhanov, Suppression of resonant auger effect with chirped x-ray free-electron laser pulse, J. Phys. B: At. Mol. Opt. Phys. 51, 035602 (2018).

[17] P. Franken, A. Hill, C. Peters, and G. Weinreich, Generation of optical harmonics, Phys. Rev. Lett. 7, 118-119 (1961).

(19)

[18] R. W. Boyd, Nonlinear Optics, Academic Press, Elsevier, Singapore, 2010.

[19] S. Shwartz, M. Fuchs, J. B. Hastings, Y. Inubushi, T. Ishikawa, T. Katayama, D. A. Reis, T. Sato, K. Tono, M. Yabashi, S. Yudovich, and S. E. Harris, X-Ray second harmonic generation, Phys. Rev. Lett. 112, 163901 (2014).

[20] S. Yudovich and S. Shwartz, Second-harmonic generation of focused ultrashort x-ray pulses, J. Opt. Soc. Am. B 32, 1894-1900 (2015).

[21] T. F. Heinz and D. P. DiVincenzo, Comment on ”Forbidden nature of multipolar contributions to second-harmonic generation in isotropic fluids”, Phys. Rev. A 42, 6249-6251 (1990). [22] M. O. Scully and M. S. Zubairy, Simple laser accelerator: Optics and particle dynamics, Phys.

Rev. A 44, 2656-2663 (1991).

[23] P. W. Milonni and J. H. Eberly, Lasers Physics, John Wiley & Sons, New York, 2010. [24] L. Mandel and E. Wolf, Optical Coherence and Quantum Optics, Cambridge University Press,

New York, 1995, p. 271.

[25] C. Buth, R. Santra, and L. Young, Electromagnetically induced transparency for X-rays, Phys. Rev. Lett. 98, 253001 (2007).

[26] V. Schmidt, Electron spectrometry of atoms using synchrotron radiation, Cambridge University Press, Cambridge, England, 1997.

[27] C. Buth and R. Santra, X-ray refractive index of laser-dressed atoms, Phys. Rev. A 78, 043409 (2008).

[28] J. L. Campbell and T. Papp, Widths of the atomic K-N7 levels, At. Data Nucl. Data Tables 77, 1-56 (2001).

[29] M. Abramowitz and I. A. Stegun, eds., Handbook of mathematical functions, Dover, New York, 1965.

(20)

m=-1

0

1

1 = n p

0 = 1s

!

r

10

 ↑↑ z ↑↑ 

k

!

FIG. 1: The dipole moment r10 of the 1s → np transition in atom is parallel to k. The axis of

(21)

- 4

- 2

0

2

4

0

1 0

2 0

3 0

4 0

5 0

6 0

7 0

8 0

1 0 0 5 . 6 2 E 3 3 . 1 6 E 5 1 . 7 8 E 7 1 . 0 0 E 9 5 . 6 2 E 1 0 3 . 1 6 E 1 2 1 . 7 8 E 1 4 1 . 0 0 E 1 6



(

µ

m

)

N e

I

p

( z ,

)

0

= w

0

/ z

R

~ 1 0

- 3

- 1 5

- 1 0

- 5

0

5

1 0

1 5

0

1 0

2 0

3 0

4 0

5 0

6 0

7 0

8 0

1 0 0 5 . 6 2 E 3 3 . 1 6 E 5 1 . 7 8 E 7 1 . 0 0 E 9 5 . 6 2 E 1 0 3 . 1 6 E 1 2 1 . 7 8 E 1 4 1 . 0 0 E 1 6



(

µ

m

)

z ( c m )

A r

0

= w

0

/ z

R

~ 0 . 2 5

×

1 0

- 3

FIG. 2: 2D map of the pump intensity at t = z/c for Ne and Ar atomic vapours. The legend shows the intensity in W/cm2.

(22)

0 5 1 0 1 5 2 0 2 5 3 0 3 5 4 0 0 5 1 0 1 5 2 0 2 5 3 0 3 5 4 0



(

µ

m

)

0 . 1 0 0 0 . 4 2 2 1 . 7 8 7 . 5 0 3 1 . 6 1 3 3 5 6 2 2 . 3 7 E 3 1 . 0 0 E 4

FIG. 3: Radial distribution of Iρfor Ne atomic vapour at z = 0.02 m. Black arrows show

(23)

0

1 0

2 0

3 0

4 0

5 0

6 0

7 0

8 0



(

µ

m

)

0 . 1 0 0 0 . 7 9 4 6 . 3 0 5 0 . 1 3 9 7 3 . 1 6 E 3 2 . 5 1 E 4 1 . 9 9 E 5 1 . 5 8 E 6

I

( z ,

)

N e

- 4

- 2

0

2

4

0

1 0

2 0

3 0

4 0

5 0

6 0

7 0

8 0

1 . 0 0 E - 6 1 . 0 8 E - 5 1 . 1 6 E - 4 0 . 0 0 1 2 5 0 . 0 1 3 5 0 . 1 4 6 1 . 5 7 1 7 . 0 1 8 3



(

µ

m

)

z ( c m )

I

z

( z ,

)

FIG. 4: 2D map of the SHG intensity at t = z/c of Ne atomic vapour. The photoabsorption is negleted. The legend shows the intensity in W/cm2.

(24)

0

1 0

2 0

3 0

4 0

5 0

6 0

7 0

8 0

1 . 0 0 E - 3 0 . 0 1 0 6 0 . 1 1 2 1 . 1 9 1 2 . 6 1 3 3 1 . 4 1 E 3 1 . 4 9 E 4 1 . 5 8 E 5



(

µ

m

)

A r

I

( z ,

)

- 1 5

- 1 0

- 5

0

5

1 0

1 5

0

1 0

2 0

3 0

4 0

5 0

6 0

7 0

8 0

1 . 0 0 E - 9 1 . 4 4 E - 8 2 . 0 7 E - 7 2 . 9 7 E - 6 4 . 2 8 E - 5 6 . 1 5 E - 4 0 . 0 0 8 8 5 0 . 1 2 7 1 . 8 3



(

µ

m

)

I

z

( z ,

)

z ( c m )

FIG. 5: 2D map of the SHG intensity at t = z/c of Ar atomic vapour. The photoabsorption is neglected. The legend shows the intensity in W/cm2.

(25)

0 1 2 3 4 - 2 - 1 0 1 2 3 4 5 0 . 0 0 . 5 1 . 0 1 . 5 2 . 0 0 1 2 3 4 5 - 5 0 5 1 0 1 5 0 . 0 0 . 3 0 . 6 0 . 9 1 . 2 1 . 5 10 10 (z ) N e z ( c m ) 10 14 z (z ) A r 10 12 (z ) 10 17 z (z ) z ( c m )

FIG. 6: Energy conversion efficiencies of the transverse SH fields βρ(z) and the longitudinal SH

fields βz(z) in Ne and Ar atomic vapours. The photoabsorption is neglected. The vertical axises

show the energy conversion efficiencies multiplied by the factor 10n: For example, 1010βρ and

(26)

- 1 5 - 1 0 - 5 0 5 1 0 1 5 0 5 1 0 1 5 2 0

A r

0 . 0 1 0 0 0 . 0 3 9 2 0 . 1 5 4 0 . 6 0 2 2 . 3 6 9 . 2 5 3 6 . 3 1 4 2 5 5 7 2 . 1 8 E 3 8 . 5 6 E 3



(

µ

m

)

z ( c m )

- 4 - 2 0 2 4 0 5 1 0 1 5 2 0 0 . 0 1 0 0 0 . 0 5 5 5 0 . 3 0 8 1 . 7 1 9 . 4 7 5 2 . 5 2 9 1 1 . 6 2 E 3 8 . 9 7 E 3 4 . 9 8 E 4 2 . 7 6 E 5



(

µ

m

)

z ( c m )

N e

FIG. 7: Distribution of the transverse SH field Iρ for Ne and Ar by taking into account the

photoabsorption. The legend shows the intensity in W/cm2.

- 0 . 4 - 0 . 2 0 . 0 0 . 2 0 . 4 0 . 6 0 . 8 1 . 0 0 . 0 0 . 5 1 . 0 1 . 5 2 . 0 2 . 5 3 . 0 3 . 5 - 4 - 2 0 2 4 6 8 1 0 0 . 0 0 . 2 0 . 4 0 . 6 0 . 8 1 . 0 1 . 2 1 . 4 N e z ( c m ) 10 11 (z ) A r 10 12 (z ) z ( c m )

FIG. 8: Energy conversion efficiency βρ(z) for Ne and Ar by taking into account the

photoabsorp-tion. The red dashed lines display βρ(z) calculated by neglecting the photoabsorption (see Fig. 6).

Figure

FIG. 1: The dipole moment r 10 of the 1s → np transition in atom is parallel to k. The axis of quantization z is along the photon momentum k.
FIG. 2: 2D map of the pump intensity at t = z/c for Ne and Ar atomic vapours. The legend shows the intensity in W/cm 2 .
FIG. 3: Radial distribution of I ρ for Ne atomic vapour at z = 0.02 m. Black arrows show schemat- schemat-ically the radially polarized SHG field.
FIG. 4: 2D map of the SHG intensity at t = z/c of Ne atomic vapour. The photoabsorption is negleted
+4

Références

Documents relatifs

As this method aims at designing nanostructures able to greatly enhance the nonlinear processes, a single grating of cylindrical nanorods is studied. The behavior of this architec-

The linear dependence of the signal on the applied field indicates that these longer pulses could still be useful in applications where the timescale of the electric field variation

The buildup of the space charge field then causes a self-focusing effect that efficiently traps the beam; at the same time the initial negative phase mismatch is counterbalanced by

To verify whether director reorientation occurs m the elastomer the behaviour of the second harmonic as a function of the incident laser power respectively as a function of the

Moreover, by analyzing similar wave-breaking dynamics in a PPLN crystal in the presence of a higher nonlinear quadratic response, we observe that the spatial beam breaking may

In particular, collagen has been shown to exhibit endogenous Second Harmonic Generation (SHG) signals and SHG microscopy has proved to enable the visualization of collagen

Ces dispo- sitifs ont pour originalité de s’appuyer sur une collaboration avec le monde professionnel : l’instauration des jurys articulés au stage de fin de premier cycle et

The virial expansion for cold two-component Fermi and Bose atomic gases is considered in the presence of a waveguide and in the vicinity of a Feshbach resonance.. The