You are on page 1of 130

UNIVERSITY OF BUCHAREST

Generalization of space-translation
method for atom laser interaction in
relativistic regime

Madalina Boca

thesis supervisor

Professor Viorica Florescu

2007
Reality is that which, when you stop
believing in it, doesn’t go away.
Philip K. Dick
Acknowledgments

It is my pleasure to take this opportunity to thank a number of people who, in a way


or the other, contributed to the completion of this thesis.

Firstly, I would like to thank my supervisor, professor Viorica Florescu. Her ideas and
tremendous support made this thesis possible. I am also grateful for excellent courses
in Theoretical Physics she taught, for giving me the opportunity to work in her research
group, for helping me to develop as a researcher myself, for the long discussions from
which I have learned so much, for her constant trust and encouragement.

Very special thanks to Professor Mihai Gavrila for suggesting the subject of this Thesis,
and for his enthusiastic support and guidance during all these years.

I am deeply indebted to Professor Harm Geert Muller for sharing with me his vast
experience in Computational Physics and for his help in developing the numerical codes
on which the last part of the Thesis is based.

I also wish to express my gratitude to Professor Mihai Dondera for numerous discussions
related to various scientific problems, for the time he spent in building and maintaining
the computer network I used, for many advices he gave me, for his constant support.

The numerical applications presented in this Thesis were realized in the Computing
Laboratory of Centre for Advanced Quantum Physics.

I wish to express my appreciation of the stimulating atmosphere created by my profes-


sors and colleagues from the Centre for Advanced Quantum Physics.

Last, but not least, I wish to thank all members of the Faculty of Physics who guided
my steps during my undergraduate and master studies.
Contents

List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii


List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v

1 Introduction 1
1.1 Theoretical methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Numerical simulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.4 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

2 Atomic stabilization: the non-relativistic case 11


2.1 Quasistationary stabilization . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.1.1 Numerical examples . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.2 Dynamic stabilization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.2.1 Numerical example . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.3 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

3 The Volkov solutions 33


3.1 Calculation of Volkov solutions . . . . . . . . . . . . . . . . . . . . . . . . 34
3.2 Properties of Volkov solutions . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.2.1 Proof of the orthogonality of the Volkov solutions . . . . . . . . . . 36
3.2.2 Proof of the completeness of the Volkov solutions . . . . . . . . . . 38
3.3 Other Volkov solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.4 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

4 Relativistic generalization of the space translation method 43


4.1 The general framework . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.2 Properties of the transformation operator . . . . . . . . . . . . . . . . . . . 45
4.3 The atomic low momentum regime approximation of T . . . . . . . . . . . 49
4.4 The low momentum regime approximation of the transformed potential V 0 56
4.5 Position and velocity operators in the generalized translated Dirac picture 61
4.6 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

5 The generalized translated Schrödinger equation 63


5.1 Derivation and properties of the generalized translated Schrödinger equation 63
5.2 The generalized translated potential . . . . . . . . . . . . . . . . . . . . . . 65

i
5.3 Equivalent forms of the generalized translated Schrödinger equation . . . . 68
5.3.1 The generalized Schrödinger equation in the “R” picture . . . . . . 68
5.3.2 The generalized Schrödinger equation in the “L” picture . . . . . . 72
5.4 Qualitative study of retardation and relativistic effects . . . . . . . . . . . 73
5.5 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

6 Numerical method and results 77


6.1 Numerical method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
6.1.1 The kinetic propagator Uk . . . . . . . . . . . . . . . . . . . . . . . 79
6.1.2 The potential propagator UV . . . . . . . . . . . . . . . . . . . . . . 82
6.1.3 The dual grid method . . . . . . . . . . . . . . . . . . . . . . . . . 84
6.2 Numerical results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
6.3 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

7 Outlook and conclusions 93

A Notations and conventions 97

B The Dirac equation for the free electron 101


B.1 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102

C Classical motion of an electron in a laser pulse 103


C.1 Initial conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
C.2 Non-relativistic case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
C.3 Relativistic case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
C.4 Numerical example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
C.5 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112

D Summary of approximations 113


D.1 Non-relativistic dipole approximation . . . . . . . . . . . . . . . . . . . . 113
D.1.1 Laboratory frame, velocity gauge . . . . . . . . . . . . . . . . . . . 113
D.1.2 Kramers-Henneberger frame . . . . . . . . . . . . . . . . . . . . . . 113
D.2 Non-relativistic approximation, first order retardation corection included . 114
D.2.1 Laboratory frame, velocity gauge . . . . . . . . . . . . . . . . . . . 114
D.3 Relativistic approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
D.3.1 Generalized Kramers-Henneberger picture . . . . . . . . . . . . . . 114
D.3.2 The “R” picture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
D.3.3 The “L” picture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115

ii
List of Figures

2.1 The energy levels E0 (full thick line, marked b0 ), E1 (full thin line, marked
b1 ) and the antibound state energy Ea (dashed line, marked a1 ), for the
two δ-function potential with the potential strength γ = 2 au, as function
of a . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.2 The dressed double δ-function potential for γ = 2 au and a = 2 au: (a)
α0 = 0.5 au; (b) α0 = 2 au; (c) α0 = 10 au. . . . . . . . . . . . . . . . . . 17
2.3 Bound and antibound levels of the dressed dressed double δ-function po-
tential for γ = 2 au, a = 2 au: even bound states - full thick lines, odd
bound states - full thin lines, even antibound states - thick dashed/dotted
lines, odd antibound states - thin dashed/dotted lines. . . . . . . . . . . . 18
2.4 The probability density of localization in the case γ = 2 au, a = 2 au
for the first four dressed states of the dressed dressed double δ-function
potential: (a) α0 = 0; (b) α0 = 0.5; (c) α0 = 2; (d) α0 = 10. . . . . . . . . 18
2.5 The total width ΓHF 1 and the partial width ΓHF 1,1 of the level E0 for the
double δ - function potential for frequencies ω = 2 (full thin line−total
width, dashed thin line−partial width) and ω = 4 (full thick line−total
width, dashed thick line−partial width). . . . . . . . . . . . . . . . . . . . 19
2.6 Trajectory of A(t) for a Gaussian pulse, for different widths τp , indicated
on graphs. The moment t = 0 is marked with a dot, and the sense of
motion by an arrow. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.7 Time dependence of the electric field magnitude E(t)/E0 for a Gaussian
pulse and two values of τp marked on each graph. . . . . . . . . . . . . . . 24
2.8 Ionization probability Pion of ground-state Hydrogen for √ various envelopes
1/2
and pulse widths τp (in cycles), as a function of I0 = 2E0 (in au). Black
curves - Gaussian pulses; red curves - sech pulses; blue curves - Lorentzian
pulses. Each panel corresponds to a fixed value of the frequency ω; the
value of τp is specified next to the Pion curves coalescent at small I0 . . . . 26
2.9 Black curve: ionization rate Γ (in au) for ground - state Hydrogen and
1/2
circular polarization at ω = 2 au, as a function of I0 , calculated from
the TDSE solutions. Red curve: Γ for linear polarization calculated by
Dondera et al. Blue curve: Γ for the LOPT one-photon ionization rate. . . 28
2.10 Comparison of Pion of ground - state Hydrogen, calculated from TDSE
(solid curves) and in the adiabatic approximation (dashed curves) for var-
ious envelopes and pulse durations. The values of τp are indicated. The
τp for Gaussian pulses carry no asterisk, for sech pulses they carry one
asterisk, and for Lorentzian pulses, two asterisks. . . . . . . . . . . . . . . 29
2.11 Comparison of Pion for circular and linear polarizations; sech pulses are
considered ω = 2 au and the indicated widths τp . Black curves: circular
polarization; red curves: linear polarization. . . . . . . . . . . . . . . . . . 29

5.1 The graph of A(φ) for a Gaussian pulse with FWHM τp = 1 cycle . . . . . 67

iii
5.2 The common trajectory of γ(t) and −rs (t). The positions at several mo-
ments of time are marked in red for γ(t) and in blue for −rs (t). The pulse
parameters are written on the figure. . . . . . . . . . . . . . . . . . . . . . 67
5.3 Plot of the distorted potential VR (r, t) in the plane y = 0 at several mo-
ments of time, marked on each panel. The laser pulse is Gaussian, with
parameters E0 = 100 au, ω = 1 au, τp = 1 cycle. . . . . . . . . . . . . . . . 71

6.1 Sketch of the algorithm used for the propagation over a time step ∆t . . . 79
6.2 The algorithm corresponding to the standard representation of the kinetic
propagator. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
6.3 The algorithm used for the propagation with the minimum number of copy-
ing operations to/from auxiliary vectors. . . . . . . . . . . . . . . . . . . . 81
6.4 Schematic two dimensional representation of the two grids, for the case
houter = 2hinner . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
6.5 Square modulus of the two parts of the wavefunction |Ψouter |2 (a) and
|Ψinner |2 (b) in the plane y = 0. The snapshot was taken at the moment
t = 1.5 T during the interaction of the atom with a gaussian pulse with
τp = 1 cycle, ω = 1 au and E0 = 50 au, (calculation done within the
nonrelativistic dipole approximation). . . . . . . . . . . . . . . . . . . . . . 86
6.6 Snapshots of density probability P(x, z, t) at the moments of time t =
−2T, −T, −T /2, −T /4, 0. The calculations were done within the NR-ND
approximation (left panel) and R approximation (right panel) for the pulse
for a Gaussian pulse with τp = 1 cycle and ω = 1 au at E0 = 100 au . . . . 89
6.7 Snapshots of density probability P(x, z, t) at the moments of time t =
T /4, T /2, T, 2T, 3T . The calculations were done within the NR-ND ap-
proximation (left panel) and R approximation (right panel) for the pulse
for a Gaussian pulse with τp = 1 cycle and ω = 1 au at E0 = 100 au . . . . 90

C.1 The vector potential A in the origin of the reference frame, for the case of
a linearly polarized Gaussian pulse of τp = 2, ω = 1 au. . . . . . . . . . . . 110
C.2 Case 1: the relativistic (in black) and non-relativistic (in red) trajectories
for a lineraly polarized pulse with τp = 2 cycles, ω = 1 au and for three
values of the amplitude A0 . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
C.3 Case 2, initial condition v0 = 0: the relativistic trajectory for monochro-
matic linearly polarized field of amplitude A0 = 10 au. . . . . . . . . . . . 111
C.4 Case 2, initial condition v0 = V0 n: the relativistic trajectories for
monochromatic linearly polarized field and four values of the amplitude
A0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

iv
List of Tables

2.1 Hydrogen quasienergies W = E − 2i Γ in a circularly polarized field at ω = 2


au and various α0 = Eω20 , according to our TDSE calculation and the Floquet
code of Potvliege. The latter values bear the subscript P. . . . . . . . . . . 27

6.1 Comparison of the ionization probabilities obtained with the HGM code
and R code for a Gaussian pulse with τp = 1 cycle and ω = 1 au at several
peak intensities; the calculations were done within the NR-D and NR-ND
approximations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
6.2 Ionization probability within the R model for a Gaussian pulse with τp = 1
cycle and ω = 1 au at different peak intensities. . . . . . . . . . . . . . . . 88

v
Author’s contribution

Papers

• C. Chirila, M. Boca, V. Dinu and V. Florescu, The dressed states of an electron in


a one-dimensional two δ-functions potential, Eur. Phys. J. D 27, 15 (2003).

• M. Boca, H. G. Muller and M. Gavrila, Dynamic stabilization of ground state


Hydrogen in super-intense circularly polarized laser pulses, J. Phys. B 37, 147
(2004).

Contributions to conferences

• M. Boca, H. G. Muller and M. Gavrila, Dynamic stabilization of ground state


Hydrogen in super-intense circularly polarized laser pulses, oral communication, 11th
International Laser Physics Workshop, Bratislava, 2002.

• M. Boca, V. Florescu and M. Gavrila, Generalized space-translated Dirac and


Pauli/Schrdinger equations for super-intense laser-atom interactions, poster com-
munication, 43th International School of Quantum Electronics, Erice, 2006.

• M. Boca, V. Florescu and M. Gavrila, A relativistic generalization of the Kramers


- Henneberger transformation, poster communication, IAMPI 2006, Szeged.

vii
Chapter 1

Introduction

The advances in the laser technology in the last decades led to a tremendous progress in
both experimental and theoretical study of the laser matter interaction. At the present
moment intensities up to 1021 W/cm2 are currently available, in a frequency range spread-
ing from IR to XUV. In this regime the electric field intensity is greater by orders of
magnitude than the electric field felt by a bound electron due to the Coulomb interaction.
Also the duration of laser pulses became shorter and shorter, durations of the order of
attoseconds (10−18 s) being presently available. In these conditions the atomic structure
is completely distorted by the external laser field, and new phenomena are revealed. We
shall briefly present some of them in the following.
Multiphoton ionization (MPI) takes place when an atom is ionized during the inter-
action with a laser field whose frequency is smaller than the ionization threshold, by
absorption of more than one photon. MPI was experimentally observed for the first time
in 1965 in two experiments performed by Voronov and Delone [1] and Hall et al [2]. For
low laser intensities I < 1011 W/cm2 the observed rates obey the prediction of the pertur-
bation theory. For larger intensities new phenomena appear: the atom can be ionized by
absorption of more than the minimum number of photons (the so-called above threshold
ionization - ATI) and the photoelectron spectrum consists of several maxima separated by
the photon energy ~ω. The first experimental evidence of ATI was obtained by Agostini
et al in 1979 [3]. For increasing intensity the observed photoelectron spectrum changes,
being no longer in agreement with the perturbation theory predictions: new maxima
appear at higher energies and the low energy peaks decrease or even disappear. This
phenomenon, named “peak suppression” is due to the AC Stark shift of the bound states
of the atom.
The harmonic generation (HG) process is the emission of odd order harmonics of the
incident radiation by an atom irradiated by a laser beam. It was observed experimentally
for the first time by New and Ward [4] and since then in several experiments harmonics
of very high order have been obtained. Recently it was proved that the HG spectrum
emitted by an atom can be used in order to obtain attosecond laser pulses. For a review
of this topic see, for example [5].

1
2 Introduction

1.1 Theoretical methods


The theoretical description of atoms in external electromagnetic fields can be done either
in a fully quantized formalism (quantized electron motion and radiation field) or in the
semiclassical approximation (quantized electron motion and classical external field). In
the very high intensity regime (∼ 1 au) and for frequencies of the order of atomic unit the
number of photons per mode is very large, such that the two descriptions are equivalent.
In this Thesis we shall adopt only the semiclassical method.
If the velocity acquired by the electron in the external field of amplitude E0 and
frequency ω is negligible with respect to the speed of light
eE0
c (1.1)

the non-relativistic dipole approximation is valid. The time evolution of the system is
governed by the Schrödinger equation and the laser fields can be considered a function
of time only. Beyond this point the magnetic Lorentz force is non-negligible, and the
retardation must be included in the Schrödinger equation. For even higher intensities,
when the velocity of the electron is of the same order of magnitude to the speed of light
eE0
∼c (1.2)

the relativistic effects become important and the Dirac equation must be used. We men-
tion here that for the frequency ω ∼ 0.05 au the intensity value corresponding to the
above limit is I ∼ 1018 W/cm2 , and for ω ∼ 1 au the intensity is I ∼ 1021 W/cm2 , values
which are in the range of currently available laser sources. For each of these three regimes
the behavior of the atomic systems is quite different, and distinct theoretical models are
required. Most of the theoretical methods developed so far are applicable within the
non-relativistic dipole approximation, but due the recent advances of the laser technology
more and more theoretical studies are dedicated to the non-dipole and relativistic effects.
At low intensity of the external electric field the time dependent perturbation theory
can be used in order to describe the atom-laser interaction [6]. Its simplest form is the
lowest order perturbation theory (LOPT) which predicts that the rate in a n-photon
ionization process is proportional to the n-th power of the external radiation intensity.
Discrepancies from the I n power law indicate the breakdown of LOPT. A problem related
to the lowest order perturbation theory is the fact that for resonant multiphoton processes
the transition matrix elements are singular. In order to avoid such difficulties the so-called
semiperturbative methods were developed. One of them is the essential states method in
which some resonant states considered important for the studied process are treated in a
non-perturbative way, while for the others perturbation theory is used.
The first non-perturbative approach is the so-called KFR method, based on the work
of Keldysh [7], Faisal [8] and Reiss [9]; widely used and subjected to a very large number of
subsequent extensions, it consists in approximating the ionization process by a transition
between an initial unperturbed bound atomic state and a final Gordon-Volkov state. At
low frequencies the so-called simple man (or three step) model [10, 11] succeeded to explain
Introduction 3

the main features of the experimental ATI and HG spectra and also the non-sequential
double ionization.
For the case of monochromatic radiation a very powerful fully non-perturbative ap-
proach is the Floquet theory [12] which allows calculation of ionization rates as solutions
of an eigenvalue problem. In the high frequency limit the high-frequency Floquet theory
(HFFT) developed by Gavrila and Kaminsky [13] predicts the so-called atomic stabiliza-
tion (the tendency of the atom to become stable against ionization for large enough field
intensity). Recently, results obtained for atoms interacting with finite laser pulses have
been interpreted in the frame of Floquet theory (see [14] and the references therein). Note
that not only ionization but also assisted electron atom scattering can be treated using
the Floquet method. The R-matrix Floquet theory, introduced by Burke, Francken and
Joachain [15, 16], combines the R-matrix and the Floquet methods, into a theory which
treats ionization, harmonic generation and laser assisted scattering in a unified way.

The brief description of the theoretical approaches presented in this Section is based
on following review papers:
M. Gavrila, Atomic structure and decay in high-frequency fields, in Atoms in Intense
Laser Fields, Academic Press 1992, pp 435.
C. J. Joachain, M. Dörr and N. J. Kylstra, High-intensity laser atom physics, in Adv.
At. Mol. Phys 42, pp 225 (1999).
M. Gavrila, Atomic stabilization in superintense laser fields, J. Phys. B 35, R147
(2002)
C. J. Joachain, Atoms in intense laser fields, in Proceedings of the International School
of Quantum Electronics, Erice 2006.

1.2 Numerical simulations


Due to the very fast development of the computers, numerical simulations became in the
last decades a very powerful and largely used tool to check various theoretical methods
and models developed to describe the laser-atom interaction. In this Section we briefly
review some of the main results published in relation to the problem of ionization of an
atom interacting with a laser field.
The problem of numerical solving the Floquet system of differential equation is a chal-
lenging one, even at the present level of computers performance. Nevertheless, Floquet
calculations have been performed for simple one-dimensional models but also for realistic
three dimensional systems. Floquet results for one dimensional Gaussian model poten-
tials have been presented by Bardsley and Comella [17], Yao and Chu [18], Marinescu
and Gavrila [19]. Another frequently used one-dimensional model is the so-called “soft-
Coulomb” potential, a Coulomb potential whose singularity in the origin is smoothed.
Comprehensive Floquet maps have been calculated by Wells et al [20, 21] and later by
Stroe et al [22, 23], who have also discussed the results in the frame of high-intensity
4 Introduction

high-frequency Floquet theory. Results for the one dimensional δ potential were given by
Groshdanov et al [24], Sanpera et al [25], LaGatutta [26] and by Dörr and Potvliege [27].
The realistic three dimensional Floquet problem is a much more difficult one; however,
high frequency studies were done for the Hydrogen atom (Pont et al [28, 29, 30, 31], Voss
and Gavrila [32]), H2+ (Shertzer et al [33]), H − (Gavrila and Shertzer [34], Muller and
Gavrila [35] and Duijin et al [36]). A powerful method was developed by Potvliege [37]
for solving the exact Floquet system at finite frequency, and also accurate results were
obtained using the complex scaling method developed by Moiseyev [38].
A completely different approach to find the Floquet solutions is based on the study of
the time evolution of atomic systems interacting with external electromagnetic pulses in
the frame of the multistate Floquet theory. A description of the theoretical basis and a
numerical example are presented in Chapter 2.

The numerical integration of the time-dependent Schrödinger/Dirac equation is the


most direct and powerful method to study the interaction of atomic systems with external
fields. However, the problem is a very difficult one and until now there are no exact
calculations for systems with more than two active electrons.
The first calculations were done within the non-relativistic dipole approximation, for
model one or two dimensional systems in the beginning, and then for realistic three dimen-
sional cases. We mention the work of Eberly and coworkers [39], for an one-dimensional
soft-Coulomb model potential and the extensive study of Patel et al [40] for a screened
Coulomb potential. The one dimensional δ potential was also intensely studied (Geltmann
[41], Su et al [42], Sanpera et al [25], Dörr and Potvliege [27]) following a controversy over
the presence of the dynamic stabilization in this model. The advance of computers made
possible the numerical integration of the time-dependent Schrödinger equation for two
and three dimensional systems. Two dimensional soft-Coulomb potentials and arbitrary
polarization have been considered by Protopapas et al [43] Patel et al [44] and Chism
et al [45], their results showing dynamic stabilization. A very powerful code written by
Muller [46] allows the integration of the time-dependent Schrödinger equation for three
dimensional systems within the single active electron (SAE) approximation. It was used
by Dondera et al [47, 48] for a detailed study of the stabilization of the ground state Hy-
drogen in linearly polarized laser fields, and also for numerical simulations of the above
threshold ionization of Argon [49, 50]. A study [51], realized with the same code, of the
Hydrogen stabilization in a circularly polarized laser field is presented in Chapter 2 of the
Thesis. Stabilization of the atom within single active electron approximation was also
studied by Gajda et al [52], Piraux and Potvliege [53], Bauer and Ceccherini [54], Choi
and Chism [55]. We mention here also the numerical integration of the time dependent
Schrödinger equation for Helium presented by Muller [56, 57].
As the experimental radiation sources became capable to produce more and more
intense fields, the necessity appeared to take into account in numerical simulations the
coordinate dependence of the laser field. In the last decade several calculation were done.
It was shown that the first retardation correction term is in fact equivalent to a new,
fictitious, time dependent component of the electric field oriented along the propagation
Introduction 5

direction. This property allowed the use of the codes already developed for the case of the
dipole approximation for description of retardation effects; so, results for realistic three-
dimensional models were obtained (Bugacov et al [58], Muller [59]). The simulations
have shown that the main effect of the magnetic field component, taken into account
through the inclusion of the non-dipole terms is the atomic destabilization; the reason
is that the Lorentz force pushes the electron along the field propagation direction such
that at the end of the laser pulse the electron is left very far from the nucleus so it
can not be recaptured. Later Kylstra et al [60] and Vasquez de Aldana et al [61] have
tested the validity of this approximation by comparing the results with those obtained
for the exact non-dipole Schrödinger equation in a two dimensional model calculation.
They have also shown that by using two counter propagating pulses polarized along the
same direction one can not recover the atomic stabilization, although the magnetic drift
is suppressed. In 2002 Vasquez de Aldana and Roso [62] have published results for the
exact nondipole three dimensional Schrödinger equation; they confirmed the previous
two dimensional simulations and have also shown that the atomic stabilization could
be reobtained by using counter-propagating pulses with crossed polarizations. A different
approach to include retardation effects in the Schrödinger equation was proposed by Førre
et al [63, 64] who defined a nondipole generalization of the Kramers-Henneberger space
translation method. A similar method was used by Mahmoudi et al [65] for the case of a
free electron in a laser pulse.
The Dirac equation for an atom in intense laser field is a much more difficult problem;
until now only a few one- (Protopapas et al [66], Kylstra et al [67], Lenz et al [68]) or two-
dimensional simulations ([69]) have been published. For the realistic three dimensional
case, although several algorithms were proposed ([70, 71, 72]), only results for the free
electron in an external plane wave laser field exist [73, 74, 75]. A promising method
was proposed by Krstic and Mittelman [76, 77]. It consists in defining a relativistic
generalization of the Kramers-Henneberger transformation, that would allow reducing the
Dirac equation to a Schrödinger like-equation for a relatively large class of phenomena. It
was applied by Ermolaev [78] to the high frequency Floquet problem of a one-dimensional
model atom. It also constitutes the starting point of the approach presented in this Thesis
in Chapters 4 - 6.

1.3 Outline
This Thesis is dedicated to the problem of ionization of an atom interacting with a super-
intense laser. We present the basics of the theory, relevant results present in the literature
and also new theoretical and numerical results.
In Chapter 2 is studied the interaction of one-electron atoms with intense electro-
magnetic fields in the frame of non-relativistic dipole approximation. There are presented
elements of Floquet theory and numerical calculations of quasienergies for a simple one-
dimensional model potential. The second part of the Chapter is a study of the stabilization
of the ground state Hydrogen atom in a super-intense laser pulse.
6 Introduction

Chapter 3 is dedicated to the study of the Volkov solutions which are used in the
theoretical model presented in the following Chapter. Their properties (orthogonality,
normalization and completeness) are discussed.
In Chapter 4 is presented a relativistic generalization of the space-translation method
for atom-laser interaction, which is applicable for arbitrary plane wave laser pulses. We
also develop the “low momentum regime approximation” of the theory, valid for light
atoms initially in a superposition of low energy states and arbitrary intense laser pulses.
The evolution equation in the atomic approximation, called “the generalized translated
Schrödinger equation”, its properties and qualitative predictions on the time-evolution of
the system are presented in the Chapter 5.
In Chapter 6 are presented results of numerical simulations of the interaction of
atoms with superintense laser pulses in the frame of the atomic approximation discussed
before.
Appendix A contains a list of notations used in the Thesis.
In Appendix B are presented the solution of the free Dirac equation in the standard
and Foldy-Wouthuysen representations.
In Appendix C is presented the solution of the equation of motion of an classical elec-
tron in a external laser pulse in non-relativistic and relativistic treatment. Also numerical
examples are presented.
Appendix D contains a list of the main results obtained in Chapters 4 and 5.

1.4 References
[1] G. Voronov and N. Delone, JETP Letters 1, 66 (1965).

[2] J. Hall, E. Robinson, L. Branscomb, Phys. Rev. Lett. 14, 1013 (1965).

[3] P. Agostini, F. Fabre, G. Mainfray, G. Petite, N. Rahman, Phys. Rev. Lett. 42, 1127
(1979).

[4] C. New, J. Ward, Phys. Rev. Lett. 19, 556 (1967).

[5] P. Agostini and L. F. DiMauro, Rep. Prog. Phys., 67, 813 (2004).

[6] F. H. M. Faisal, Theory of multiphoton processes, Plenum Press New York (1986).

[7] L. V. Keldysh, Sov. Phys. JETP 20, 1307 (1965).

[8] F. H. M. Faisal, J. Phys. B 6 L32 (1973).

[9] H. R. Reiss, Phys Rev A 22, 1786 (1990).

[10] K. C. Kulander, K. J. Schafer and J. L. Krause in: B. Piraux, A. L’Huilier and K.


Rzazewski (eds.) Super-Intense Laser-atom Physics 316, Plenum Press New York 1993.

[11] P. Corkum, Phys. Rev. Lett. 71, 1993 (1994).


Introduction 7

[12] J. H. Shirley, Phys. Rev. B 138, 979 (1965).

[13] M. Gavrila and J. Z. Kaminski, Phys. Rev. Lett. 52, 614 (1984).

[14] M. Gavrila, J. Phys. B 35, R147 (2002).

[15] P. G. Burke, P. Francken and C. J. Joachain, Europhys. Lett 13, 617 (1990).

[16] P. G. Burke, P. Francken and C. J. Joachain, J. Phys. B 24, 761 (1991).

[17] J. N. Bardsley and M. J. Comella, Phys. Rev. A 39, 2252 (1989).

[18] G. Yao and S. I. Chu, Phys. Rev. A 45, 6735 (1992).

[19] M. Marinescu and M. Gavrila, Phys. Rev. A 53, 2513 (1996).

[20] J. C. Wells, I. Simbotin and M. Gavrila, Phys. Rev. Lett. 80, 3479 (1998).

[21] J. C. Wells, I. Simbotin and M. Gavrila, Phys. Rev. Lett. 82, 665 (1999).

[22] I. Simbotin, M. Stroe, and M. Gavrila, Laser Phys. 14, 482 (2004)

[23] M. Stroe, PhD thesis, Bucharest 2004.

[24] T. P. Grozdanov, P. S. Krstic and M. H. Mittleman, Phys. Lett. A 149, 144 (1990).

[25] A. Sanpera, Q. Su and L. Roso-Franco, Phys. Rev. A 47, 2312 (1993).

[26] K. J. LaGattuta, Phys. Rev. A 49, 144 (1994).

[27] M. Dorr and R. M. Potvliege, J. Phys. B 33, L233 (200).

[28] M. Pont Phys. Rev. A 40, 5659 (1989).

[29] M. Pont, N. R. Walet, and M. Gavrila, Phys. Rev. Lett 61, 939 (1988).

[30] M. Pont, N. R. Walet and M. Gavrila, Phys. Rev. A 41, 477 (1990).

[31] M. Pont and M. Gavrila, Phys. Rev. Lett 65, 2362 (1990).

[32] R. J. Vos and M. Gavrila, Phys. Rev. Lett. 68, 170 (1992).

[33] J. Shertzer, A. Chandler and M. Gavrila Phys. Rev. Lett. 73, 2039 (1994).

[34] M. Gavrila and J. Shertzer, Phys. Rev. A 53, 343 (1996).

[35] H. G. Muller and M. Gavrila, Phys. Rev. Lett 71, 1693 (1993)

[36] E. van Duijin, M. Gavrila and H. G. Muller, Phys. Rev. Lett 77, 3759 (1996).

[37] R. M. Potvliege, Comput. Phys. Commun. 114, 42 (1998).

[38] N. Moiseyev, Phys. Rep. 302, 211 (1998).


8 Introduction

[39] J. H. Eberly, R. Grobe, C. K. Law and Q. Su, in Atoms in Intense Laser Fields, Academic
Press 1992.

[40] A. Patel, N. J. Kylstra and P. L. Knight, J. Phys. B 32, 5759 (1999).

[41] S. Geltman, Phys. Rev. A 45, 5293 (1992), J. Phys. B 27, 1497 (1994), J. Phys. B 32,
853 (1999).

[42] Q. Su, B. P. Irving, C. W. Johnson and J. H. Eberly, J. Phys. B 29, 5755 (1996).

[43] M. Protopapas, D. G. Lappas and P. L. Knight, Phys. Rev. Lett. 79, 4550 (1997).

[44] A. Patel, M. Protopapas, D. G. Lappas and P. L. Knight, Phys. Rev. A 58, R2652
(1998).

[45] W. Chism, D. I. Choi and L. E. Reichl, Phys. Rev. A 61, 054702 (2000).

[46] H. G. Muller, Laser Phys. 9, 138 (1999).

[47] M. Dondera, H. G. Muller and M. Gavrila, Laser Phys. 12, 415 (2002)

[48] M. Dondera, H. G. Muller and M. Gavrila, Phys. Rev. A 65, 031405(R) (2002)

[49] H. G. Muller and F. C. Kooiman, Phys. Rev. Lett 81, 1207 (1998).

[50] M. J. Nandor, M. A. Walker, L. D. Van Woerkom, H. G. Muller, Phys. Rev. A 60, R1771
(1999).

[51] M. Boca, H. G. Muller and M. Gavrila, J. Phys. B 37, 147 (2004).

[52] M. Gajda, B. Piraux and K. Rzazewski K, Phys. Rev. A 50, 2528 (1994).

[53] B. Piraux and R. M. Potvliege, Phys. Rev. A 57, 5009 (1998) .

[54] D. Bauer and F. Ceccherini, Phys. Rev. A 66, 053411 (2002).

[55] D. I. Choi and W. Chism, Phys. Rev. A 66, 025401 (2002).

[56] H. G. Muller, Optics Express 8, 86 (2001).

[57] H. G. Muller, Optics Express 8, 417 (2001).

[58] A. Bugacov, M. Pont and R. Shakeshaft, Phys. Rev. A 48, R4027 (1993).

[59] H.G. Muller: Weakly relativistic stabilization : the effect of the magnetic field. In: Super-
Intense Laser-Atom Physics, edited by B. Piraux and K. Rzazewski. (Kluwer, 2001), p.
339-344.

[60] N. J. Kylstra, R. A. Worthington, A. Patel, P. L. Knight, J. R. Vázquez de Aldana, and


L. Roso, Phys. Rev. Lett. 85, 1835 (2000).
Introduction 9

[61] J. R. Vázquez de Aldana, N. J. Kylstra, L. Roso, P. L. Knight, A. Patel, and R. A.


Worthington, Phys. Rev. A 64, 013411 (2001).

[62] J. R. Vázquez de Aldana and L. Roso, JOSA B 19, 1467 (2002).

[63] M. Førre, S. Selstø, J. P. Hansen, and L. B. Madsen, Phys. Rev. Lett 95, 043601 (2005).

[64] M. Førre, J. P. Hansen, L. Kocbach, S. Selstø, and L. B. Madsen, Phys. Rev. Lett 97,
043601 (2006).

[65] M. Mahmoudi, Y. I. Salamin and C. H. Keitel, Phys. Rev. A 72, 033042 (2005).

[66] M Protopapas, C H Keitel and P L Knight, J. Phys. B 29, L591 (1996).

[67] N. J. Kylstra, A. M. Ermolaev, and C. J. Joachain, J. Phys. B 30, L449 (1997).

[68] E. Lenz, M. Dörr and W. Sandner,Laser Phys. 11, (2001) 216-220.

[69] U. W. Rathe, C. H. Keitel, M. Protopapas and P. L. Knight, J. Phys. B 30 L531 (1997).

[70] J. W. Braun, Q. Su, and R. Grobe, Phys. Rev. A 59, 604 (1998).

[71] K. C. B. New, K. Watt, C. W. Misner, and J. M. Centrella, Phys. Rev. D 58, 064022
(1998).

[72] J. Yepez, arXiv.org:quant-ph/0210093.

[73] J. San Roman, L. Roso and H. R. Reiss, J. Phys B 33, 1896 (2000).

[74] J. San Roman, L. Roso and L. Plaja, J. Phys. B 36, 1 (2003).

[75] J. San Roman, L. Roso and L. Plaja, J. Phys. B 36, 2253 (2003).

[76] P. S. Krstic and M. H. Mittleman, Phys. Rev. A 42, 4037 (1990).

[77] P. S. Krstic and M. H. Mittleman, Phys. Rev. A 45, 6514 (1992).

[78] A. M. Ermolaev, J. Phys. B 31, L65 (1998).


Chapter 2

Atomic stabilization: the


non-relativistic case

The topic discussed in this Chapter is the “atomic stabilization”, one of the most spectac-
ular phenomena which take place when an atomic system interacts with an superintense
laser pulse. The atomic stabilization is defined as the tendency of an atom interacting
with an ultraintense electromagnetic radiation to become stable against ionization when
the intensity increases beyond a certain limit.
Two types of stabilization are defined: the quasistationary stabilization and the dy-
namic stabilization. The close relation between them has been only recently fully under-
stood. The so-called “quasistationary (or adiabatic) stabilization” is the tendency of the
ionization rate of an atom in an external monochromatic field to decrease when the field
intensity increases at fixed frequency. In contrast, the term “dynamic stabilization” is
used in relation to the behavior of an atom interacting with a finite pulse and it repre-
sents the property of the ionization yield to decrease when the pulse intensity increases.
Both quasistationary and dynamic stabilization are in contradiction with the predictions
of the low order perturbation theory and they can only take place in an intensity regime
well beyond the perturbative one.
Since its theoretical prediction two decades ago important progress has been achieved
in understanding the atomic stabilization, and a very large number of papers treating
various aspects of this problem have been published. Also the experimental evidence of
the dynamic stabilization was obtained in two experiments carried at FOM-AMOLF. A
recent review of this topic was published by Gavrila in 2003 [1].
In this Chapter we shall briefly present the theoretical aspects of the quasistationary
(Section 2.1) and dynamic (Section 2.2) stabilization and also original numerical results
related to this issue. The validity of the non-relativistic dipole approximation is assumed.

2.1 Quasistationary stabilization


The quasistationary stabilization was predicted by M. Gavrila in the frame of the high
frequency Floquet theory; we shall briefly present here some of the main results, for a

11
12 Atomic stabilization: the non-relativistic case

detailed review see [1], [2].


The external electromagnetic field treated in the dipole approximation is described by
the vector potential

A(t) = A0 [e1 cos(ωt) + e2 tan δ sin(ωt)] (2.1)

where A0 is a real amplitude, e1 and e2 are orthogonal unity vectors oriented along the
main axes of the polarization ellipse, and the parameter δ gives the polarization. Linear
polarization along the e1 direction is obtained for δ = 0 and circular polarization for
δ = π4 . The amplitude of the corresponding electric field is E0 = Aω0 and the field intensity
is
A20
I= . (2.2)
2ω 2 cos2 (δ)
The classical quiver motion of an electron interacting with the electromagnetic field is
given by the equation
Zt
e
α(t) = − dt A(t); (2.3)
m

we shall denote its amplitude by α0 = − eA mω


0
.
The time dependent Schrödinger equation written in the velocity gauge for an atom
interacting with the external field (2.1) is
" #
(P − eA(t))2 ∂Ψv
+ V (r) Ψv (r, t) = i~ . (2.4)
2M ∂t

It is convenient to perform two unitary transformation on this equation: first, we shall


eliminate the term in A2 (t) by a phase transformation and then we shall perform a spatial
translation by −α(t). The result is the Schrödinger equation written in the so-called
Kramers-Henneberger frame (a reference frame oscillating along the trajectory α(t) of
the free electron).
 2 
P ∂Ψ
+ V (r + α(t)) Ψ(r, t) = i~ . (2.5)
2M ∂t

The previous equation was discovered by Pauli and Fierz [3], extensively used by Kramers
[4] and rediscovered by Henneberger [5], and also by Faisal [6].
Since the Hamiltonian in the previous equation is periodic in time with the period
T = 2π ω
of the electromagnetic field the Floquet theorem applies [7]; it states that the
equation admits particular solutions of the form
 
i
Ψ(r, t) = exp − W t Φ(r, t), (2.6)
~
Atomic stabilization: the non-relativistic case 13

where W is generally a complex number named quasienergy and Φ(r, t) is periodic in time
with the period T .

Φ(r, t + T ) = Φ(r, t). (2.7)

The next step is to insert in Eq.(2.5) the Fourier expansions of the translated potential
V (r + α(t)) and of the periodic part Φ(r, t) of the solution

X ∞
X
−inωt
V (r + α(t)) = e Vn (r; α0 ), Φ(r, t) = e−inωt φn (r); (2.8)
n=−∞ n=−∞

(in the above formulae the parametric dependence on α0 of the Fourier components of
the potential has been explicitly written). One obtains the Floquet system of differential
equations

~2
X   
−δnm ∆ + Vn−m (r; α0 ) φm (r) = (W + n~ω) φn (r),
m=−∞
2M
n = −∞, . . . , +∞. (2.9)

In order to obtain solutions which describe ionization states of the atom one must solve
the previous system with Gamow-Siegert boundary conditions
r
exp (i kn r) 2M (W + n~ω)
φn (r) ∼ fn (r̂, kn ) , kn = ± . (2.10)
r→∞ r ~2
With this conditions1 the quasienergies W take discrete complex values. Depending on
the sign chosen in front of the square root in the previous equation, a Floquet solution
can be physical if its components decrease exponentially in every channel n that obeys
the condition <(W + n~ω) < 0 (the so-called closed channels) and increase exponentially
for the channels with <(W + n~ω) > 0 (the open channels)

=(kn ) > 0 if <(W + n~ω) < 0, =(kn ) < 0 if <(W + n~ω) > 0, (2.11)

or unphysical if the above conditions are not satisfied in at least one channel. One can
prove that for physical solutions the imaginary part of the quasienergy of a Floquet state
is always negative and the ionization rate Γ in that state is twice the absolute value of
the imaginary part of the quasienergy.
Γ
=(W ) = − (2.12)
2
The concept of quasistationary stabilization was discussed for the first time in the
frame of the high frequency Floquet theory. This theory, introduced by Gavrila and
Kaminsky [8] in relation to the laser assisted scattering problem, is a formalism which
allows the expansion of the solutions of the Floquet system of differential equations as
1
For Coulombic tail potentials the logarithmic phase must be included in the outgoing wave.
14 Atomic stabilization: the non-relativistic case

power series in 1/ω. In the limit ω → ∞ the Floquet system of differential equations re-
duces to the so-called structure equation, which is the usual time-independent Schrödinger
equation for the zeroth Fourier component of the potential (2.8)
~2
− ∆φ0 (r) + V0 (r, α0 )φ0 (r) = W φ0 (r) (2.13)
2M
The potential V0 (r, α0 ) in the previous equation is called “the dressed potential” and is
the time average of the potential seen by an electron oscillating along the trajectory (2.3)
ZT
1
V0 (r, α0 ) = dt V (r + α(t)). (2.14)
T
0

The ionization rate writes as a sum over the open channels of partial rates; in the first
order of the high-frequency approximation its expression is:
Z
(1)
X
(1) m X (−)
Γ = Γn = 2
kn dΩhukn |Vn |u0 i, (2.15)
n≥n
(2π~) n≥n
0 0

(−)
where u0 is a bound state of the field-free potential, and ukn is the outgoing ionization
state of the dressed potential, normalized in the energy scale. The validity criterion for
the high-frequency Floquet theory is
~ω  W0 (α0 ) (2.16)
where W0 (α0 ) is the binding energy of the dressed potential V0 (r, α0 ). It follows that for
high enough frequencies the Floquet solutions are almost identical with the eigenstates of
the dressed potential V0 (r, α0 ), and the corresponding quasienergies are very close to the
(real) energy levels of the structure equation, their imaginary part being small.
The behavior of the Floquet quasienergies at fixed frequency and variable intensity is
the following: in the zero field limit they originate from the field free eigenstates of the
atomic potential; in the small intensity range the perturbation theory is valid, and the
ionization rates have a pronounced tendency to increase with the intensity which leads to
a maximum of Γ (the so-called “death valley”). If the field intensity increases further for
almost all the realistic model potentials the ground state energy of the dressed potential
starts to decrease, the condition (2.16) is satisfied and the Floquet quasienergies get closer
and closer to the real eigenvalues of V0 (r, α0 ).
A recently developed version of the high frequency Floquet theory is the so-called high
intensity high frequency Floquet theory [9, 10]. It is based on the fact that the ground
state energy of the dressed potential decreases with increasing α0 ; as a consequence, for
any frequency the condition (2.16) is obeyed if the intensity in large enough. Then, even
for frequencies that are much lower than the zero field ground state energy of the atomic
potential the high frequency Floquet theory is applicable in the large intensity limit and
the Floquet quasienergies tend to become identical with the (α0 dependent) eigenstates of
the dressed potential. One must note, however, that if the frequency is very small it may
Atomic stabilization: the non-relativistic case 15

happen that the intensity required is greater than the validity limit of the non-relativistic
dipole approximation which was assumed, such that in fact the theory does not apply.
Also other phenomena have been extensively discussed in the frame of Floquet theory;
among them are the appearance of the light-induced states and the avoided crossings.
The light induced states are unphysical Floquet states which become physical when the
field intensity increases. Sometimes, this appearance of new states is accompanied by the
disappearance of another physical state, i.e. its transformation into a unphysical one.
For a very comprehensive discussion on light-induced and light-suppressed states in short
range potentials see [10] and the references therein. The so-called avoided crossings are a
consequence of the impossibility for a pair of Floquet states to have equal quasienergies at
the same field intensity; this is a generalization of the non-crossing rule of the eigenvalues
of a hermitian Hamiltonian dependent on a real parameter [11]. When two quasienergies
tends to come very close, they may have a extremely abrupt change in their dependence
of the field intensity; if the distance between them is represented as a function of the field
intensity it has a very sharp minimum which is known as avoided crossing. A theoretical
study of the avoided crossings in the frame of Floquet theory was given by Potvliege and
Shakeshaft [12], also new results were recently presented by Stroe [10].

2.1.1 Numerical examples


We present in this Section the detailed study of the bound state solutions of the struc-
ture equation for the one dimensional two δ - function potential [13], and their widths
calculated in the frame of high frequency Floquet theory. Atomic units are used.

We consider an electron in the field of the symmetric two δ - function potential


γ
V (x) = − [ δ(x − a) + δ(x + a) ] , a > 0, (2.17)
2
which can be looked at as an extremely simplified model for a diatomic ion [14, 15]. Its
eigenvalues problem in the absence of the external electromagnetic field is a textbook one
[16]: with the notations
κ2
E=− , η = 2γa, (2.18)
2
the equation for the bound levels is
2κb a = η (1 ± e−2 κb a ) , (2.19)
the upper sign corresponding to even solutions, and the lower one to odd solutions. Sim-
ilarly, the equations for the antibound states is
2κa a = −η (1 ± e2 κa a ) . (2.20)
For η < 1 Eq.(2.19) has only one solution, for the upper sign, and the equation (2.20) has
also only one solution but for the lower sign, i.e. the system has an even bound state and
16 Atomic stabilization: the non-relativistic case

b1

-0.5 a1

-1
E (a.u.)

-1.5

b0

-2

-2.5
0.001 0.01 0.1 1 10
a (a.u.)

Figure 2.1: The energy levels E0 (full thick line, marked b0 ), E1 (full thin line, marked
b1 ) and the antibound state energy Ea (dashed line, marked a1 ), for the two δ-function
potential with the potential strength γ = 2 au, as function of a .

an odd antibound one. For η > 1 the bound states equation admits two solutions, even
and odd respectively, and the equation for the antibound states doesn’t have any, i.e. at
the value η = 1 the antibound state turns into a bound one. In Fig.2.1 the energies of
the two states are represented as functions of the parameter a, for γ = 2 au.
For a = 0 the potential reduces to a single δ potential, of strength γ; for a > 2 the
exponential function in Eq.(2.19) becomes negligible and the two bound states are almost
degenerate. The following scaling laws for the energy levels can be written
1
En (γ, a) = En (η, 1) = γ 2 En (1, η). (2.21)
a2
Note that in the absence of the field the system has also an infinity of resonances (see
[17]), which shall not be discussed here.
The dressed potential V0 (α0 , x) in this case has the simple analytic expression:
V0 (α0 ; x) = V0δ (α0 ; x − a) + V0δ (α0 ; x + a) , (2.22)
with
 γ
 − π √α20 −x2 , | x | < α0
1

V0δ (α0 ; x) = (2.23)

 0, | x | > α0
In Fig.2.2 the dressed potential with the parameters γ = a = 2 au is represented for
three values of α0 : 0.5, 2 and 10 au. For α0 < a, the dressed potential is non-vanishing
in two non-overlapping regions x ∈ (−a − α0 , −a + α0 ) and x ∈ (a − α0 , a + α0 ), each of
them bounded by singularities. At α0 = a the singularities xs = a − α0 and xs = −a + α0
Atomic stabilization: the non-relativistic case 17

0 0 0
(a) (b) (c)
V0 (x) (a.u.)

-1 -1 -1

-2 -2 -2

-3 -1.5 0 1.5 3 -5 -2.5 0 2.5 5 -15 -10 -5 0 5 10 15


x (a.u.) x (a.u.) x (a.u.)

Figure 2.2: The dressed double δ-function potential for γ = 2 au and a = 2 au: (a)
α0 = 0.5 au; (b) α0 = 2 au; (c) α0 = 10 au.

coalesce in the origin and the non-vanishing regions become adjacent. They overlap for
α0 > a in the interval (a − α0 , −a + α0 ) , along which both terms in (2.22) contribute.
Unlike in the field-free case, only one scaling law for the energy levels of the dressed
potential can be found

E(γ, α0 , a) = γ 2 E(1, α0 , η) . (2.24)

Due to the singularities of the dressed potential, the corresponding structure equation
cannot be directly solved with good accuracy. The numerical method used is a general-
ization of that described in [18] for the simpler problem of a single δ function. It uses a
change of variable which eliminates the singularities, allowing the solving with very high
accuracy for arbitrary large α0 .
In Fig.2.3 are presented the energies of the first seven bound levels and of ten antibound
states as functions of α0 for γ = 2 au, a = 2 au. The first two bound states exist in the
zero field limit, while the others are light-induced. At α0 = 0 the two bound states b0
and b1 are almost degenerate (see Fig.2.1); they split apart around α0 ≈ 2 and for large
α0 the odd-even degeneracy sets in again. For the higher states there is also present
the tendency to coalesce in pairs at large α0 . One must note, however, that the dressed
double δ potential has four singular points; as a consequence the corresponding bound
states have a “polichotomic” structure rather than the normal dichotomic one. In order
to study the appearance of this polichotomic structure we have represented in Fig.2.4 the
squared modulus of the first four eigenfunctions vn (x) of the dressed potential as functions
of x in the region 0 < x < 20. The same values γ = 2 au, a = 2 au have been chosen
for the potential parameters. Fig.2.4(a) corresponds to α0 = 0. In this case, only the
ground state b0 and the first excited state b1 exist; both present the discontinuity of the
first derivative at x = ±a associated to the δ function. At α0 = 0.5 (fig. 2.4(b)), the first
derivative is smooth, but the two bound states still have narrow maxima around x = ±a
(dichotomic structure). The maximum of the probability density in the ground state
18 Atomic stabilization: the non-relativistic case

,
a6
-0.0001
,
a5 b6

,
a4 a6
-0.001 a5 b5
a3
, b4
b3 a4
E (a.u.)

-0.01 , b2 a3
a2

a2
b1
-0.1

b0

-1
0 5 10 15 20 25 30 35 40 45 50
α0 (a.u.)

Figure 2.3: Bound and antibound levels of the dressed dressed double δ-function potential
for γ = 2 au, a = 2 au: even bound states - full thick lines, odd bound states - full thin
lines, even antibound states - thick dashed/dotted lines, odd antibound states - thin
dashed/dotted lines.

0.3
0.25 α0 = 0 0.25
α0 = 0.5
0.2 0.2
2
|vn(x)|

0.15 0.15
b0 b0
0.1 0.1
(a) (b)
0.05 b1 0.05 b1
0 0
0 2 4 6 8 10 0 2 4 6 8 10
0.3 0.1
0.25 0.08
α0 = 2 b2 α0 = 10
0.2 b0 b1
0.06
2
|vn(x)|

0.15 b1 b0
0.04
0.1
(d)
0.05 b2 (c) 0.02
b3
0 0
0 4 8 12 16 20 0 4 8 12 16 20
x (a.u.) x (au)

Figure 2.4: The probability density of localization in the case γ = 2 au, a = 2 au for the
first four dressed states of the dressed dressed double δ-function potential: (a) α0 = 0;
(b) α0 = 0.5; (c) α0 = 2; (d) α0 = 10.
Atomic stabilization: the non-relativistic case 19

follows closely the behavior of the potential in Fig.2.2: with increasing α0 it moves from
x close to 2 au toward lower values. For α0 = 2 (fig. 2.4(c)) this maximum has reached
the origin and the distribution is narrow. For larger values of α0 the distribution becomes
broader and the central maximum splits in two maxima at x ≈ α0 − a and x ≈ −α0 + a;
it is also visible the tendency to form another pair of maxima located at x ≈ α0 + a and
x ≈ −α0 − a (the onset of polichotomic structure). At α0 = 2 (fig.2.4(c)) the dressed
potential support three bound states, and four bound states at α0 = 10 (fig.2.4(d)). For
the excited states the maxima located around the four singularities of the potential are
also present, although for them the picture is complicated by the presence of the nodes.
We have also calculated the widths acquired by the ground state of the structure
equation in the first order of the high-frequency approximation, for two values of the
frequency: ω = 2 au and ω = 4 au. In Fig.2.5 are represented, as function of α0 , for the
case a = 2 au, γ = 2 au, the partial one photon ionization rates (with thin lines) and
the total ionization rates for the two frequencies above. A number of 120 channels was
needed to reach the convergence of the total width at ω = 2 au, and 100 channels for
ω = 4 au. The behavior of ΓHF illustrates the quasistationary stabilization.
0.2

0.15
Γ (a.u.)

0.1

0.05

0
0 1 2 3 4 5
α0 (a.u.)

Figure 2.5: The total width ΓHF 1 and the partial width ΓHF 1,1 of the level E0 for the
double δ - function potential for frequencies ω = 2 (full thin line−total width, dashed
thin line−partial width) and ω = 4 (full thick line−total width, dashed thick line−partial
width).

2.2 Dynamic stabilization


An atom acted upon by a laser pulse is ionized; for small intensities (in the so-called
perturbative domain) the ionization probability increases with the laser intensity accord-
ingly to the predictions of the perturbation theory. For large intensities of the external
field a new regime may set in: the total ionization probability tends to level off at a
value lower than 1, or even to decrease with increasing field intensity. This phenomenon,
known as “dynamic stabilization” was the subject of many theoretical studies during the
20 Atomic stabilization: the non-relativistic case

last two decades. Also it was experimentally confirmed in two experiments performed at
FOM-AMOLF [19, 20]. Note, however, that the atomic stabilization is not always en-
countered in the laser-atom interaction; in fact, it requires several conditions on the laser
pulse intensity, shape, duration and frequency to be simultaneously met. In this section
we present briefly theoretical aspects of the atomic stabilization, and also original results
of numerical simulations.
In a numerical experiment the ionization of an atom in an intense field is studied
by direct numerical integration of the time-dependent Schrödinger equation. Usually the
initial condition is chosen as a bound state of the atom at a moment ti before the pulse has
reached the atom and the wavefunction is propagated in time up to a moment tf when the
external field extinguished and the atom is free again. Usually, in a stabilization problem
the quantities of interest are the photoelectron spectrum and the survival probability of
the atom; they are calculated by projecting the wavefunction at the moment tf on the
field-free atomic system eigenstates.

P(E) = |huE |Ψ(tf )i|2 , Pn = |hun |Ψ(tf )i|2 . (2.25)

The ionization probability is defined as the total probability to find the electron in the
continuum
Z∞
Pion = dE P(E) (2.26)
0

If one is interested only to calculate the total ionization probability and not the photo-
electron spectrum itself usually is more convenient to use use the formula
X
Pion = 1 − Pn , (2.27)
n

as is easier to calculate the projection of the wave-function on the bound states of the
system than on the continuum states. Note that most of the model potentials used in
numerical experiments support an infinity of bound levels; in this case in the previous
equation one must include a number of terms large enough such that the convergence of
the result is reached. The equations and definitions above are written for a one active
electron model. They can be extended for many-electron systems; in this case one defines
differential or total ionization probabilities relative to each electron. However, at the
moment there are no exact time dependent calculations for models with more than two
active electrons.

Besides the numerical integration of the Schrödinger equation, which provides exact
results, some rough informations on the time evolution of the wavepacket can be obtained
from very simple qualitative considerations. The simplest method is to apply the Ehrenfest
theorems neglecting the atomic potential; this approximation is justified as for the high
field intensities involved in calculations its contribution is very small with respect to that
Atomic stabilization: the non-relativistic case 21

of the external field. Then the Ehrenfest theorems predict that the mean coordinate and
velocity of the electron are given by the classical expressions
Zt Zt
e 0 0 e
hri(t) = − A(t )dt , hvi(t) = E(t0 )dt0 (2.28)
m m
ti ti

where A(t) and E(t) are the vector potential and the electric field respectively. From
the equations (2.25) one can easily see that in order to obtain non-negligible survival
probabilities Pn the wavefunction at the moment tf must be localized around the nucleus,
such that the total displacement vanishes
Ztf
e
hri(tf ) = − A(t0 )dt0 ≈ 0 (2.29)
m
ti

Also a rigorous mathematical study was published by Faria [21] who has shown that the
dynamic stabilization can not be obtained unless the total displacement and the drift
velocity of the classical particle at the end of the pulse vanish. This is equivalent to the
conditions on the laser pulse
Ztf Ztf
A(t) dt = 0, E(t) dt = 0. (2.30)
ti ti

The above - very restrictive - conditions have been shown to be obeyed for any pulse
which can be experimentally produced [22]. However, we emphasize here that they are
necessary but not sufficient to obtain the atomic stabilization; as we shall see in the
numerical example in the next Section the frequency, the shape and the length of the
pulse play an important role too.
A physical interpretation of the dynamic stabilization can be given in terms of Floquet
theory. In its simplest version one assumes that at every moment of time the laser
field which can be described by the time-dependent vector potential (A.23) or (A.25)
can be approximated as monochromatic with fixed frequency ω and the “instantaneous
amplitude” A0 (t) = A0 f Ttτ and that the system continously evolves along the Floquet


state originating at zero field from the initial bound state of the atom. A more elaborate
approach was developed by Gavrila who introduced the multistate Floquet theory. It is
based on the assumption that the wavepacket describing an atom interacting with a not
too short laser pulse can be written at any moment of time as a superposition of Floquet
states corresponding to the frequency ω of the pulse and the instantaneous amplitude
A0 (t)

Ψ(r, t) = S c (t)ψ
ν
ν
(ν)
(r, t; A0 (t)) (2.31)

S
The symbol in the previous equation indicates that the summation is performed over the
discrete and continuum part of the Floquet spectrum. Although a rigorous mathematical
22 Atomic stabilization: the non-relativistic case

proof of the completeness of the Floquet solutions was not given yet, the formula (2.31)
was used with good results in model calculations. An exhaustive check of its validity was
done by Wells et al [23] in a one-dimensional model calculation. If in the expansion (2.31)
only one term contributes then the atomic wavefunction can be approximated as
Ψ(r, t) = ψ (in) (r, t; A0 (t)) (2.32)
where ψ (in) (r, t; A0 (t)) is the Floquet state originating from the initial bound state of the
system2 , In this case the evolution is called “adiabatic” and, at any moment of time the
ionization rate of the system is given by the width of the quasienergy of the instantaneous
Floquet state. If the Floquet quasienergy is known for every intermediate instantaneous
amplitude of a given pulse then the total ionization probability can be calculated according
to the formula:
 ∞ 
Z
ad
Pion = exp − Γ(E0 (t))dt (2.33)
−∞

In practice, this formula is used to give a rough estimation of the adiabaticity of the
evolution, by comparison with the value of the ionization probability calculated directly
from the numerical integration of the time-dependent Schrödinger equation. Several nu-
merical experiments have revealed that even relatively short pulses, of about 10 optical
cycles lead to a very good degree of adiabaticity. On the other hand, from the time evo-
lution of the system one can extract the quasienergy widths: if one calculates, during the
flat-top of an adiabatic turned-on pulse the projection of the wave-function of the initial
bound state of the system one must obtain an oscillation with the frequency of the field
modulated by a exponential decay. From the slope of the exponential the ionization rate
Γ = −2= W (0) of the Floquet state can be calculated. Numerical examples obtained
for a three-dimensional Coulomb-potential will be presented in the next Section.

2.2.1 Numerical example


In this section we present a numerical study of the interaction of the ground state Hy-
drogen atom with a laser pulse of circular polarization [24]. Our aim was to study the
observability of dynamic stabilization for ground state Hydrogen interacting with a cir-
cularly polarized laser pulse.
We considered a circularly polarized laser pulse described in the dipole approximation
by the vector potential (A.25)
A(t) = Ax ex + Ay ey , (2.34)
 
t
Ax (t) = A0 f sin(ωt), (2.35)

Ay (t) = Ax (t − T /4) = −A0 f (t − T /4) cos(ωt) (2.36)
2
Note that, however, a Floquet solution is not normalizable, so it can not describe alone a real system.
A small contribution from the continuum must be allways added to the expansion (2.32).
Atomic stabilization: the non-relativistic case 23

for three possible shapes of the pulse - Gaussian, sech and Lorentzian - the corresponding
envelopes being

fG (t) = exp[−(1.177 t/τp )2 ] (2.37)

fsech (t) = sech(1.763 t/τp ) (2.38)

fL (t) = [1 + (1.29 t/τp )2 ]. (2.39)

1 1
τp = 0.25 τp = 1

0.5 0.5
Ay/A0

0 0

-0.5 -0.5

-1 -1
-1 -0.5 0 0.5 1 -1 -0.5 0 0.5 1
1 1
τp = 2 τp = 7

0.5 0.5
Ay/A0

0 0

-0.5 -0.5

-1 -1
-1 -0.5 0 0.5 1 -1 -0.5 0 0.5 1
Ax/A0 Ax/A0

Figure 2.6: Trajectory of A(t) for a Gaussian pulse, for different widths τp , indicated on
graphs. The moment t = 0 is marked with a dot, and the sense of motion by an arrow.

The parameter τp in the above formulae is the pulse full width at half maximum (FWHM)
for A2 ; for each of the above envelopes we have explored a large domain of τp for several
values of the frequency ω, all of them larger than the ionization threshold of the ground
state Hydrogen Eg = 0.5 au. In Fig.2.6 is presented the trajectory of the vector potential
A(t) for a Gaussian pulse; the values of τp in units of T = 2π ω
are written on each
graph. One can see that for very short pulses they have very little resemblance with the
“usual”circular polarization image, but reduce to it if the pulse envelope varies slowly
with respect to ωt. Another important mention is that our choice of the vector potential
implies that the conditions (2.30)

Z∞ Z∞
A(t) dt = 0, E(t) dt = 0 (2.40)
−∞ −∞

are satisfied for any pulse shape and length. We remind here that the pulses obeying
the above conditions have been shown, on one hand, to favor stabilization [21], and, on
24 Atomic stabilization: the non-relativistic case

the other hand, to be the only ones possible to obtain in an experiment (see [22]). The
nominal electric peak value E0 is defined as
E0 = ωA0 . (2.41)
In Fig.2.7 the electric field magnitude is represented as a function of time, in units of
E0 , for two Gaussian pulses of widths τp = 1 and τp = 5. One can see that, even for a
very short pulse, E0 is very close to the real maximum value of the field magnitude, and
that the agreement improves when the pulse gets longer. In the following we shall use
either E0 or the related nominal field intensity I0 = 2E02 in order to characterize the pulse
strength.
1.1 1.1

1 1

0.9 τp = 1 0.9 τp = 5

0.8 0.8
Electric field magnitude / E0

Electric field magnitude / E0

0.7 0.7

0.6 0.6

0.5 0.5

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0
-3 -2 -1 0 1 2 3 -16 -12 -8 -4 0 4 8 12 16
t (c) t (c)

Figure 2.7: Time dependence of the electric field magnitude E(t)/E0 for a Gaussian pulse
and two values of τp marked on each graph.

We have numerically integrated the time dependent Schrödinger equation for a Hy-
drogen atom initially in the ground state, interacting with the laser pulse previously
described; at the end of the pulse the survival probability Psurv was calculated by project-
ing the wave-function on the bound states of the Hydrogen atom. Also the total ionization
probability was calculated as
X
Pion = 1 − Psurv = 1 − |hunlm |Ψ(tf )i|2 . (2.42)
n,l,m

In the previous expression the number of states included in the sum was chosen suffi-
ciently large to provide an accuracy better than 0.1% of Pion ; a typical value for the cases
considered here is n ≤ 30. The numerical integration of TDSE was performed with an
extremely powerful code written by Muller [25], which allows the integration of TDSE for
three-dimensional one-electron systems and arbitrary electromagnetic field, with accuracy
better than 10−4 , even on very modest computers.
Atomic stabilization: the non-relativistic case 25

Our results are presented in Fig.2.8. The ionization probability Pion is represented as
a function of the nominal peak intensity I0 , for the three envelopes previously presented
(Gaussian (2.37) in black, sech (2.38) in red, and Lorentzian (2.39) in blue), for several
values of the pulse width τp ; the four panels correspond to the values of the frequency
ω = 0.65, 2, 4, 8 au. The maximum values of the peak intensity E0 at a given ω was
chosen at the validity limit of the nonrelativistic dipole approximation; this limit was
estimated at Emax = 20 au for ω = 2 in a model one-dimensional calculation by Kylstra
et al [26]; moreover, the relativistic corrections have been shown to scale as E0 /ω. We
have adapted their estimations for our case taking into account the fact that the relevant
parameter here is not the electric field peak value, but rather the peak intensity, which is
defined differently for the linear and circular polarization
I0, circ = 2E02 , I0, lin = E02 . (2.43)
As a consequence, the validity limit of the non-relativistic dipole approximation was taken
as
r
ω I0,max (ω = 2)
E0,max (ω) = , I0,max (ω = 2) = 400 au. (2.44)
2 2
The values of the FWHM τp at each frequency have been chosen such that: a) the resulted
atomic responses cover a domain as large as possible, and b) the values considered are in
the range of accessibility for the laser sources existent or under construction.
All curves in Fig.2.8 have several common features, which will be discussed in the
following. At small peak intensities the ionization probabilities increase very fast with I0 ,
in agreement with the predictions of the perturbation theory. Starting from a certain value
of the intensity I0 , the slope of Pion dramatically decreases, and the ionization probability
tends to level off at a value less than 1 for Lorentzian pulses and even decreases for
Gaussian and sech pulses; this behavior is the evidence of dynamic stabilization of ground
- state Hydrogen atom. In the non-perturbative regime the ionization probabilities at
fixed frequency and pulse width have a very strong dependence on the pulse shape, unlike
in the perturbative domain: Pion has systematically the lowest value for Gaussian pulse,
and the highest one for Lorentzian pulses
G sech L
Pion < Pion < Pion . (2.45)
For fixed frequency and at a given pulse shape Pion increases with τp , eventually getting
indistinguishable from 1. At fixed pulse width τp and for the frequencies ω = 0.65, 2, 4 au
one can see near the upper limit of the intensity the tendency of Pion to increase which is
the onset of the destabilization regime. It should be noted, however that the study of the
atomic destabilization in the nonrelativistic dipole approximation is at most a problem of
academic interest. Relativistic corrections have to be taken into account in order to get
a correct picture of the atomic response at very high field intensities.
We have also studied the possibility to interpret the TDSE results in relation with the
predictions of the multistate Floquet theory. Its simplest version, “single state Floquet
theory” states that for a pulse sufficiently long and smooth the atom can be described
26 Atomic stabilization: the non-relativistic case

1 1
10
2

0.9 0.9
5

0.8 0.8
1

0.7 0.7 3
Ionization probability

Ionization probability
0.6 0.6

0.25
0.5 0.5

0.4 0.4

0.3 0.3

0.2 ω = 0.65 au 0.2


0.25 ω = 2 au

0.1 0.1

0 0
0 1 2 3 4 5 6 7 8 9 10 0 5 10 15 20
1/2 1/2
I0 (au) I0 (au)

1 1

0.9 25 0.9 100


ω = 8 au

0.8 0.8
50

0.7 10 0.7
Ionization probability

Ionization probability

0.6 0.6

25
0.5 0.5

5
0.4 0.4

0.3 0.3
10

1
0.2 0.2

0.25 5
0.1 0.1
1
ω = 4 au 0.25

0 0
0 10 20 30 40 0 10 20 30 40 50 60 70 80
1/2 1/2
I0 (au) I0 (au)

Figure 2.8: Ionization probability Pion of ground-state


√ Hydrogen for various envelopes and
1/2
pulse widths τp (in cycles), as a function of I0 = 2E0 (in au). Black curves - Gaussian
pulses; red curves - sech pulses; blue curves - Lorentzian pulses. Each panel corresponds
to a fixed value of the frequency ω; the value of τp is specified next to the Pion curves
coalescent at small I0 .
Atomic stabilization: the non-relativistic case 27

at any moment of time by a single Floquet state. This property gives us a very accurate
method - although extremely time consuming - to calculate the Floquet quasienergies.
In the case of an atom initially in a bound state u0 acted upon by a pulse with a very
smooth turn-on followed by a flat top the projection of the wave-function on u0 during
the flat top of the pulse will have the form
 
i
hu0 |ψ(t)i = exp − W t φ(t) (2.46)
~

where φ(t) is a periodic function with the period T of the field, and W is the complex
quasienergy of the Floquet state which is continuously connected with u0 . If one solves the
TDSE for this case it is possible to extract the quasienergy W from the numerical values
of hu0 |ψ(t)i. This method is extremely time consuming but it allows the calculation
of the quasienergies for intensities domains which are inaccessible with other methods.
In the table 2.1 is presented the ground state quasienergy for the Hydrogen atom, at
the frequency ω = 2 au for different intensities. Also the values given by the Floquet
code of Potvliege [27] are presented; one notices the very good agreement which proves
the correctness of the single state Floquet theory. In Fig.2.9 the ionization rate Γ is

α0 Re(W ) Re(WP ) Γ ΓP
0.5 -0.44597 -0.44597 0.06514 0.06518
1 -0.34138 -0.34141 0.05585 0.05590
2 -0.25335 -0.25333 0.005815 0.005829
3 -0.20176 -0.20175 0.002413 0.002424
5 -0.14372 – 0.0004356 –
8 -0.10103 – 0.0001069 –

Table 2.1: Hydrogen quasienergies W = E − 2i Γ in a circularly polarized field at ω = 2


au and various α0 = Eω20 , according to our TDSE calculation and the Floquet code of
Potvliege. The latter values bear the subscript P.

represented (in black line) as a function of the field intensity I0 ; also the LOPT ionization
rate is represented (blue line) and the ionization rate for linearly polarized radiation (red
line); the latter have been calculated by Dondera et al [28]. Note that at low intensity
all the three curves coincide, in agreement with the perturbative prediction. At larger
intensities the LOPT prediction diverges while the result of the Floquet calculations
present stabilization. It is also worth noticing that the ionization rates for circular and
linear polarization are quite similar, when represented as a function of I0 .
Next, we have tested the adiabaticity of the evolution for the pulses previously studied.
At the frequency ω = 2 au we have compared the exact ionization probability calculated
numerically for the pulses (2.37 - 2.39) with the adiabatic approximation (2.33); the results
are presented in Fig.2.10. One can see that the agreement, which is an indication of the
adiabaticity of the evolution, is surprisingly good even for pulses of FWHM as small as
28 Atomic stabilization: the non-relativistic case

0.09

0.08 LOPT

0.07
ω = 2 au

0.06

Γ (au) 0.05

0.04

0.03

0.02

0.01

0
0 2 4 6 8 10 12 14
1/2
I0 (au)

Figure 2.9: Black curve: ionization rate Γ (in au) for ground - state Hydrogen and circular
1/2
polarization at ω = 2 au, as a function of I0 , calculated from the TDSE solutions. Red
curve: Γ for linear polarization calculated by Dondera et al. Blue curve: Γ for the LOPT
one-photon ionization rate.

1 cycle. As a further test, we have compared the ionization probabilities in the circular
and linear case for a sech pulse with ω = 2 and various lengths. The results, represented
again as a function of I0 are very similar. To this similarity, two factors contribute: on
the one hand, because the evolution is almost adiabatic, the approximate formula (2.33)
should work in both cases; on the other hand, we saw that the ionization rates Γ in the
linear and circular polarization are not too different, so the results of (2.33) are also very
close to each other. This latter observation comes to further support the physical reality
of the single state Floquet theory.
Atomic stabilization: the non-relativistic case 29

1
**
10 *
10
10

0.95
**
5

0.9 *
5

0.85 5

Ionization probability
**
3
0.8
ω = 2 au

0.75 *
3

0.7
3

0.65

0.6
0 5 10 15 20
1/2
I0 (au)

Figure 2.10: Comparison of Pion of ground - state Hydrogen, calculated from TDSE (solid
curves) and in the adiabatic approximation (dashed curves) for various envelopes and
pulse durations. The values of τp are indicated. The τp for Gaussian pulses carry no
asterisk, for sech pulses they carry one asterisk, and for Lorentzian pulses, two asterisks.

10

0.9

0.8

0.7 3
Ionization probability

0.6

0.5

0.4 1

0.25
0.3

0.2 ω = 2 au

0.1

0
0 5 10 15 20
1/2
I0 (au)

Figure 2.11: Comparison of Pion for circular and linear polarizations; sech pulses are
considered ω = 2 au and the indicated widths τp . Black curves: circular polarization; red
curves: linear polarization.
30 Atomic stabilization: the non-relativistic case

2.3 References
[1] M. Gavrila, J. Phys. B 35, R147 (2002)

[2] M. Gavrila, in Atoms in Intense Laser Fields, Academic Press 1992.

[3] W. Pauli and M. Fierz, Nuovo Cim. 15, 167 (1938).

[4] H. A. Kramers, Collected scientific papers. Notrth-Holland, Amsterdam 1956.

[5] W. C. Henneberger, Phys. Rev. Lett 21, 838 (1968).

[6] F. H. Faisal, J. Phys. B 6, L89 (1973).

[7] J. H. Shirley, Phys. Rev. B 138, 979 (1965).

[8] M. Gavrila and J. Z. Kaminski, Phys. Rev. Lett. 52, 614 (1984).

[9] I. Simbotin, M. Stroe, and M. Gavrila, Laser Phys. 14, 482 (2004)

[10] M. Stroe, PhD thesis, Bucharest 2004.

[11] E. Teller, J. Phys. Chem. 41, 109 (1937).

[12] R. M. Potvliege and R. Shakeshaft, Phys. Rev. A 40, 3061 (1989).

[13] C. Chirila, M. Boca, V. Dinu and V. Florescu, Eur. Phys. J. D 27, 15 (2003).

[14] A.A. Frost, J. Chem. Phys. 25, 1150 (1956).

[15] I. Richard Lapidus, Am. J. Phys. 38, 905 (1970).

[16] A. Galindo and P. Pascual, Quantum Mechanics, vol. I (Springer Verlag, Berlin, 1990)

[17] For a detailed discussion see Ciprian Chirila, Master thesis, Bucharest, 2003.

[18] M. Boca, C. Chirila, M. Stroe, and V. Florescu, Phys. Lett. A 286, 410 (2001).

[19] M. P. de Boer, J. H. Hoogenraad, R. B. Vrijen, R. C. Constantinescu, L. D. Noordam


and H. G. Muller, Phys. Rev. Lett. 71, 3263 (1993), M. P. de Boer, J. H. Hoogenraad,
R. B. Vrijen, R. C. Constantinescu, L. D. Noordam and H. G. Muller, Phys. Rev. A 50,
4085 (1994).

[20] N. J. van Druten, R. Constantinescu, J. M. Schins, H. Nieuwenhuize and H. G. Muller,


Phys. Rev. A 55, 622 (1997).

[21] C. F. M. Faria, A. Fring and R. Schrader, Laser Phys. 9, 379 (1999)

[22] M. Gavrila, Multiphoton processes (AIP Conf. Proc. vol. 525), ed. I. Di Mauro, R. R.
Freeman and K. C. Kullander, (New-York, American Institute of Physics) p 107.

[23] J. C. Wells, I. Simbotin and M. Gavrila, Phys. Rev. A 56, 3961 (1997).
Atomic stabilization: the non-relativistic case 31

[24] M. Boca, H. G. Muller and M. Gavrila, J. Phys. B 37, 147 (2004).

[25] H. G. Muller, Laser Phys. 9, 138 (1999).

[26] N. J. Kylstra, R. A. Worthington, A. Patel, P. L. Knight, J. R. Vazquez de Aldana and


L. Roso, Phys. Rev. Lett. 85, 1835 (2000).

[27] R. M. Potvliege, Comput. Phys. Commun. 114, 42 (1998).

[28] M. Dondera, H. G. Muller and M. Gavrila, Laser Phys. 12, 415 (2002)
Chapter 3

The Volkov solutions

The main goal of the Thesis is the study of the interaction of one-electron atoms with very
intense laser pulses in the relativistic regime. A first step required in order to accomplish
this is to review the interaction of free electrons with supraintense laser pulses. As in this
case the electron acquires relativistic velocities the Schrödinger equation is no more valid,
and the dynamics of the system is governed by the Dirac equation. For the case of plane
wave laser pulses its solutions can be written in closed analytical form, were obtained for
the first time in 1935 by D. M. Volkov [1] and are known as Volkov solutions.
As we shall see in the following Chapter the Volkov solutions constitute the main
building blocks of the formalism developed for the study of the atom-laser interaction in
the relativistic regime. As a consequence a detailed study of their properties is needed
and it constitutes the topic of this Chapter. We present their derivation in Section 3.1,
the most important properties in Section 3.2 and, finally, an alternative expression in the
last Section of this Chapter.
As discussed in Appendix A a plane wave laser pulse propagating along the n direction
can be described by the vector potential
A (x) ≡ (0, A(ct − n · r)) , A(ct − n · r) ≡ A(φ), A · n = 0. (3.1)
In the following we shall choose the coordinate system such that n is along the Oz di-
rection. Although in numerical applications only the linear polarization case will be
considered, the Volkov solutions are presented for the case of arbitrary polarization. The
Dirac equation for a particle of mass m and electric charge e in the external field A is
∂ψ(x) 
= cα · (P − eA(φ)) + mc2 β ψ(x),

i~ (3.2)
∂t
where α and β are the Dirac matrices (A.8). The Volkov bispinors are four linearly
independent solutions of the above equation defined by the condition that in the zero
field limit they reduce to the plane waves solutions (B.2) of the free Dirac equation (B.1).
Their expression is
 
i
ψi (x; p) = exp − i (x · p) + Λi (φ; p) Ωi (φ; p) ξi (p), i = 1, . . . , 4 (3.3)
~

33
34 The Volkov solutions

where i is the sign factor (B.4) equal to 1 for i ∈ {1, 2} and to −1 for i ∈ {3, 4}, the
four-vector n is defined as n = (1, n) (A.11), ξi (p) are the bispinors defined in Eq.(B.2),


 
i 1
2eA(φ0 ) · p − i e2 A2 (φ0 ) dφ0 
 
Λi (φ; p) =  (3.4)
~ 2(n · p)

and
 
e
Ωi (φ; p) = 1 − i (1 + α · n) (α · A(φ)) . (3.5)
2(n · p)
In the previous equations a · b is the notation for the four product defined in (A.3), p is a
four-vector of norm mc; in the zero field limit ψi (x; p) reduce to the free particle bispinors
χi (x; p) (B.2) and p is the particle four-momentum.

3.1 Calculation of Volkov solutions


We present in this section the derivation of the Volkov solution (3.3). We shall look for
the solutions ψi of the equation (3.2) of the form
   
∂ 2 i
ψi (x; p) = i~ + cα · (P − eA(φ)) + mc β exp − i (p · x) Φi (φ; p),
∂t ~
i = 1, . . . , 4 (3.6)
The equation obeyed by Φi (φ; p)
   
d 2 2 dA(φ)
i ~(n · p) − e A (φ) − 2i eA(φ) · p + i~e (1 − α · n) α · Φi (φ; p) = 0
dφ dφ
(3.7)
can be easily solved, the solution being
 
e  
Φi (φ; p) = 1 + i (1 − α · n) (α · A(φ)) Zi exp Λi (φ; p) (3.8)
2(n · p)
where Λi (x; p) given in Eq.(3.4) and Zi is an arbitrary constant bispinor. If one uses
the condition that for A(φ) = 0 the Volkov solutions ψi (x; p) reduce to the free particle
bispinors (B.2) one finds that Zi must be chosen equal to ζi defined in (B.3). Finally,
using (3.4) and (3.8) in (3.6) one obtains by direct calculation the result (3.3).
As we shall see, for some applications it is convenient to use an equivalent more
compact expression of the Volkov solutions; after some straight algebraic manipulations
one finds that the bispinors Ωi (φ; p)ξi (p) can be written as
" #
p̂ − i eÂ(φ) + i mc
Ωi (φ; p)ξi (p) = n̂ξi (p); (3.9)
2(n · p)
The Volkov solutions 35

in the previous equation the notation â ≡ a · γ = γ 0 a0 − γ · a defined in the (A.9) was


used. It will be also useful an equivalent expression of the phase (3.4),
i
Fi (x; p) ≡ − i (x · p) + Λi (φ; p) (3.10)
~

 
i  1 1
(eA(φ0 ) − i p⊥ )2 + m2 c2 dφ0 − (r · n + ct)(n · p) .
 
= i p⊥ · r⊥ −
~ 2(n · p) 2

Due to the one-to-one correspondence between the Volkov solutions and the free Dirac
bispinors ui (x, p) ψ1,2 (x; p) will be briefly called “positive energy Volkov bispinors” and
ψ3,4 (x; p) “negative energy Volkov bispinors”, but we must keep in mind that at non-zero
field intensity the Volkov solutions are not eigenvectors of the projectors (B.7) on the
positive/negative energy subspaces.

3.2 Properties of Volkov solutions


One can easily check
 that the above
 defined Volkov solutionsare eigenvectors of the
∂ ∂ ∂
operators P⊥ ≡ −i~ ex ∂x + ey ∂y and (n · P ) ≡ i~ ∂(ct) + ∂x∂ 3 :

P⊥ ψi (x; p) = i p⊥ ψi (x; p), (3.11)

(n · P )ψi (x; p) = i (n · p)ψi (x; p). (3.12)

An interesting question regarding the Volkov solutions is whether they form an or-
thonormal basis, i.e. whether they obey the orthogonality
Z
hψi (x; p )|ψj (x; p)i ≡ dr ψi+ (x; p)ψj (x; p0 ) = δij δ(p − p0 )
0
(3.13)
R3

and completeness relations


XZ
dp ψi (t, r; p) ψi+ (t, r0 ; p) = δ(r − r0 ) I4 . (3.14)
i=1,4
R3

Note that in the previous equation the dependence on the four vector x was explicitly
written as x → (t, r) in order to emphasize the fact that both ψi (t, r; p) and ψi+ (t, r0 ; p)
are taken at the same moment of time.
The issue of orthogonality and completeness was largely disputed in the literature, a
complete answer being only recently given. A fist attempt was made by Ritus in 1972 [2]
who gave a different form of orthogonality and completeness
Z Z
d x ψi (x; p)ψi (x; p ) = δ (p − p ), d4 p ψi (x; p)ψi+ (x0 ; p) = δ (4) (x − x0 ). (3.15)
4 + 0 (4) 0

R4 R4
36 The Volkov solutions

The first proof of the orthogonality relation in the form (3.13) was given by Ritus in 1979
[3] and rederived by Filipowicz [4] in 1985 for the case of positive energy Volkov solutions.
Recently a rigorous mathematical proof was presented by Zakowicz [5].
The completeness problem was also addressed by Bergou and Varró [6] who obtained
a completeness relation along the hyperplane (n · x) = const. The validity of Eq.(3.14)
was stated several times in the literature, but no direct proof was given up to now.
Since the orthogonality and completeness relations in the sense (3.13), (3.14) are es-
sential for the formalism developed in the next Chapter, the following subsections are
dedicated to their proofs. In Section 3.2.1 we briefly present the derivation given by Fil-
ipowicz for the relation (3.13), generalized for the case when both positive and negative
energy Volkov solutions are considered. In Section 3.2.2 we present a direct calculation
for (3.14) which to the best of our knowledge is a new result.

3.2.1 Proof of the orthogonality of the Volkov solutions


The goal of the calculation is to prove the validity of the equation

hψi (x; p0 )|ψj (x; p)i = δij δ(p − p0 ) (3.16)

Using Eqs.(3.3), (3.9), and (3.10) we have


 
Z
Fi∗ (x;p0 )+Fj (x;p) 
hψi (x; p0 ) | ψj (x; p)i = hξi (p0 ) |  dr Ω+ 0
i (φ; p )Ωj (φ; p) e | ξj (p)i. (3.17)
R3

In the previous equation the integral over r⊥ can be directly calculated and its value is
(2π~)2 δ(i p⊥ − j p0⊥ ). Then one obtains

hψi (x; p0 ) | ψj (x; p)i = (2π~)2 δ(i p⊥ − j p0⊥ ) ×


  ∞  
 Z 
0 + 0 Qij (z;p,p0 ) 
× hξi (p ) |  dz Ωi (φ; p )Ωj (φ; p) e | ξj (p)i , (3.18)
 
−∞ p⊥ =i j p0⊥

with
i 
Qij (z; p, p0 ) = − i (n · p0 ) − j (n · p) ×

(3.19)
 ~ 
1 Zφ
i j  2 2 2
 
× (z + ct) − (p⊥ − i eA(φ0 )) + m c dφ0 .
2 2(n · p)(n · p0 ) 

Using the identities

n̂p̂ = n̂p̂⊥ + n̂γ 0 (n · p), p̂0 n̂ = p̂0 ⊥ n̂ + γ 0 n̂(n · p0 ) (3.20)


The Volkov solutions 37

and Eq.(3.9) one can show that the matrix under the integral over z is
!
0 1 i j [p⊥ − i eA(φ)]2 + m2 c2
Ω+
i (φ; p )Ωj (φ; p) = 1+ (γ 0 n̂)2 .
p⊥ =i j p0⊥ 4 (n · p)(n · p0 )
(3.21)

The next step is to perform in the integral the change of variable


 
1 i j
(p⊥ − i eA(φ0 ))2 + m2 c2 dφ0  .
 
ρ= z + ct − (3.22)
2 (n · p)(n · p0 )

As the Jacobian of the transformation


!
dρ 1 i j [p⊥ − i eA(φ)]2 + m2 c2
= 1+ (3.23)
dz 2 (n · p)(n · p0 )

is already formed in the expression (3.21) of Ω+ 0


i (φ; p )Ωj (φ; p) the remaining integral over
ρ can be easily calculated, final result being

hψi (x; p0 )|ψj (x; p)i = (2π~)3 δ(i p⊥ − j p0⊥ ) ×


× δ(i (n · p0 ) − j (n · p)) hξi (p0 )|(γ 0 n̂)2 |ξj (p)i. (3.24)

As the four products (n · p) and (n · p0 ) are always positive, the second δ− function if
the previous equation vanishes unless i, j ∈ {1, 2} or i, j ∈ {3, 4}, i.e. the inner product
between a positive energy Volkov solution and a negative energy one is zero. Further,
using the relations

E
δ((n · p0 ) − (n · p)) = δ(pk − p0k ) (3.25)
c(n · p)

and

1 2c(n · p)
hξi (p)|(γ 0 n̂)2 |ξj (p)i = δ ij , i, j ∈ {1, 2} or i, j ∈ {3, 4} (3.26)
(2π~)3 E

one obtains the desired relation

hψi (x; p0 )|ψj (x; p)i = δij δ(p − p0 ) (3.27)

which ends the proof of orthogonality and normalization of the Volkov solutions.
38 The Volkov solutions

3.2.2 Proof of the completeness of the Volkov solutions


We present a proof based on direct calculation of the fact that the Volkov solutions
normalized in the momentum scale (3.3) form a complete set at any moment of time, i.e.
they obey the equation
XZ
0
C(r, r , t) ≡ dp ψi (r, t; p) ψi+ (r0 , t; p) = δ(r − r0 ) I4 . (3.28)
i=1,4
R3

As in Eq.(3.14), we have explicitly indicated the fact that all the Volkov spinors in the
previous equation are taken at the same moment of time.
Our approach, based on the direct evaluation of the integrals in the expression of the
operator C(r, r0 , t) starts with a changes of variable in (3.28)
{p⊥ , pk } −→ {i p⊥ , (n · p)}. (3.29)
Using Eqs.(3.9) and (3.10) in the expression of Volkov solutions, after lengthy but straight
algebraic manipulations one finds that the operator C(r, r0 , t) can be written as
Z
0 1 i 0
C(r, r , t) = 3
dp⊥ e ~ p⊥ ·(r⊥ −r⊥ ) × (3.30)
(2π~)
R2
 
1 0 0 0
× Γ0 (p⊥ , φ, φ )C0 (a, b) + Γ−1 (p⊥ , φ, φ )S−1 (a, b) + Γ−2 (p⊥ , φ, φ )C−2 (a, b) .
2
Here C0,−2 (a, b) and S−1 (a, b) are defined as
Z∞   Z∞  
b b
Cn (a, b) = dv v n cos av − , Sn (a, b) = n
dv v sin av − . (3.31)
v v
0 0

with the parameters


z − z0
a= , (3.32)
2~
Zφ0
1
dφ0 (eA(φ0 ) − p⊥ )2 + m2 c2 ,
 
b= (3.33)
2~
φ

and the three matrices Γi , i = 1, . . . , 3 are given by:


Γ0 (p⊥ , φ, φ0 ) = 1 − α3 (3.34)
nh  i   o
0 0 0
Γ−1 (p⊥ , φ, φ ) = −i 2(p̂⊥ + mc) − e Â(φ) + Â(φ ) γ + e Â(φ) − Â(φ ) γ 3
0

(3.35)
h  
Γ−2 (p⊥ , φ, φ0 ) = p2⊥ + m2 c2 + e Â(φ)p̂⊥ + p̂⊥ Â(φ0 ) − e2 Â(φ)Â(φ0 )+
 i
+ mc eÂ(φ0 ) − eÂ(φ) (1 + α3 ). (3.36)
The Volkov solutions 39

One can prove that the integrals C0,−2 (a, b) and S−1 (a, b) for real values of parameters a
and b have the expressions
 
p
S−1 (a, b) = πJ0 (2 |a| |b|) θ(a)θ(−b) − θ(−a)θ(b) , (3.37)

s  
|b| 0 p
C0 (a, b) = πδ(a) + π J (2 |a| |b|) θ(a)θ(−b) + θ(−a)θ(b) , (3.38)
|a| 0

s  
|a| 0 p
C−2 (a, b) = πδ(b) + π J (2 |a| |b|) θ(a)θ(−b) + θ(−a)θ(b) . (3.39)
|b| 0

In the above relations θ is the Heaviside function and J0 is the Bessel function of order
0. From (3.32,3.33) on can see that in our case the parameters a and b have always the
same sign; as a consequence Eqs.(3.37-3.39) become

S−1 (a, b) = 0, C0 (a, b) = πδ(a), C−2 (a, b) = πδ(b). (3.40)

Next, we use the identities

z − z0
 
δ = 2~δ(z − z 0 ), (3.41)
2~

Zφ0
 
1 2~
dφ0 (eA(φ0 ) − p⊥ )2 + m2 c2  = δ(z − z 0 ) (3.42)
 
δ
2~ (eA(φ) − p⊥ )2 + m2 c2
φ

and

Γ−2 (p⊥ , φ, φ0 ) = (eA(φ) − p⊥ )2 + m2 c2 (1 + α3 ).


 
(3.43)
z=z 0

With these one gets


Z  
0 1 i
p ·(r⊥ −r0⊥ ) 1 1
C(r, r , t) = dp⊥ e ~ ⊥ (1 + α3 ) + (1 − α3 ) δ(z − z 0 ) (3.44)
(2π~)2 2 2
R2

or

C(r, r0 , t) = δ(r − r0 ) (3.45)

which is the desired identity.


40 The Volkov solutions

Derivation of Eqs.(3.37-3.39)
One defines the integrals

Z∞   Z∞  
b b
Cn (a, b) = dv v n cos av − , Sn (a, b) = n
dv v sin av − . (3.46)
v v
0 0

For real a and b, S−1 (a, b) is calculated using the equations (3.868.1) and (3.868.3) of
[7]
 
p
S−1 (a, b) = πJ0 (2 |a| |b|) θ(a)θ(−b) − θ(−a)θ(b) . (3.47)

Starting from this result one obtains the expressions of C0 (a, b) and C−2 (a, b) by derivation
with respect to a or b of S−1 (a, b),
s  
|b| 0 p
C0 (a, b) = πδ(a) + π J (2 |a| |b|) θ(a)θ(−b) + θ(−a)θ(b) , (3.48)
|a| 0
s  
|a| 0 p
C−2 (a, b) = πδ(b) + π J (2 |a| |b|) θ(a)θ(−b) + θ(−a)θ(b) . (3.49)
|b| 0

The integrals S−1 and C0,−2 obey the symmetry relations

S−1 (a, b) = −S−1 (−a, −b), (3.50)


C0,−2 (a, b) = C0,−2 (−a, −b), (3.51)
C0 (a, b) = C−2 (b, a). (3.52)

3.3 Other Volkov solutions


We present in this Section an expression of Volkov solutions which is slightly different
from the usual one (3.3). A set of solutions written in this form was used by Krstic and
Mittleman [8, 9] in their study of the Dirac equation for an atom in a monochromatic
external electromagnetic field.
One defines the four vector
µ
p∗ = p + n, (3.53)
(n · p)

with µ an arbitrary positive constant. One can easily prove that the four-vector p∗ has
the properties:

(n · p) = (n · p∗ ), (A(φ) · p) = (A(φ) · p∗ ), (p∗ · p∗ ) = (mc)2 + 2µ (3.54)


The Volkov solutions 41

One notices that p∗ can be considered the four momentum of a fictitious particle of mass
r

m∗ = m2 + 2 (3.55)
c
With these notation, one defines
 
∗ ∗ µ
ψ̃i (x; p ) = ψi x; p − . (3.56)
(n · p∗ )

After some straight algebraic manipulations one finds that these “translated” Volkov
solutions can be expressed as
   
∗ i ∗ ∗ ∗ ∗ µ
ψ̃i (x; p ) = exp − i (x · p ) + Λ̃i (φ; p ) Ωi (φ; p ) ξi p − (3.57)
~ (n · p∗ )

with

 
i  1
Λ̃i (φ; p∗ ) = 2eA(φ0 ) · p∗ − i e2 A2 (φ0 ) − 2µ dφ0  .
 
(3.58)
~ 2(n · p∗ )

The most striking property of the new form of the Volkov solutions ψ̃i (x; p∗ ) is that
although they obey the Dirac equation for the real electron, of mass m, in the exponent
in Eq.(3.57) appears the four-momentum p∗ of the fictive particle of mass m∗ . We also
note that ψ̃i (x; p∗ ) are not normalized in the scale of p∗ , but in the scale of the on-shell
momentum p; the p∗ scale normalization is obtained if their definition is changed as
s
µ
p∗0 − (n·p ∗)
ψ̃i (x; p∗ ) −→ ψ̃i (x; p∗ ) (3.59)
p∗0

Other solutions of the Dirac equation for an electron in an external electromagnetic


field are given by Bergou and Varró [6] and Bagrov and Gitman [10]. Note that also
studies of the Volkov solutions for the quantized field have been published; they will not
be discussed here.

3.4 References
[1] D. M. Volkov, Z. Phys 94, 250 (1935).

[2] V. I. Ritus, Ann. Phys 69, 555 (1972).

[3] V. I. Ritus, Trudy FIAN 111, 5 (1979).

[4] Piotr Filipowicz, J. Phys. A 18, 1675 (1985).


42 The Volkov solutions

[5] S. Zakowics, J. Math. Phys 46, 032304 (2005).

[6] J. Bergou and S. Varró, J. Phys A 13, 2823 (1980).

[7] I. S. Gradstein and I. M. Rijik, Tables of Integrals, Sums, Series and Products, Moscow
1962 (in Russian)

[8] P. S. Krstic and M. H. Mittleman, Phys. Rev. A 42, 4037 (1990)

[9] P. S. Krstic and M. H. Mittleman, Phys. Rev. A 45, 6514 (1992)

[10] V. Bagrov and A. Gitman, arXiv:hep-th/0502106.


Chapter 4

Relativistic generalization of the


space translation method

4.1 The general framework


We shall consider an one-electron atom whose nucleus is fixed in the origin of the reference
frame, with the atomic potential V (r), acted upon by an external plane wave laser pulse of
arbitrary polarization. The external electromagnetic field propagating along the direction
n is described in the Lorentz gauge by the four potential A(φ) ≡ (0; A(ct − n · r)), as
discussed in the Appendix A, Eq.(A.15). In the frame of the single-electron Dirac equation
∂Φ 
= cα · (P − eA(φ)) + mc2 β + V (r) Φ

i~ (4.1)
∂t
a series of high energy phenomena involving electrons of any speed, photons of any energy
and atomic potentials of arbitrary strength can be treated. However, we must note that
the numerical solving of Dirac equation is an extremely difficult problem even for model
atomic systems with reduced dimensionality. The purpose of this Chapter is to develop a
formalism allowing the inclusion of the relativistic effects in numerical study of atom-laser
interaction. The method is applicable only for the case of atomic phenomena deriving
from low energy initial conditions like: light atom structure in superintense optical fields,
multiphoton ionization and low energy free-free transitions in such fields. Another class of
phenomena, taking place for high energy initial conditions and/or heavy atoms such as fast
electron - heavy nucleus collision in intense laser field, resulting in emission/absorption of
high energy photons or pair creation, can not be described in the frame of the formalism
presented here, and they shall not be discussed in the Thesis.
We consider the case in which at the initial moment, chosen very far in the past,
when the laser pulse is infinitely far from the origin, the electron is in a bound state or
a superposition of low energy continuum states. Due to the interaction with the super
intense plane wave laser field the electron can acquire relativistic velocities, but after the
pulse has passed it slows down again at non-relativistic velocities and the atom is left in
a superposition of low energy states.

43
44 Relativistic generalization of the space translation method

Our approach is to obtain a generalization of the non-relativistic space translation


method (see its definition in Section 2.1), i.e. to find a unitary transformation leading
to a Dirac equation that contains the electromagnetic field only through a transformed
potential. Such a transformation is

Φ = T Ψ, T = S |ψ ihφ |
i
i i (4.2)

where ψi is a set of solutions of the Dirac equation for the particle in the electromagnetic
field
∂ψi 
= cα · (P − eA(φ)) + mc2 β ψi

i~ (4.3)
∂t
and φi is a set of solutions corresponding to a field independent Hamiltonian H0
∂φi
i~ = H0 φi . (4.4)
∂t
The conditions imposed on ψi and φi are to be indexed with the same set of quantum
numbers denoted here by {i} and to form complete and orthogonal sets. In this case the
transformation operator T is unitary and the equation obeyed by Ψ is
∂Ψ
i~ = [H0 + V 0 ] Ψ, V 0 = T +V T (4.5)
∂t
the entire dependence on the field being contained in the transformed potential V 0 .
This approach has been used by Kaminski [1] in relation to the laser assisted electron-
atom scattering problem. Also Krstic and Mittleman [2, 3] have performed a similar
transformation of the Dirac equation for an atom in a monochromatic external field
π
A = A0 [ex sin(ωt − kz) + tan δ ey cos(ωt − kz)] , δ = 0, (4.6)
4
in order to study the relativistic corrections to the Floquet problem. As the
 set ψi they
eA0 2
have used the translated Volkov solutions defined in (3.56) with µ = 2 and H0 was
chosen the free Dirac Hamiltonian in the standard representation for a fictitious particle
of mass
r
e2 A20
m∗ = m2 + . (4.7)
2c2
Note that their choice leads to the transformed equation
∂Ψ 
= cα · P + m∗ c2 β + V 0 Ψ

i~ (4.8)
∂t
which explicitly contains the dressed mass m∗ discussed in the Appendix C, Eqs.(C.42),
(C.45).
Similar generalizations of the space-translation method transformation have been pro-
posed by Protopapas et al [4] for the one-dimensional Dirac equation and by Førre et al
[5] for the non-dipole Schrödinger equation.
Relativistic generalization of the space translation method 45

Our method is very similar to that used by Krstic and Mittleman; however, it is ap-
plicable to the case of a laser pulse instead of the monochromatic radiation. We choose
the set ψi (x; p) as the set of the standard Volkov solutions (3.3) and for H0 the free Dirac
Hamiltonian in the Foldy-Wouthuysen representation [6]

H0 = c2 P2 + m2 c4 β (4.9)
with the set of solutions vi (x; p) given by Eq.(B.9). The equation for the new wavespinor
Ψ is
∂Ψ h√ 2 2 i
i~ = c P + m2 c4 β + T + V T Ψ (4.10)
∂t
with
X Z
0
hr|T |r i = dp ψi (x, p)vi+ (x0 , p). (4.11)
i=1,4
R3

In the sense discussed above the transformation (4.2) is a relativistic generalization of the
spatial translation used in the non-relativistic dipole approximation, and Eq.(4.10) will be
named the generalized space translated Dirac equation; the picture defined by the operator
T will be called the generalized translated Dirac picture. In the previous equation x and x0
are the coordinate four vectors x ≡ (ct, p r) and x0 ≡ (ct, r0 ) taken at the same moment and
p is the onshell four momentum p ≡ ( m2 c2 + p2 , p) ≡ ( Ec , p). Since both the Volkov
solutions ψi (x, p) and the free solutions vi (x, p) form orthogonal and normalized basis sets
the operator T is unitary. One notices that at a time chosen very far in the past T reduces
to the unitary operator S defined in Appendix B, Eq.(B.10), such that (4.10) becomes
∂Ψ h√ 2 2 i
i~ = c P + m2 c4 β + S + V S Ψ, t −→ −∞. (4.12)
∂t

4.2 Properties of the transformation operator T


For further calculations it is convenient to separate the contribution of the positive and
negative frequency solutions in T . After replacing in Eq.(4.11) the Volkov and the free
solutions by their explicit form one obtains
hr|T |r0 i = hr|T+ |r0 i + hr|T− |r0 i, (4.13)
Z
0 1
dp t± φ, p exp f± r, r0 , t; p
  
hr|T± |r i = 3
(4.14)
(2π~)
R3

with
 

i  1 
r, r0 , t; p = ±p · (r − r0 ) + dχ 2eA(χ) · p ∓ e2 A2 (χ)
 
f± (4.15)
~ 2(n · p) 
−∞
46 Relativistic generalization of the space translation method

and
(1 + α · n) α · (eA(φ)) (E + cα · p + mc2 ) 1 ± β
 

t± φ, p = 1∓ p =
2(n · p) 2E(E + mc2 ) 2
(E + cα · p + mc2 ) 1 ± β
= Ωi (φ; p) p . (4.16)
2E(E + mc2 ) 2

In the previous equation Ωi (φ; p) is the matrix defined in Eq.(3.5) and i must be 1 or 2
for t+ and 3 or 4 for t− One notices that the two matrices t± (φ, p) in T have only the
first/last two columns non-zero. Their explicit form is
 
2 eA·cp eA×cp+(E+mc2 )n×eA
E + mc − 2(n·p) − iσ · 2(n·p)
0
 1  
t+ φ, p = p  ,
2E(E + mc2 )  h i 
i(n×eA)·cp eA(E+mc2 )−cp×(eA×n)
− 2(n·p) + σ · cp + 2(n·p)
0
(4.17)

 h i 
i(n×eA)·cp eA(E+mc2 )−cp×(eA×n)
0 2(n·p)
+ σ · cp − 2(n·p)
 1  
t− φ, p = p  ;
2E(E + mc2 )  eA·cp eA×cp+(E+mc2 )n×eA

0 E + mc2 + 2(n·p)
+ iσ · 2(n·p)

(4.18)

in the previous equations argument of the vector potential A is φ = ct − z.


The adjoint of the operator T is

hr|T + |r0 i = hr|T++ |r0 i + hr|T−+ |r0 i, (4.19)

where
Z
1 ∗T
hr|T±+ |r0 i dp t± φ0 , p exp −f± r0 , r, t; p ;
 
= (4.20)
(2π~)3
R3

and (t± )∗T is the conjugate transpose of the matrix t± .


As both {ψi } and {vi } form complete and orthonormalized sets the operator T is
unitary. Using the orthogonality and the normalization properties of the Volkov solutions
{ψi } and the completeness of the set of free bispinors {vi } one can write the identity
X Z Z Z
00 0
+
hr |T T | r i = dr dp dp0 vi (x, p)ψi+ (x0 , p)ψj (x0 , p0 )vj+ (x00 , p0 )
i,j=1,4 3
R R3 R3
X Z
= dp vi (x, p)vi+ (x00 , p) = δ(r − r00 )I4 . (4.21)
i=1,4
R3
Relativistic generalization of the space translation method 47

Similarly, T T + can be calculated using the completeness of the Volkov solutions and the
orthogonality and normalization of the free solutions.
X Z Z Z
+ 00 0
hr |T T | r i = dr dp dp0 ψi (x, p)vi+ (x0 , p)vj (x0 , p0 )ψj+ (x00 , p0 )
i,j=1,4 3
R R3 R3
X Z
= dp ψi (x, p)ψi+ (x00 , p) = δ(r − r00 )I4 . (4.22)
i=1,4
R3

The above unitarity relations can be explicitly written in terms of t± and f± : Eq.(4.21)
is equivalent to
Z Z Z
1 0
h
0
i+
dr dp 1 dp 2 t± (φ , p ) t± (φ0 , p2 ) ×
(2π~)6 1
h    i 1±β
× exp f± r0 , r00 , t; p2 − f± r0 , r, t; p1 = δ(r − r00 ) , (4.23)
2

Z Z Z
1 0
h i+
dr dp1 dp2 t± (φ0 , p1 ) t∓ (φ0 , p2 ) ×
(2π~)6
h    i
0 00 0
× exp f∓ r , r , t; p2 − f± r , r, t; p1 = 0 (4.24)

and (4.22) writes


Z Z Z
1 0
h
00
i+
dr dp1 dp2 t± (φ, p1 ) t± (φ , p2 ) ×
(2π~)6
h    i 1±β
× exp f± r, r , t; p1 ) − f± r , r , t; p2 = δ(r − r00 )
0 00 0
, (4.25)
2

Z Z Z
1 0
i+ h
00
dr dp1 dp2 t± (φ, p1 ) t∓ (φ , p2 ) ×
(2π~)6
h    i
× exp f∓ r, r0 , t; p1 − f± r00 , r0 , t; p2 = 0. (4.26)

Note that in the previous equations the arguments φ, φ0 and φ00 which appear in t± and
f± must be all taken at the same time t.

For further applications is also useful to explicitly write the action of T on an arbitrary
bispinor Ψ(r, t). Using Eqs.(4.14) - (4.16) we have
Z Z
0 0 0
hr|T |Ψi = dr hr|T+ |r iΨ(r , t) + dr0 hr|T− |r0 iΨ(r0 , t) (4.27)
R3 R3
48 Relativistic generalization of the space translation method

with
Z Z Z
0 0 1 0 0
Ψ(r0 , t) ×

dr hr|T± |r iΨ(r , t) = 3
dp dr t ± φ, p
(2π~)
R3 R3 R3
  
i Zφ
0 1  2 2
 
× exp ±p · (r − r ) + dχ 2eA(χ) · p ∓ e A (χ)  =
~ 2(n · p) 
−∞
Z
1 
= dp t± φ, p ×
(2π~)3/2
R3
  
i Zφ
1  
dχ 2eA(χ) · p ∓ e2 A2 (χ) 

× exp ±p · r +
~ 2(n · p) 
−∞
Z  
1 i
× 0
dr exp ∓ p · r Ψ(r0 , t).
0
(4.28)
(2π~)3/2 ~
R3

In the last integral in the previous equation appears the Fourier transform hp|Ψ(t)i of the
wavespinor Ψ(r, t); with this the action of the operator T becomes
Z
1 
hr|T± |Ψi = 3/2
dp t± φ, p h±p|Ψ(t)i ×
(2π~)
R3
  
i Zφ
1  2 2
 
× exp  ±p · r + dχ 2eA(χ) · p ∓ e A (χ)  . (4.29)
~ 2(n · p) 
−∞

One notices that to the integral over p contributes only the domain in which the Fourier
transform of Ψ(r, t) is nonvanishing.
We shall explicitly write also the action of T + on a wavespinor Φ. From Eqs.(4.19) -
(4.20) we have
Z Z
0
+
hr|T |Φi = dr hr|T++ |r0 iΦ(r0 , t) + dr0 hr|T−+ |r0 iΦ(r0 , t), (4.30)
R3 R3

with
Z Z Z
0 1 +
dr hr|T±+ |r0 iΦ(r0 , t) = dp dr0 t± φ0 , p Φ(r0 , t) ×
(2π~)3
R3 R3 R3
Zφ0
  
 i 1  
− ∓p · (r − r0 ) + dχ 2eA(χ) · p ∓ e2 A2 (χ) 

× exp (4.31)
 ~ 2(n · p) 
−∞
Relativistic generalization of the space translation method 49

Unlike in the previous case, we can not get a simple expression of the result in terms of
the Fourier transform of the function Φ(r). However it is useful to calculate the Fourier
transform hq|T±+ |Φi of the whole result T±+ Φ. We get
Z Z Z  
1 0 i 0
+
+
hq|T± |Φi = dr dp dr exp − q · r t± φ , p Φ(r0 , t) ×
(2π~)9/2 ~
R3 R3 R3
Zφ0
  
 i 1 
0 2 2
 
× exp −  ∓p · (r − r ) + dχ 2eA(χ) · p ∓ e A (χ)  . (4.32)
 ~ 2(n · p) 
−∞

In the previous equation the integral over r can be directly calculated, with the result
(2π~)3 δ(p ∓ q)
pand then the δfunction is used to calculate the integral over p. Using the
notation q ≡ q2 + m2 c2 , q , we obtain
Z
1 +
hq|T±+ |Φi
= 3/2
dr0 t± φ0 , q Φ(r0 , t) ×
(2π~)
R3
Zφ0
  
 i 1  
× exp − ±q · r0 + dχ 2eA(χ) · q ∓ e2 A2 (χ)  .

(4.33)
 ~ 2(n · q) 
−∞

The explicit expressions (4.29) and (4.33) shall be used in the following Section in order
to obtain approximate expressions of operators T and T + .

4.3 The atomic low momentum regime approxima-


tion of T
The potential V 0 = T + V T in Eq.(4.10) is a very complicate non-local matrix operator,
which makes this form of Dirac equation even more difficult to handle numerically than
the initial equation. These difficulties can be avoided in the atomic low momentum regime
approximation discussed in this Section.
If a light atom interacting with a laser pulse is described in the generalized translated
Dirac picture by a wavespinor Ψ(r) with the property that its Fourier transform hq|Ψi
contains only low momenta

hq|Ψi = 0, if |q| > λ(mc), λ  1 (4.34)

then we shall denote this situation as being in atomic low momentum regime. For example,
an isolated light atom in a bound state or a superposition of low energy continuum states is
in low momentum regime; note that for this example not only condition (4.34) is verified,
but also negative energy states are not populated. However, our definition allows the
presence of negative energy components in a low momentum regime wavespinor.
50 Relativistic generalization of the space translation method

The ensemble of all low momentum wavespinors forms a subspace D in the quantum
mechanical Hilbert space H. In this Chapter we shall prove that if a light atom is described
by a low momentum wavespinor at the initial moment, chosen before the laser pulse has
reached the atom, then this regime will be maintained during the interaction with the
laser

Ψ(r, −∞) ∈ D −→ Ψ(r, t) ∈ D, (∀)t. (4.35)

In the following we shall derive the low momentum regime approximation of T and T + .
In order to simplify the calculations we define the dimensionless quantity

eA(φ)
a(φ) = (4.36)
mc
and we introduce the notations

Zφ Zφ
1
A(φ) = dχ a(χ), B(φ) = dχ a2 (χ), (4.37)
2
−∞ −∞

R± (r) = r ± A(φ) − n B(φ), φ = ct − z. (4.38)

We shall denote the solutions r0 of the equations r = R± (r0 ) by R−1


± (r); also we shall use
1
the identity

1
δ(R± (r) − r0 ) = a2 (φ)
δ(r − R−1 0
± (r )). (4.39)
1+ 2

The approximation of T on D
If Ψ is an low momentum wavespinor Ψ ∈ D then its Fourier transform appearing in
Eq.(4.29) limits the domain of the integral over p to a sphere of radius |p| = λ(mc),
λ  1. It is then justified to use an approximation of T obtained by replacing the
p
integrand in the definition (4.13) by its series expansion up to the first order in mc . The
approximate expression T̃ of the operator T in the low momentum regime approximation
is obtained by replacing in Eq.(4.14) the matrices t± (φ, p) and the exponent f± (r, r0 , t; p)
p
by their expansions up the first order in mc . Using Eqs.(4.15), (4.17), (4.18) we obtain

(0) p (1)
t± (φ, p) ≈ t± (φ) + · t (φ) = t̃± (φ, p) (4.40)
mc ±
1
P δ(τ −τn )
This identity is based on the well-known property of the δ function δ(f (τ )) = |f 0 (τn )| , where τn
n
are simple zeros of f .
Relativistic generalization of the space translation method 51

with

1 + 2i σ · (n × a(φ)) 0
 
(0)
t+ (φ) =  , (4.41)
− 21 σ · a(φ) 0
1
 
0 2
σ · a(φ)
(0)
t− (φ) =  , (4.42)
i
0 1− 2
σ · (n × a(φ))
1
 
4
[−a(φ) + i(n × a(φ))(σ · n) + in(σ · (n × a(φ)))] 0
(1)
t+ (φ) =  , (4.43)
σ 1
2
+ [i(n × a(φ)) − a(φ)(σ · n) − n(σ · a(φ))]
4
0
σ
+ 14 [−i(n × a(φ)) + a(φ)(σ · n) + n(σ · a(φ))]
 
0 2
(1)
t− (φ) = , (4.44)
1
0 4
[a(φ) − i(n × a(φ))(σ · n) − in(σ · (n × a(φ)))]

and
i i
f± (r, r0 , t; p) ≈ ∓ (mc)B(φ) ± p · (R± (r) − r0 ) . (4.45)
~ ~
Using the above formulas in the definition (4.14) of T one can easily write the approxi-
mation T̃ = T̃+ + T̃− of T

i~∇0
  
0 0 i (0) (1)
hr|T± |r i ≈ hr|T̃± |r i = exp ∓ (mc)B(φ) t± (φ) ± t± (φ) · δ(R± (r) − r0 )
~ mc
(4.46)

where ∇0 denotes the gradient with respect to r0 .


The explicit action of T̃± on a bispinor Ψ can be easily obtained from the previous
relation
 "   #
i (0) (1) i~
hr|T̃± |Ψi = exp ∓ (mc)B(φ) t± (φ)Ψ(R± (r)) ± t± (φ) · ∇Ψ (4.47)
~ mc R± (r)

and it is a good approximation of the exact T as long as the spinor Ψ is in the subspace
D; as a consequence it makes sense to define T̃ only on D. However we must note that the
subspace D is not invariant under the action of T . We shall denote its (time-dependent)
range by ∆(t) = T (D) ≈ T̃ (D).

The approximation of T + on ∆(t)


The Fourier components of low momentum of T±+ Φ in Eq.(4.33) can be approximated by
q
expanding the integral up to the first order in mc . However, if the wavespinor Φ obeys
52 Relativistic generalization of the space translation method

the condition that T + Φ contains only low momenta then the approximate expression is
valid for all q in the Fourier spectrum of T±+ Φ. For such wavespinors we obtain
Z
1 0

(0)+ 0 q (1)+ 0

+
hq|T± |Φi ≈ dr t± (φ ) + · t (φ ) Φ(r0 , t) ×
(2π~)3/2 mc ±
3
 R   
i 0 i 0
× exp ± (mc)B(φ ) exp ∓ q · R± (r ) , ∀ q ∈ R3 (4.48)
~ ~
which can be transformed back to the position representation with the result
( )
0
 i 
±

exp (mc)B(φ ) (0)+ (1)+ i~
hr|T±+ |Φi ≈ ~
a2 (φ0 )
t± (φ0 ) Φ(r0 , t) ∓ t± (φ0 ) · ∇Φ (4.49)
1+ 2 mc
r0 =R−1 (r)
±

In the previous equations we have used the notations defined in Eqs.(4.36) - (4.38). Also
(0)+ (1)+ (0) (1)
t± and t± are the transpose conjugate of the matrices t± and t± defined in (4.41)
- (4.44). We notice that in fact in this way we have obtained an approximation of T + .
This approximation, which will denoted by T̃ + = T̃++ + T̃−+ , is valid on those functions Φ
having the property that T + Φ ∈ D. Since we previously saw that T (D) = ∆(t) and the
operator T is unitary, i.e. T + = T −1 it follows that T + (∆(t)) = D so the approximation
T̃ + is valid only on ∆(t). As a consequence T̃ + will be defined only on ∆(t) and its
range is D. From Eq.(4.49) we can easily derive the expression of T̃ + in the position
representation:
i~∇0
  
+ 0 + 0 i 0 (0)+ 0 (1)+
hr|T± |r i ≈ hr|T̃± |r i = exp ± (mc)B(φ ) t± (φ ) ∓ t± (φ) · δ(R± (r0 ) − r).
~ mc
(4.50)

Properties of the operators T̃ and T̃ +


Up to now we have defined a subspace D of the quantum mechanical Hilbert space H,
as the subspace of low-momentum wavespinors (wavespinors whose Fourier transform
contains only momenta which are low in comparison to mc (|p| < λ(mc), λ  1)). Then
we have found an approximation T̃ = T̃+ + T̃− of T , defined on D and with the range
∆(t) 6= D, and an approximation T̃ + = T̃++ + T̃−+ of T + , defined on ∆(t) and with the
range D. However we must note that, in spite of the notations used for the two, T̃ + is
not the adjoint of T̃ ; in fact none of them has an adjoint, as their range does not coincide
with the domain. However, in the following we shall prove that T̃ + is the inverse of T̃ up
to the lowest order in λ.
For the beginning we shall prove that the operator T̃ is isometric on its domain D
to the lowest order in λ. Let us take two low momentum wavespinors f1 (r), f2 (r) ∈ D
and calculate the inner-product hT̃ f1 |T̃ f2 i; as T̃ is a sum of two term, the inner-product
writes a a sum of four terms
X X Z  ∗
hT̃ f1 |T̃ f2 i = dr hr|T̃σ1 |f1 i hr|T̃σ2 |f1 i. (4.51)
σ1 ={−,+} σ2 ={−,+} R3
Relativistic generalization of the space translation method 53

Note that in the previous equation the star attached to the first term indicates the complex
conjugation and transposition of the bispinor. For each of hr|T̃σ1,2 |f1,2 i we shall use the
explicit expression (4.46); however, we must keep in mind that if f is a low momentum
wavespinor then its gradient i~∇mc
f is of the order O(λ) so it can be neglected in comparison
to the first term in (4.46) which is O(λ0 ). With this approximation, after a straight
calculation one obtains
Z  
(0)+ (0)
hT̃ f1 |T̃ f2 i = drf1∗ (R+ (r)) t+ (φ) t+ (φ) f2 (R+ (r)) + (4.52)
R3
Z  
(0)+ (0)
+ drf1∗ (R− (r)) t− (φ)t− (φ) f2 (R− (r)) +
R3
Z  

(0)+ (0)
 i
+ drf1∗ (R+ (r)) t+ (φ)t− (φ)
f2 (R− (r)) exp 2 (mc)B(φ) +
~
R3
Z  


(0)+ (0)
 i
+ drf1 (R− (r)) t− (φ)t+ (φ) f2 (R+ (r)) exp −2 (mc)B(φ) .
~
R3

By direct calculation one obtains


a2 (φ) 1 ± β a2 (φ)
 
(0)+ (0) (0)± (0) 1±β
t± (φ)t± (φ) = 1 + , t± (φ)t∓ (φ) = α·n . (4.53)
2 2 2 2
From the explicit expressions before one can see that the products of matrices in the
previous integrand are smooth functions of z. On the other hand, using the expression
(4.37) of B(φ) one can easily write its derivative with respect to z
∂ [(mc)B(φ)] mc
= − a2 (φ). (4.54)
∂z 2
As a = eA(φ)
mc
is of the order of unity it follows that B is a very fast oscillating function of
z, so the last two integrals in (4.52) vanish. In the remaining first two integrals we shall
perform the change of variable r → ρ = R± (r). As the corresponding Jacobian in both
cases is directly formed under the integral, one can easily obtain the result
 
1+β 1−β
Z

hT̃ f1 |T̃ f2 i = dρf1 (ρ) + f2 (ρ) = hf1 |f2 i + O(λ) (4.55)
2 2
R3

which end the proof of the isometry of T̃ to the desired order in λ.


Since T̃ is isometric on its domain D it has an inverse [7]. In the following we shall
prove by direct calculation that its inverse is T̃ + . In order to do so, we shall calculate the
action of the operator Ĩ ≡ T̃ + T̃ on a low momentum bispinor Ψ ∈ D, keeping only the
terms of order O(λ0 )
As both operators T̃ and T̃ + write as a sum of two terms, Ĩ is the sum of four terms
Ĩ = T̃++ T̃+ + T̃++ T̃− + T̃−+ T̃+ + T̃−+ T̃− . (4.56)
54 Relativistic generalization of the space translation method

Due to the presence of the factors 1±β in the expression of the operators T̃± the structure
of the four terms in the previous equation is very simple: for each of them only a 2 × 2
corner is non-zero. In order to simplify the notations we will redefine each of the four
terms in (4.56) as a 2 × 2 matrix containing the only non-vanishing corner of the original
4 × 4 matrix. With this redefinition one can write
hr|Ĩ++ |r0 i hr|Ĩ+− |r0 i
 

hr|Ĩ|r0 i =  , (4.57)
0 0
hr|Ĩ−+ |r i hr|Ĩ−− |r i
with
Z
0
hr|Ĩ±± |r i = dr00 hr|T̃±+ |r00 ihr00 |T̃± |r0 i,
R3
Z
0
hr|Ĩ±∓ |r i = dr00 hr|T̃±+ |r00 ihr00 |T̃∓ |r0 i. (4.58)
R3

The four terms in the previous equation are calculated using (4.46), (4.50) and also the

identity (4.39). The result, up to the first order in (mc) is
∇0
 
0 00 ∇ 10 00 10 00
hr|Ĩ±± |r i ≈ I±± ∓ i~ · I ±± (r ) ± i~ · I ±± (r ) ×
mc mc r00 =R−1 (r)
±

× δ(r − r0 ), (4.59)
∇0
 
0 00 00 ∇ 10 00 10 00
hr|Ĩ±∓ |r i ≈ I±∓ (r ) ∓ i~ · I ±∓ (r ) ∓ i~ · I ±∓ (r ) ×
mc mc r00 =R−1 (r)
±

× δ(R∓ (R−1
± (r))
0
− r ). (4.60)
In the above equations the following notations have been used:
00
I±± = I2 , I2 = the 2 × 2 identity matrix (4.61)
a2 (φ)  
00 2 i
I±∓ (r) = 2 (σ · n) exp ±2 B(φ) , (4.62)
1 + a 2(φ) ~
a2 (φ)
 
01 1 a(φ) (n × a(φ))
I ±± (r) = 2 ∓ ∓i (σ · n) + (n − i(σ × n)) , (4.63)
1 + a 2(φ) 2 2(mc) 4
a2 (φ)
  
01 1 σ (n × a(φ)) a(φ)
I ±∓ (r) = 2 1 − ± i ± (σ · n)+
1 + a 2(φ) 2 2 2 2
a2 (φ)
  
i
+ n(σ · n) exp ±2 B(φ) , (4.64)
2 ~
+
I 10 01
±± (r) = I ±± (r) , (4.65)

+
I 10
±∓ (r) = I 01
∓± (r) . (4.66)
Relativistic generalization of the space translation method 55

The symbols ∇ and ∇0 in Eqs.(4.59), (4.60) are the gradients with respect to r or r0 .
The second step in our approach is to explicitly calculate the action of Ĩ (4.56) on

an arbitrary spinor Ψ ∈ D, keeping only the lowest order terms in mc Ψ. After a straight
calculation one obtains
 
00 P 10 −1 10 −1 P
hr|I±± |Ψi = I±± ± · I ±± (R± (r)) ± I ±± (R± (r)) · Ψ(r). (4.67)
mc mc

(Note that from now on we must go back to the 4 × 4 form of the matrices I in order to

apply them on a bispinor.) In order to correctly evaluate the orders in mc in the previous
expression one must explicitly write the action of the momentum operator in its second
term. We have
 
P 10 −1 ∇ −1 −1 P
· I ±± (R± (r))Ψ(r) = −i~ · I ±± (R± (r)) Ψ(r) + I 10
10
±± (R± (r)) · Ψ(r). (4.68)
mc mc mc

If the wavespinor Ψ is in the subspace D, i.e. its Fourier transform does not contain
components of momenta larger that λ(mc) with λ  1, then the third term in the
previous equation and the third term in Eq.(4.67) are of order λ; since also the second
term in Eq.(4.68) is neglijible2 the final expression is
00
hr|I±± |Ψi = I±± Ψ(r) + O(λ) ≡ I2 Ψ(r) + O(λ) (4.69)

A similar analysis has to be done for the off-diagonal components I±∓ but, as we shall
see, in this case the calculation is more difficult. We have
 
P P
hr|I±∓ |Ψi = 00
T±∓ (R−1
±
± (r))· I 10 −1 10 −1
±∓ (R± (r)) ∓ I ±∓ (R± (r)) · ×
mc mc
−1
× Ψ(R∓ (R± (r))). (4.70)

As before, the term in discussion is the second one in the previous equation
 
P −1 −1 ∇
10
· I ±± (R± (r))Ψ(R∓ (R± (r))) = −i~ · I ±∓ (R± (r)) Ψ(R∓ (R−1
10 −1
± (r))) +
mc mc
−1 P
+ I 10
±∓ (R± (r)) · Ψ(R∓ (R−1
± (r))) (4.71)
mc

The complication arises from the exponential present in I 10


±∓ , whose derivative is (4.54)

  " a2 (φ)  #
∇ i 2 i
−i~ exp ±2 B(φ) = ∓n 2 exp 2 B(φ) (4.72)
mc ~ r=R−1 (r) ±
1 + a (φ)2
~
r=R−1
± (r)

2
See the discussion at the end of Section 4.4
56 Relativistic generalization of the space translation method

One can prove 3 that the other terms in Eq.(4.71) are neglijible and also the last term in
Eq.(4.70) such that
"
00
hr|I±∓ |Ψi = I±∓ (R−1
± (r))− (4.73)

a2 (φ)  !
2 i  Ψ(R∓ (R−1
− 2 (σ · n) exp ±2 B(φ) ± (r))) +
1 + a 2(φ) ~
r=R−1
± (r)

+O(λ)
00
Taking into account the expression (4.62) of I±∓ one notices that the two terms in the
parenthesis cancel, such that the final result is

hr|I±∓ |Ψi = O(λ) (4.74)

From the results (4.69) and (4.74) one can see that when acting on a spinor from the
subspace D the operator I can be approximated as
δ(r − r0 )I2
 
0
hr|Ĩ|r0 i ≈   + O(λ) = δ(r − r0 )I4 + O(λ). (4.75)
0 δ(r − r0 )I2

In conclusion, we have proven that T̃ + T̃ = I + O(λ) ; since, on the other hand we already
know that T̃ has an inverse, it follows that to the zeroth order of λ we have T̃ + = T̃ −1 .

4.4 The low momentum regime approximation of the


transformed potential V 0
We are looking for an approximate expression of the potential V 0 = T + V T which must
correctly describe the action of V 0 on a low momentum wavespinor Ψ ∈ D. Then in the
expression of V 0 we can replace the rightmost operator T by its approximation T̃ , defined
on D. On the result which is in ∆(t) one applies the potential V ; if V describes a light
atom then its Fourier transfom does not contain high momenta and ∆(t) is invariant to its
action at any time (see the proof an the end of this Section). Then the leftmost operator
T + can be replaced by its approximation T̃ + , defined on ∆(t) and with the range D. This
way we have found the approximation Ṽ of V 0

Ṽ = T̃ + V T̃ (4.76)

and we have also proven that D in invariant under the action of Ṽ .


In the following we shall explicitly calculate the transformed potential Ṽ . The cal-
culation is very similar to that presented in the previous section for Ĩ = T̃ + T̃ ; we only
3
See the discussion at the end of Section 4.4
Relativistic generalization of the space translation method 57

briefly present the main steps. As a first step we obtain an expression of Ṽ by using in its
definition (4.76) the formulas (4.46) and (4.50). In the second step we explicitly calculate
the action of Ṽ on an arbitrary low momentum wavespinor and keep only the lowest term
p
in mc .
As in the case of Ĩ, Ṽ is the sum of four terms

Ṽ = T̃++ V T̃+ + T̃++ V T̃− + T̃−+ V T̃+ + T̃−+ V T̃− . (4.77)

which shall be redefined as 2 × 2 matrices. We obtain


hr|Ṽ++ |r0 i hr|Ṽ+− |r0 i
 

hr|Ṽ |r0 i =  , (4.78)


hr|Ṽ−+ |r0 i hr|Ṽ−− |r0 i

with
Z
0
hr|Ṽ±± |r i = dr00 hr|T̃±+ |r00 iV (r00 )hr00 |T̃± |r0 i,
R3
Z
0
hr|Ṽ±∓ |r i = dr00 hr|T̃±+ |r00 iV (r00 )hr00 |T̃∓ |r0 i. (4.79)
R3


The four terms in the previous equation up to the first order in (mc)
are

∇0
 
0 00 00 ∇ 10 00 10 00
hr|Ṽ±± |r i ≈ V±± (r ) ∓ i~ · V±± (r ) ± i~ · V±± (r ) ×
mc mc r00 =R−1
± (r)

× δ(r − r0 ), (4.80)
∇0
 
0 00 00 ∇ 10 00 10 00
hr|Ṽ±∓ |r i ≈ V±∓ (r ) ∓ i~ · V±∓ (r ) ∓ i~ · V±∓ (r ) ×
mc mc r00 =R−1
± (r)

× δ(R∓ (R−1 0
± (r)) − r ). (4.81)

In the above equations we have used the notations:


ij ij
Vσ1σ2 (r) = V (r)Iσ1σ2 , σ1 , σ2 ∈ {+, −}, i, j ∈ {0, 1} (4.82)

The second step in our approach is to explicitly calculate the action of the approxi-
mated Ṽ (4.80), (4.81) on an arbitrary low momentum spinor Ψ, keeping only the lowest

order terms in mc Ψ. Going back again to the 4 × 4 representation we get
 
00 −1 P 10 −1 10 −1 P
hr|V±± |Ψi = V±± (R± (r)) ± · V±± (R± (r)) ± V±± (R± (r)) · Ψ(r).(4.83)
mc mc
The second term in the above expression can be written as
 
P −1 ∇ −1 P
10
· V±± (R± (r))Ψ(r) = −i~ 10
· V±± (R± (r)) Ψ(r) + V±±10
(R−1
± (r)) · Ψ(r); (4.84)
mc mc mc
58 Relativistic generalization of the space translation method

the previous equation is very similar to Eq.(4.68) and using a similar argument one can
prove that both terms in the previous equation and also the third term in Eq.(4.83) are
negligible such that the final expression is
00
hr|V±± |Ψi = V±± (R−1 −1
± (r))Ψ(r) + O(λ) = V (R± (r))Ψ(r) + O(λ) (4.85)

For the off-diagonal components V±∓ we have


 
00 −1 P 10 −1 10 −1 P
hr|V±∓ |Ψi = V±± (R± (r)) ± · V±∓ (R± (r)) ∓ V±∓ (R± (r)) · ×
mc mc
× Ψ(R∓ (R−1
± (r))). (4.86)

With the identity


 
P −1 −1 ∇
10
· V±± (R± (r))Ψ(R∓ (R± (r))) = −i~ · V±∓ (R± (r)) Ψ(R∓ (R−1
10 −1
± (r))) +
mc mc
P
10
+ V±∓ (R−1
± (r)) · Ψ(R∓ (R−1
± (r))) (4.87)
mc
and, using again (4.72) and the same type of argument as before we have
"
00
hr|V±∓ |Ψi = V±∓ (R−1
± (r))− (4.88)

a2 (φ)  !
2 i  Ψ(R∓ (R−1
− V (r) 2 (σ · n) exp ±2 B(φ) ± (r))) +
1 + a 2(φ) ~
r=R−1
± (r)

+O(λ)
00
Taking into account the expression (4.62), (4.82) of V±∓ one notices that the two terms
in the parenthesis cancel, such that the final result is

hr|V±∓ |Ψi = O(λ) (4.89)

Form the results (4.85) and (4.89) and with the definition (4.78) one can see that when
acting on a low momentum wavespinor the potential V 0 can be approximated as

V (R−1 0
 
+ (r))δ(r − r )I2 0
hr|Ṽ |r0 i ≈  =
−1 0
0 V (R− (r))δ(r − r )I2
 
0 −1 1+β −1 1−β
= δ(r − r ) V (R+ (r)) + V (R− (r)) (4.90)
2 2

where I2 is the 2 × 2 unity matrix.


Relativistic generalization of the space translation method 59

The form (4.10) of the Dirac equation is unitarily equivalent to the standard form of
the Dirac equation (4.1) as long as for V 0 is used the exact expression (4.5). However we
saw that in low momentum regime the use of the approximate form (4.90) of the potential
is justified; in this case also the free Hamiltonian H0 can be approximated as
√ P2
 
2 2 2 4 2
c P + m c β ≈ mc + β. (4.91)
2m

In this approximation, since both the free Hamiltonian and the potential are diagonal

P2
   
2 −1 1+β −1 1−β
H ≈ H̃ = mc + β + V (R+ (r)) + V (R− (r)) (4.92)
2m 2 2

the upper and lower components of the bispinor Ψ


 
Ψe
Ψ= (4.93)
Ψp

are uncoupled, the equations obeyed by them being

P2
 
∂Ψe 2 −1
i~ = mc + + V (R+ (r)) Ψe (4.94)
∂t 2m
P2
 
∂Ψp 2 −1
i~ = −mc − + V (R− (r)) Ψp . (4.95)
∂t 2m

We note that the subspace D of low momentum spinors is invariant under the action of
the free Hamiltonian H0 and, as we saw earlier, it is also invariant under Ṽ . It follows that
if the solution Ψ is in D at the initial moment, it will remain in that subspace during the
interaction with the pulse, which confirms that the atomic low momentum approximation
is valid at any time. Since the positive, respectively negative, energy solution are described
by spinors of the form
 
Ψe
Ψelectronic = , (4.96)
0
 
0
Ψpositronic = (4.97)
Ψp

the absence of coupling between Eqs(4.94) and (4.95) indicates the fact that pair creation
does not take place. If the system wavespinor in the generalized translated Dirac picture
has the form (4.96) at the initial moment - as is the case for a bound state of a light atom
-, then it have the same form at any time, and in fact only the Schrödinger-like equation
(4.94) has to be solved. We also note that spin effects are not present in Eq.(4.94); they
could be obtained by taking higher order in λ in the low momentum regime approximation.
60 Relativistic generalization of the space translation method

Proof of the invariance of ∆(t) to the action of V


Let us take a spinor Φ(r) in the subspace ∆(t); then it can be written as
Φ(r, t) = T̃ (t)Ψ(r) (4.98)
where Ψ(r) is in D. In order to prove that
V Φ(r) ≡ V T̃ (t)Ψ(r) ∈ ∆(t), (4.99)
according to the definition of ∆(t), we have to show that
T̃ + (t)V φ(r) ≡ T̃ + (t)V T̃ (t)Ψ(r) ∈ D (4.100)
i.e. that the Fourier transform of T̃ + (t)V T̃ (t)Ψ(r) contains only low momenta. From the
result (4.90) it follows that in fact we have to investigate the Fourier transform of the
spinor
 
−1 1+β −1 1−β
Ξ(r, t) ≡ V (R+ (r)) + V (R− (r)) Ψ(r). (4.101)
2 2
Next, using the property that the Fourier transform of a product is the convolution of
Fourier transforms of the two factors, and the fact that Ψ(r) is in D, one finds that we
have to prove that the Fourier transform of V (R−1
± (r)) does not contain high momentum
components. We shall make the assumption that, since V describes an light atom, the
Fourier transform of V (r) does not have high momenta components. Also we shall use the
observation that that the property of a function f of having only low momenta Fourier
components
hq|f i ≈ 0, if |q| ≥ λ(mc), λ  1 (4.102)
if equivalent to
P
f (r) = O(λ). (4.103)
mc
The consequence of the above remarks is that in our case it is enough to investigate the
gradient of V (R−1
+ (r)).
Let us take an arbitrary function f (r). Using the definition (4.38) one can easily prove
that

−1 ∂f ∂f
∇f (R± (r)) = ex + ey (4.104)
∂x ∂y
!#
1 ∂f ax (φ) ∂f ay (φ) ∂f
+ez 2 ± 2 ± 2
1 + a (φ) ∂z
2
1 + a (φ) ∂x 1 + a (φ) ∂y
2 2 −1
R± (r)

where R−1 ± (r) is that defined in Section 4.3. From the above identity it follows that if
P
mc

f (r) ≡ −i~ mc P
f (r) is negligible, then mc f (R−1
± (r)) can be also neglected, which when
applied for the case of the atomic potential V ends the proof of invariance of ∆(t) to the
action of V . (Note that this observation applies also for the terms which were discarded
in Eqs.(4.67), (4.70), (4.83), (4.86).)
Relativistic generalization of the space translation method 61

4.5 Position and velocity operators in the generalized


translated Dirac picture
As we shall see in the next Chapter, the position and velocity operators are particularly
important for the physical interpretation of the generalized translated Dirac picture and
also their knowledge allows us to make a series of qualitative predictions concerning the
behavior of the atomic systems in very intense plane wave laser fields. Their expressions
are obtained in the following.

The position operator


The position operator in the generalized translated Dirac picture is calculated by applying
the unitary transformation T to the “mean position operator” defined by Foldy and
Wouthuysen [6].
 
0 + βα 2 β(α · P)P + (σ × P)p
R =T R + i~c − i~c T,
2Ep 2Ep (Ep + mc2 )p

Ep = c P2 + m2 c2 , p = |P|. (4.105)

In the atomic low momentum regime approximation the operator T must be replaced by
its approximation T̃ given by Eq.(4.46), T̃ + (4.50) must be used instead of T + and also
the Foldy’s mean position operator must be approximated up to the first order in 1/c.
With these approximations one obtains

R0 ≈ R̃ = T̃ + RT̃ . (4.106)

In order to evaluate the previous expression one must use exactly the same method as for
the atomic low momentum regime approximation of the potential Ṽ ; this can be easily
understood as the atomic potential V (r) is in fact a function of the position operator R.
Based on these considerations, from Eq.(4.90) one can directly write the result
 −1
R+ (r)δ(r − r0 )I2

0
hr|R̃|r0 i =  =
−1 0
0 R− (r)δ(r − r )I2
 
0 −1 1+β −1 1−β
= δ(r − r ) R+ (r) + R− (r) . (4.107)
2 2

The velocity operator


An approximation of the velocity operator in the generalized translated Dirac picture is
calculated as
 
dR̃ 1
Ṽ = + R̃, H̃ (4.108)
dt i~
62 Relativistic generalization of the space translation method

where H̃ and R̃ are given by the approximate expressions (4.92), (4.107). After a lengthy
but straight calculation the result is
 
Ṽ+ I2 0
Ṽ = (4.109)
0 Ṽ− I2
with
" #
Px eAx (Φ± ) 1 1 ax (Φ± ) ax (Φ± )
Ṽ±,x = − 2 + Pz a2 (Φ± )
+ P
a2 (Φ± ) z
(4.110)
m m 1 + a (Φ 2
±) 2m 1 + 2
1 + 2
" #
Py eAy (Φ± ) 1 1 ay (Φ± ) ay (Φ± )
Ṽ±,y = − 2 + Pz a2 (Φ± )
+ P
a2 (Φ± ) z
(4.111)
m m 1 + a (Φ 2
±) 2m 1 + 2
1 + 2
" #
2 2
e A (Φ± ) 1 1 1 1
Ṽ±,z = + Pz + Pz , (4.112)
2m2 c 1 + a2 (Φ± ) 2m 2
1 + a (Φ± )
2
1 + a (Φ± )
2 2 2

In the previous equation Px,y,z are the Cartesian components of the derivation operator
P = −i~∇ and Φ± must be calculated as

Φ± = ct − n · R−1
± (r). (4.113)

4.6 References
[1] J. Z. Kamiński, J. Phys. A 18, 3365 (1985).

[2] P. S. Krstic and M. H. Mittleman, Phys. Rev. A 42, 4037 (1990)

[3] P. S. Krstic and M. H. Mittleman, Phys. Rev. A 45, 6514 (1992)

[4] M. Protopapas, C. H. Keitel and P. L. Knight, J. Phys. B 29, L591 (1996).

[5] M. Førre, S. Selstø, J. P. Hansen, and L. B. Madsen, Phys. Rev. Lett. 95, 043601 (2005).

[6] L.L. Foldy and S.S. Wouthuysen, Phys. Rev. 78, 26 (1950).

[7] N. I. Akhiezer and I. M. Glazman, Theory of linear operators in Hilbert space, Dover,
1994.
Chapter 5

The generalized translated


Schrödinger equation

In this Chapter we study the lowest order approximation of the generalized translated
Dirac equation In this approximation the evolution equation for an electronic wavepacket
reduces to a Schrödinger-like equation, (4.94) which shall be called the generalized trans-
lated Schrödinger equation. Several equivalent forms of it will be derived, and also a
comparison with the non-relativistic results will be presented. The general predictions
obtained on a intuitive basis here are confirmed by the numerical calculations presented
in the next Chapter.

5.1 Derivation and properties of the generalized


translated Schrödinger equation
In the low momentum regime approximation approximation described in the previous
Chapter the Dirac equation in the generalized translated picture picture (4.10) reduces
to a system of uncoupled equations (4.94), (4.95) for the electronic and positronic parts
of the wavespinor. In these conditions the time evolution of an electron is described by
the Schrödinger-like equation

P2
 
∂ψ(r, t) 2 −1
i~ = mc + + V (R+ (r)) ψ(r, t) (5.1)
∂t 2m

form which the term mc2 can be eliminated by the usual unitary transformation
 
i 2
ψ(r, t) −→ exp − mc t ψT (r, t) (5.2)
~
the final form being
∂ψT (r, t)
i~ = HT ψT (r, t). (5.3)
∂t
63
64 The generalized translated Schrödinger equation

with
P2
HT = + V (R−1
+ (r)). (5.4)
2m
The previous equation is the lowest order approximation of the generalized space
translated Dirac equation. As in the non-relativistic dipole case, the entire field effect is
contained in the modified potential

Ṽ (r, t) = V (R−1
+ (r)). (5.5)

We must note that in this case, unlike in the non-relativistic case, due to the inclusion
of the relativistic effects, the modified potential is not simply translated, but is is also
distorted during the interaction with the pulse; however, for simplicity Ṽ (r) will be named
in the following the generalized translated potential and the corresponding Schrödinger-
like equation (5.3) - the generalized translated Schrödinger equation, or the Schrödinger
equation in the generalized translated picture. It is instructive here to discuss the relation
between the explicit expression of the generalized translated potential and the classical
relativistic trajectory γ(t) of an electron interacting with the laser field. By comparing
Eqs.(C.34,C.35) with the definition of R−1+ (r) (4.38) one can easily see that the generalized
translated potential can be written as
  
−1 1 −1
V (R+ (r)) = V r + γ t − n · R+ (r) . (5.6)
c

We remind here that in the non-relativistic case the translated potential is (2.5)

Vtransl (r) = V (r + α(t)) , (5.7)

where α(t) is the classical nonrelativistic trajectory, so using the term “generalization of
the space translation method” is justified in this sense.

The position and velocity operators in the generalized translated picture are the upper
corners of the expressions (4.107) and (4.108), respectively

RT = R−1
+ (r), (5.8)

0
VT = V+ (5.9)

It is also instructive to find the probability density in terms of the wavefunction ψT in


the generalized translated picture. If the system is described by an electronic wavespinor
Φ(r, t), solution of the Dirac equation (4.1) then the probability to find the particle at
the moment t in the infinitesimal volume dr at the point r in the laboratory frame is

dp = Φ+ (r, t)Φ(r, t) dr. (5.10)


The generalized translated Schrödinger equation 65

The probability density with respect to the laboratory frame P(r, t) = Φ+ (r, t)Φ(r, t) can
be expressed in terms of the wavespinor Ψ in the generalized translated picture using the
relation between them Φ = T Ψ
P(r, t) = (T Ψ(r, t))+ (T Ψ(r, t)) . (5.11)
In the low momentum regime approximation Ψ reduces to the one-component solution
ψT of the generalized translated Schrödinger equation (5.3) and T must be replaced by
its lowest-order approximation
 
i (0)
hr|T |r i ≈ exp − (mc)B(φ) t+ (φ) δ(R± (r) − r0 )
0
(5.12)
~
such that the above expression becomes
a2 (φ)
 
2
PT (r, t) = |ψT (R+ (r), t)| 1 + . (5.13)
2
We emphasize here that the above expressions of the position and velocity operators,
and of the probability density represent the respective quantities in the laboratory frame;
unlike in the non-relativistic case, the unitary transformation used does not correspond
to a change of the reference frame, so one can not define a “particle position with respect
to a translated reference frame”.

5.2 The generalized translated potential


In this Section we study the generalized translated potential Ṽ (r) (5.5). Up to now the
theory was derived for the case of an arbitrary atomic potential. In the following we
consider the case of the Hydrogen atom
e2 e2
V (r) = − , −→ Ṽ (r) = − , (5.14)
|r| |R−1
+ (r)|

the Schrödinger equation in the generalized translated picture (5.3) being


 2
e2

∂ψT (r, t) P
i~ = HT ψT (r, t) ≡ − ψT (r, t). (5.15)
∂t 2m |R−1 + (r)|

For simplicity we denote R−1


+ (r) by R0 ; it is calculated as the solution of the equation

ZΦ0 ZΦ0
1 n
R0 = r − dχeA(χ) + dχe2 A2 (χ), Φ0 = ct − n · R0 (5.16)
mc 2(mc)2
−∞ −∞

From Eq.(5.14) one can see that in the Coulomb case the generalized translated potential
has a singularity for R0 = 0. The singularity trajectory, denoted by rs (t) is given by
Zct Zct
1 n
rs (t) = dχeA(χ) − dχe2 A2 (χ). (5.17)
mc 2(mc)2
−∞ −∞
66 The generalized translated Schrödinger equation

Although, as we previously saw, we use a generalization of the space translation transfor-


mation, there is a relation between the trajectory of the singularity of Ṽ (r, t) and that
of a classical electron in the electromagnetic field which is in some respect similar to the
relation existent in the non-relativistic dipole (NR-D) case. We remind here that in the
NR-D case the potential singularity in the Kramers-Hennberger frame moves along a tra-
jectory which is identical up to a sign to the non-relativistic trajectory of the electron in
the external field.
Zct
e
rs,N R−D (t) = −α(t) = dχA(χ) (5.18)
mc
−∞

In the relativistic (R) case we shall compare the trajectory (5.17) with the relativistic
trajectory γ(t) of an electron initially at rest in the origin of the reference frame interacting
with the laser pulse A(φ). Using the results (C.34), (C.35) we have
Zφγ Zφγ
1 n
γ(t) = − dχeA(χ) + dχe2 A2 (χ), φγ = ct − n · γ. (5.19)
mc 2(mc)2
−∞ −∞

From (5.17) and (5.19) one can see that the singularity rs follows exactly the trajectory
−γ(t) but at different speed
" #
d(−rs (t)) eA(ct) e2 A2 (ct) dγ(t) 1 d(−rs (φ))
=− +n , = 2 2 .
dt m 2m2 c dt 1 + e A (φ)
2

2(mc) φ=(ct−n·γ)
(5.20)
Another consequence of the previous equations is that, unlike the classical relativistic
electron, the singularity rs (t) can move at speed larger than c for large enough field
intensities. We emphasize here that this is not a contradiction or a limitation of our model,
since rs (t) does not denote the trajectory of a real particle, but a fictitious potential.
We present a numerical example for the case of a Gaussian pulse linearly polarized
along the Ox axis with FWHM τp = 1 cycle and frequency ω = 1 au; the corresponding
vector potential
"  2 #  
1.1774φ φ
A(φ) = A0 exp − sin ω (5.21)
cτp c

is represented in Fig.5.1 as a function of φ (note that the unit on the Oy axis is the peak
intensity A0 ). The common trajectory of γ(t) and −rs (t) for the pulse described above
is drawn in Fig.5.2 in black line. With red/blue dots are marked instantaneous positions
of γ(t)/−rs (t) at several moments of time. At the beginning of the pulse the intensity is
low, the retardation effects are negligible and the two moves together, the instantaneous
positions being identical. For t > −T /2 appears a difference of speed between γ(t) and
−rs (t), and the distance between them can increase up to 50 au at the moment t = T /4.
The generalized translated Schrödinger equation 67

0.75

Gaussian pulse
0.5
τp = 1

0.25
A(φ) / A0

-0.25

-0.5

-0.75

-1
-2 -1.5 -1 -0.5 0 0.5 1 1.5 2
φ / (cT)

Figure 5.1: The graph of A(φ) for a Gaussian pulse with FWHM τp = 1 cycle

150

Gaussian pulse

τp = 1, ω = 1 au, E0 = 100 au. t = 2T

t=T
100
t = T/2

t = - T/8
z (au)

t = T/4
γ (t)
- r s(t)

50 t=0

t = - T/8
t = - T/4

t = - T/2

t = -T
0
t = -2T
-100 -50 0 50 100
x (au)

Figure 5.2: The common trajectory of γ(t) and −rs (t). The positions at several moments
of time are marked in red for γ(t) and in blue for −rs (t). The pulse parameters are
written on the figure.
68 The generalized translated Schrödinger equation

However, at the end of the pulse the two are left at rest in the same point located along
the Oz axis. The total displacement can be analytically calculated (see (C.37)) the result
being
r
e2 E02 π πτp
∆z = . (5.22)
2m2 cω 3 2 1.1774
Note that is scales at E02 and can become extremely large for pulses of relativistic inten-
sities.

5.3 Equivalent forms of the generalized translated


Schrödinger equation
The form (5.15) of the generalized translated Schrödinger equation is very compact and
explicitly shows the similarity with the non-relativistic Schrödinger equation written in
the Kramers-Henneberger frame (2.5). However, this form is not always convenient: for
example, it is not possible to explicitly separate the pure relativistic effects from the
retardation ones and it is also not convenient for numerical integration because of the
moving singularity of the generalized potential. In order to solve these difficulties we have
derived two equivalent forms of the equation 5.15 which are presented in the following.

5.3.1 The generalized Schrödinger equation in the “R” picture


This form is obtained by applying a pure translation by −rs (t) to the generalized trans-
lated Schrödinger equation (5.15).
 
i
ψT (r, t) −→ ψR (r, t) = UT R ψT (r, t) ≡ exp P · rs (t) ψT (r, t). (5.23)
~

The equation obeyed by ψR (r, t) is

∂ψR (r, t)
i~ = HR ψR (r, t) (5.24)
∂t
with
P2 drs (t)
HR = −P· + VR (r, t), VR (r, t) = Ṽ (r + rs (t)), (5.25)
2m dt
and will be named in the following the generalized Schrödinger equation in the “R” picture.
Taking into account the definition (5.5) of the generalized translated potential one can
easily check that the potential VR (r, t) in the previous equation can be written as

e2
VR (r, t) = − −1 , (5.26)
R+ (r)
The generalized translated Schrödinger equation 69

where R+ (r) is given by

Zφ Zφ
1 n
R+ (r) = r + dχeA(χ) − dχe2 A2 (χ), φ = ct − n · r; (5.27)
mc 2(mc)2
ct ct

−1
and R+ (r) is the solution R0 of the equation R+ (R0 ) = r.
Although the potential VR is time dependent, one can check that its singularity is fixed
in the origin of the reference frame. We shall also introduce the notation

e drs (t) eA2 (ct)


Aeff (ct) ≡ , =⇒ Aeff (ct) = A(ct) − n . (5.28)
m dt 2mc
With this notation the our evolution equation in the “R” picture becomes
∂ψR (r, t)
i~ = HR ψT (r, t) (5.29)
∂t
with
P2 e e2
HR = − P · Aeff (ct) − −1 . (5.30)
2m m |R+ (r)|

Formally, it is identical to a non-relativistic Schrödinger equation written in the mo-


mentum gauge, for the “effective atomic potential” VR and the external field treated in
the dipole approximation Aeff (ct). However, we must note that our “effective atomic
potential” is time dependent, although its singularity is fixed. This form of the evolu-
tion equation is more convenient than (5.15) since it allows the explicit separation of the
first-order retardation effects from the relativistic ones.

The retardation corrections to the non-relativistic Schrödinger equation


We shall briefly present here the problem of including the first order retardation correc-
tions into the nonrelativistic Schrödinger equation. This problem was previously treated
in several papers (see, for example [1, 2, 3, 4]); here we shall only sketch the derivation.
One starts from the non-relativistic Schrödinger equation
" #
∂ψ(r, t) (P − eA (ct − n · r))2 e2
i~ = − ψ(r, t). (5.31)
∂t 2m r

and one expands the Hamiltonian as a power series in 1c keeping only the terms up to the
first order. The result is
 2
e2 A2 (ct) dA(ct) e2

∂ψ(r, t) P e 2n · r
i~ = + − A(ct) · P − e A(ct) · − ψ(r, t).
∂t 2m 2m m mc dt r
(5.32)
70 The generalized translated Schrödinger equation

The second term in the Hamiltonian in the previous equation can be easily eliminated by
a unitary transformation, and the fourth one can be considered as due to the presence of
an electric field
eA2 (ct)
 
eA(ct) dA(ct) d
E(t) = n · =− −n (5.33)
mc2 dt dt 2mc2
treated in the dipole approximation and length gauge. Recasted in the momentum gauge
2 2
and with the term e A2m(ct) eliminated Eq.(5.32) becomes

∂ψN R−N D (r, t)


i~ = HN R−N D ψN R−N D (r, t) (5.34)
∂t
with
P2 e e2
HN R−N D = − P · Aeff (ct) − (5.35)
2m m r
and Aeff identical to that defined by (5.28).

A comparison of the Schrödinger equation in the “R” picture (5.30) with (5.34) shows
that the only difference is the replacing of the Coulomb potential by the distorted VR (r, t)
which includes the higher order retardation corrections and relativistic effects. In order
to get a general idea on the magnitude of the distortion it is instructive to draw graphical
representation of VR (r, t). In Fig.5.3 are presented plots of the distorted potential in the
plane y = 0 at several moments of time, marked on each graph; the external field is the
same as before. For |t| ≥ T /2 the external field in the region near the nucleus is low (see
Fig.5.1) and the distortion effect is negligible, the potential being practically Coulombic.
For −T /2 ≤ t ≤ T /2 the potential starts to deviate from the spherical symmetry, being
periodically stretched along one or the other of the two diagonals of the plane Oxz .

The operators of position and velocity with respect to the laboratory frame in the “R”
picture can be calculated by applying the transformation UT R to the formulae (5.8) and
(5.9), respectively. By direct calculation one finds
−1
RR = R+ (r), (5.36)

" #
Px eAx (Φ+ ) 1 1 ax (Φ+ ) ax (Φ+ )
VR,x = − 2
+ Pz a2 (Φ+ )
+ a2 (Φ+ )
Pz , (5.37)
m m 1 + a (Φ +) 2m 1 + 1 +
2 2 2
" #
Py eAy (Φ+ ) 1 1 ay (Φ+ ) ay (Φ+ )
VR,y = − 2
+ Pz a2 (Φ+ )
+ P ,
a2 (Φ+ ) z
(5.38)
m m 1 + a (Φ +) 2m 1 + 1 +
2 2 2
" #
2 2
e A (Φ+ ) 1 1 1 1
VR,z = + Pz + Pz (5.39)
2m2 c 1 + a2 (Φ+ ) 2m 2
1 + a (Φ+ )
2
1 + a (Φ+ )
2 2 2
The generalized translated Schrödinger equation 71

200 -0.256 200 -0.256

-0.064 -0.064
100 100
-0.016 -0.016
z (au) 0 z (au) 0
-0.004 -0.004

-100 -100

-200 t=-T -200 t = - T/2

-200 -100 0 100 200 -200 -100 0 100 200


x (au) x (au)

200 -0.256 200 -0.256

-0.064 -0.064
100 100
-0.016 -0.016
z (au) 0 z (au) 0
-0.004 -0.004

-100 -100

-200 t = - T/4 -200 t = - T/8

-200 -100 0 100 200 -200 -100 0 100 200


x (au) x (au)

200 -0.256 200 -0.256

-0.064 -0.064
100 100
-0.016 -0.016
z (au) 0 z (au) 0
-0.004 -0.004

-100 -100

-200 t=0 -200 t = T/8

-200 -100 0 100 200 -200 -100 0 100 200


x (au) x (au)

200 -0.256 200 -0.256

-0.064 -0.064
100 100
-0.016 -0.016
z (au) 0 z (au) 0
-0.004 -0.004

-100 -100

-200 t = T/4 -200 t = T/2

-200 -100 0 100 200 -200 -100 0 100 200


x (au) x (au)

Figure 5.3: Plot of the distorted potential VR (r, t) in the plane y = 0 at several moments
of time, marked on each panel. The laser pulse is Gaussian, with parameters E0 = 100
au, ω = 1 au, τp = 1 cycle.
72 The generalized translated Schrödinger equation

with
−1
P = −i~∇, Φ+ = ct − n · R+ (r). (5.40)

Similarly, the probability density of localization with respect to the laboratory frame,
expressed in terms of the wavefunction in the “R” picture is
a2 (φ)
 
2
PR (r, t) = |ψR (R+ (r), t)| 1 + . (5.41)
2

5.3.2 The generalized Schrödinger equation in the “L” picture


The generalized Schrödinger equation in the “L” picture is defined by the unitary trans-
formation

ψL (r, t) = UT L ψT (r, t) (5.42)

with
r
a2 (φ0 ) 0
hr|UT L |r0 i = 1+ δ(r − R+ (r)) (5.43)
2
where
eA(φ)
φ0 = ct − n · r0 , a(φ) = . (5.44)
mc
The equation obeyed by the new wavefunction ψL is
∂ψL (r, t)
i~ = HL ψL (r, t) (5.45)
∂t
with
 !2 !2 
1  eA(φ) 1 1 1 1 a(φ) · P⊥
HL = P⊥ − 2 + Pz 2 + 2 P z + 2

2m 1 + a 2(φ) 2 1 + a (φ)
2
2 1 + a (φ)
2
1 + a 2(φ)

e2 e2 A2 (φ) 1 e2 A2 (φ) 1 e2 A2 (φ) 1


− + Pz + P z − 2
r 4m2 c 1 + a2 (φ) 4m2 c 1 + a2 (φ) 2m

a2 (φ)
2 2 1+ 2

(5.46)

A important feature of the above equation is that it contains the unmodified atomic
2
potential − er . Also the position operator calculated according to

RL = UT+L RT UT L (5.47)

can be easily shown to be

RL = r (5.48)
The generalized translated Schrödinger equation 73

and the density probability of localization is


PL (r) = |ψL (r, t)|2 . (5.49)
We must note that Eq.(5.46) is not unique; any further transformation realized with an
unitary multiplicative operator
 
i
hr|Uf |r i = exp − f (r, t) δ(r − r0 )
0
(5.50)
~
leaves the atomic potential, position operator and the probability density unchanged.
The velocity operator with respect to the laboratory frame in the “L” picture is
   
Px  a2 (φ)  eAx (φ) 1 1  ax (φ) ax (φ)
VL,x = 1 + h − 2 (φ) + Pz h +h Pz 

2 a 2 2
m m 1+ 2m
2
i   2
i 2
i
1 + a 2(φ) 2 1 + a 2(φ) 1 + a 2(φ)
(5.51)
   
Py  a2 (φ)  eAy (φ) 1 1  ay (φ) ay (φ)
VL,y = 1 + − + P + i 2 Pz 

2 a2 (φ) z 2
m m 1+ 2m
 h 2
i   h 2
i h 2
1 + a 2(φ) 2 1 + a 2(φ) 1 + a 2(φ)
(5.52)
 
e2 A2 (φ) 1 1 P · a(φ) 1  1 1
VL,z = 2 (φ) + 2 (φ) + Pz h +h Pz  . (5.53)

2 a a 2 2
2m c 1 + m1+ 2m 2
i 2
i
2 2 1 + a 2(φ) 1 + a 2(φ)

The evolution equations and the expressions of the physical observables position, ve-
locity and density probability with respect to the laboratory frame in the non-relativistic
and relativistic approximations discussed in the Thesis are listed in the Appendix D

5.4 Qualitative study of retardation and relativistic


effects
A complete study of the first-order relativistic effects in the laser-atom interaction can
be done only by numerical integration of the generalized translated Schrödinger equation;
this is the subject of the next Chapter of the Thesis. However, some qualitative but very
useful informations can be found from very simple considerations. In the following we
shall analyze, within the “atomic approximation” defined in the previous Chapter the time
evolution of an atom initially in a bound state interacting with a superintense laser field for
the three approximations discussed in the Thesis: non-relativistic dipole approximation
(NR-D case), non relativistic approximation with the first retardation correction (NR-ND
case), and the relativistic approach (R case). We expect the following general predictions
to be correct in the high frequency regime, and for not extremelly high field intensity.
74 The generalized translated Schrödinger equation

It is well known that in the non-relativistic dipole approximation the electronic


wavepacket moves along the classical nonrelativistic trajectory
Zct
e
α(t) = − A(χ) dχ; (5.54)
mc
−∞

this can be proven by writing the Ehrenfest theorem for the Scrhödinger equation written
in the velocity gauge
 2
e2

∂ψN R−D (r, t) P e
i~ = − P · A(ct) − ψN R−D (5.55)
∂t 2m m r
and neglecting the atomic potential in comparison to the external field. One obtains
dhRN R−D i(t) hpi(t) e
= − A(ct). (5.56)
dt m m
As the mean canonical momentum hpi(t) is small with respect to the external field
hpi(t)  eA at any moment of time, using the initial condition hri(0) = 0 one gets

hRN R−D i(t) ≈ α(t). (5.57)

It follows that for all realistic laser pulses

hRN R−D i(∞) ≈ α(∞) = 0. (5.58)

(see the discussion in the Chapter 2) such that at the end of the laser pulse the wavepacket
will be localized around the nucleus and the survival probability of the atom
X
Psurv = |hun |ΨN R−D (∞)i|2 (5.59)
n

can be considerable large even at very high field intensity, this being the atomic stabiliza-
tion phenomenon.

If the retardation corrections are taken into account (NR-ND case), for large enough
intensity, the behavior of the atomic systems is completely different from the case of
the dipole approximation. By the same method as before one can easily prove that the
Ehrenfest theorem applied to Eq.(5.34) leads to
Zct
e
hRN R−N D i(t) ≈ − dχAeff (χ) ≡ αeff (t). (5.60)
mc
−∞

Note that the electron trajectory in this approximation is identical, up to a sign to the
trajectory of the singularity of the generalized translated potential (5.17)

hRN R−N D i(t) ≈ −rs (t) (5.61)


The generalized translated Schrödinger equation 75

(see Fig.5.2 for a numerical example). At the end of the pulse the electron will be localized
around a point
Z∞ Z∞
e n
hRN R−N D i(∞) = − dχAeff (χ) = dχe2 A2 (χ) (5.62)
mc 2(mc)2
−∞ −∞

which can be very far from the nucleus even for relatively low field intensities (see Eq.(5.22)
for a numerical example). The consequence of this mismatch between the final position
of the electron and the nucleus is the atomic destabilization.
It is important to note that the electron mean velocity, as predicted by this non-
relativistic theory is
dhRN R−N D i(t) e
hVN R−N D i(t) = ≈ − Aeff (ct) (5.63)
dt m
and can take values greater that c if the peak intensity of the external field is large (eA0 &
mc). In fact, the above observation gives the validity limit of the NR-ND approximation:
e
hVN R−N D i(t) ≈ − Aeff (ct) < c, (∀) t ∈ (−∞, ∞). (5.64)
m

The relativistic approach (R case) can be discussed in any of the unitarily equivalent
pictures “generalized space translated”, “R” or “L”; we have chosen the “R” picture since
it allows easily the comparison with the NR-ND approach. First, we note that the two
Hamilton operators (5.25) and (5.35) are almost identical, the only difference being that
in the relativistic case the pure Coulomb potential is replaced by the distorted time-
dependent potential VR (r, t). However, we saw from the numerical example previously
discussed that even for a strong external field (E0 = 100 au) the distortion of the potential
is not too strong (see Fig.5.3). On the other hand, at relativistic intensities the strongest
influence on the system is that of the external field Aeff which drives the wavefunction
along the trajectory αeff (t) defined in (5.60). As a consequence, we can state that the
wavefunctions ψR (r, t) and ψN R−N D (r, t) will be almost identical at any moment of time,
for identical external fields and initial conditions;

ψR (r, t) ≈ ψN R−N D (r, t). (5.65)

We must note, however that although the two wavefunctions are identical, their physical
significance is quite different: while in the non-relativistic case the probability density is
simply the squared modulus of the wavefunction, in the “R” picture one must use the more
complicated expression (5.41). A rough indication on the electron motion as predicted
by the relativistic model can be obtained as follows: using the expression (5.41) of the
probability density

a2 (φ)
 
2
PR (r, t) = |ψR (R+ (r), t)| 1 + (5.66)
2
76 The generalized translated Schrödinger equation

and taking into account the fact that ψR (r, t) is centered about the point
Zct
e
αeff (t) = − dχAeff (χ) (5.67)
mc
−∞

one finds that the probability density will be nonvanishing in a a region centered about
−1
rm (t) = R+ (αeff (t)). (5.68)
Next, using the definition (5.27) of the function R(r) one can easily check that
−1
R+ (αeff (t)) = γ(t) (5.69)
i.e. the electronic wavepacket moves along the classical relativistic trajectory γ(t). This
result can be understood since the system studied is a light atom whose atomic potential
is negligible with respect to the very intense external electromagnetic field. The most
important consequence is that the mean velocity predicted for the electron will be
dγ(t)
hVR i(t) ≈ <c (∀) A0 (5.70)
dt
less than the speed of light for any field, which means that, unlike in the NR-ND case,
the relativistic model is valid at arbitrary intensities.
Another important observation is that at a moment of time t −→ ∞, after the field
has passed, the observables in all approximations discussed here reduce to the field-free
expressions. If one take into account also the fact that the wavefunctions in the R and
NR-ND models are almost identical at any moment of time, it follows that at the end of
the pulse the expectation values of any observable calculated with the NR-ND or R model
will be equal, with very good accuracy. In other words, one can say that although for
high intensity the NR-ND model is “wrong” at intermediate times, it becomes “correct”
again at the end of the pulse.
The qualitative predictions discussed here are confirmed by the numerical simulations
presented in the next Chapter.

5.5 References
[1] A. Bugacov, M. Pont and R. Shakeshaft, Phys. Rev. A 48, R4027 (1993).

[2] N. J. Kylstra, R. A. Worthington, A. Patel, P. L. Knight, J. R. Vázquez de Aldana, and


L. Roso, Phys. Rev. Lett. 85, 1835 (2000).

[3] J. R. Vázquez de Aldana, N. J. Kylstra, L. Roso, P. L. Knight, A. Patel, and R. A.


Worthington, Phys. Rev. A 64, 013411 (2001).

[4] H.G. Muller: Weakly relativistic stabilization : the effect of the magnetic field. In: Super-
Intense Laser-Atom Physics, edited by B. Piraux and K. Rzazewski. (Kluwer, 2001), p.
339-344.
Chapter 6

Numerical method and results

In this Chapter we present numerical solutions of the evolution equation for a Hydrogen
atom in the three approximations discussed in the previous Chapter (non-relativistic
dipole approximation (NR-D case), non-relativistic approximation with first retardation
corection included (NR-ND case), and the relativistic approach (R case). The external
field is a plane wave laser pulse propagating along the Oz direction and linearly polarized
along the Ox direction, described by the vector potential A(φ) ≡ ex A(ct − z).
The NR-D equation written in the laboratory frame and velocity gauge is
 2
e2

∂ΨN R−D (r, t) P e
i~ = − P · A(ct) − ΨN R−D (r, t); (6.1)
∂t 2m m r
for the NR-ND approximation we have the form
 2
e2

∂ΨN R−N D (r, t) P e
i~ = − P · Aeff (ct) − ΨN R−N D (r, t) (6.2)
∂t 2m m r
with
eA2 (ct)
Aeff (ct) = A(ct) − n , (6.3)
2mc
and the evolution equation in the R case will be solved in the “R picture”
 2 
∂ΨR (r, t) P e
i~ = − P · Aeff (ct) + Veff (r, t) ΨR (r, t) (6.4)
∂t 2m m
with
e2
Veff (r, t) = − −1 , (6.5)
R+ (r)

Zφ Zφ
1 n
R+ (r) = r + dχeA(χ) − dχe2 A2 (χ), φ = ct − n · r. (6.6)
mc 2(mc)2
ct ct

77
78 Numerical method and results

6.1 Numerical method


The cases NR-D and NR-ND are relatively simple, as the atomic potential is spherically
symmetric. We have solved them with the code developed by H. G. Muller [1, 2] which
provides very accurate results while the computer resources required are small. The
relativistic case R is much more difficult, mainly due to the fact that the atomic potential
has no spatial symmetry. In the following we present the numerical method developed in
collaboration with H. G. Muller for treating this case.
The obvious choice is to represent the wavefunction as a three-dimensional matrix on
a Cartesian grid with the spatial step h:

Ψ(r, t) −→ Ψ(t), Ψijk (t) = Ψ(ex ih + ey jh + ez kh, t). (6.7)

Similarly, the time dependent effective atomic potential is represented as a three-


dimensional matrix

Veff (r, t) −→ V(t), Vijk (t) = Veff (ex ih + ey jh + ez kh, t). (6.8)

For the propagation we shall use the so called split-operator technique. The total prop-
agation time interval (tin − tf ) is split in N very small equal time steps ∆t, chosen such
that the effective atomic potential Veff (r, t) and the effective field Aeff (t) can be considered
constant along it, and the evolution operator along such time interval is approximated as
i P2 − 2eP · Aeff (ct)
     
i
U (t + ∆t, t) ≈ exp − H(t)∆t = exp − + Veff (r, t) ∆t
~ ~ 2m
i P2
     
i e
≈ exp − Veff (r, t + ∆t)∆t exp − − P · Aeff (c(t + ∆t/2)) ∆t ×
2~ ~ 2m m
 
i
× exp − Veff (r, t)∆t (6.9)
2~
The term “split-operator technique” comes from the previous splitting of the time-
evolution operator in three factors. Obviously, since the momentum operator P and
the atomic potential Veff do not commute, this splitting involves an error which can be
shown to be of the order of (∆t)3 . In the following we shall define the “kinetic” and
“potential” propagators
i P2
   
e
Uk (t; ∆t) = exp − − P · Aeff (ct) ∆t , (6.10)
~ 2m m
 
i
UV (t; ∆t) = exp − Veff (r, t)∆t , (6.11)
~

such that Eq.(6.9) becomes


     
∆t ∆t ∆t
U (t + ∆t, t) ≈ UV t + ∆t; Uk t + ; ∆t UV t; . (6.12)
2 2 2
Numerical method and results 79

One notices that in the successive application of the evolution operator U over consecutive
time-steps the leftmost and the rightmost factors can be combined, such that the total
evolution operator, from tin to tf can be written as
  
∆t
U (tf , tin ) ≈ UV tf ; × Uk (tf − ∆t/2; ∆t) UV (tf − ∆t; ∆t) × . . .
2
  
∆t
× . . . UV (tin + ∆t; ∆t)Uk (tin + ∆t/2; ∆t) UV tin ; (6.13)
2

The advantage of this scheme consists in the fact that the potential evolution operator
UV (t; ∆t) is diagonal in the coordinate representation, while the kinetic evolution operator
Uk (t; ∆t) is diagonal in the momentum representation. The conversion between the two
is done using the not too expensive Fast Fourier Transform algorithm. A schematic
representation of the operator applied to the wavefunction at every time step is given in
Fig.6.1 below: first, the wavefunction is transformed in the momentum representation,
the kinetic evolution operator is applied, and, finally, is converted to the coordinate
representation again and multiplied by the potential part of the evolution operator..

Ψ FFT (Ψ) Ψ U k (Ψ) Ψ IFFT (Ψ) Ψ U V(Ψ)

(1) (2) (3) (4)

FFT = Fourier transform


IFFT = Inverse Fourier transform

Figure 6.1: Sketch of the algorithm used for the propagation over a time step ∆t

6.1.1 The kinetic propagator Uk


The main difficulty related to kinetic part of the propagator is that the function Ψ is a
three dimensional matrix, so both direct and inverse Fourier transforms must be performed
on all the three directions. Since the FFT algorithm requires that the elements of the
transformed vector are stored in consecutive memory locations it follows that for two of the
three directions elements of each “column” must be stored in temporary one-dimensional
vectors which are Fourier transformed, and then copied back to the matrix Ψ. Along
one of the directions the column elements are already located in consecutive positions,
so no auxiliary vector is needed: only the address of the first element of the column is
transferred to the FFT routine. The detailed algorithm corresponding to this standard
scheme is presented in Fig.6.2.
However one must note that the standard scheme is not optimal, as it requires a large
number of operations of copying “columns” of the wavefunction to/from auxiliary vectors
80 Numerical method and results
Numerical method and results 81

Figure 6.3: The algorithm used for the propagation with the minimum number of copying
operations to/from auxiliary vectors.
82 Numerical method and results

which are time consuming. The minimization of the total number of such operations is
obtained1 if one writes the kinetic propagator as
i Px2 i Py2
       
e
Uk (t; ∆t) = exp − − Px Aeff,x (ct) ∆t × exp − ∆t ×
~ 2m m ~ 2m
i Pz2
   
e
× exp − − Pz Aeff,z (ct) ∆t . (6.14)
~ 2m m
Note that the previous expression is exact, since the components of the momentum op-
erator commute; also the order of the three exponentials is arbitrary. Using this rep-
resentation it is possible to apply the Fourier transform, multiplication with the kinetic
propagator and the inverse Fourier transform on each auxiliary vector at once, thus re-
ducing the copying operations by half. Elements of each “column” of each “direction” of
the three dimensional matrix are copied in an auxiliary vector, then the vector is Fourier
transformed and multiplied by the corresponding part of the kinetic evolution operator;
further the vector is Fourier transformed back to the position representation and copied
in the matrix Ψ. The corresponding algorithm for this scheme if given in Fig.6.3.

6.1.2 The potential propagator UV


The application of the potential propagator UV requires only the multiplication of the
wavefunction in the coordinate representation with exponential containing the potential
energy
 
i
Ψ(r, t) → exp − Veff (r, t)∆t Ψ(r, t) (6.15)
~
The difficulty here is that the effective atomic potential is calculated by numerical in-
version of the vectorial function (6.6); even more, it is time dependent so it must be
recalculated at every moment of time. As a consequence, it is important to use a fast
method to calculate Veff (t). Also care must be taken in order to avoid wasting of too
much memory in the process. In following we present the method adopted by us. In order
to calculate the effective potential (6.5) one must solve the equation in R0
R+ (R0 ) = r. (6.16)
Written explicitly for a pulse linearly polarized along the Ox direction and propagating
along Oz the equation becomes
ct−Z
Z 0
1
X0 = x− dχeA(χ), (6.17)
mc
ct
Y 0 = y, (6.18)
ct−Z
Z 0
1
Z0 = z+ dχe2 A2 (χ). (6.19)
2(mc)2
ct
1
The method was proposed by H. G. Muller
Numerical method and results 83

In the previous system the second equation is trivial, in the third one the solution Z 0
depends only on z, and in the first equation the solution X 0 depends on x explicitly and
on z through Z 0 . We shall recast the equation for Z 0 into an equation for Φ0 = ct − Z 0 .
ZΦ0
1
Φ0 = φ − dχe2 A2 (χ), φ = ct − z. (6.20)
2(mc)2
ct

Further it will we written as the differential equation


"  #
dΦ0 e2 A2 Φ0 dΦ0 e2 A2 (ct)
=c− −c (6.21)
dt 2(mc)2 dt 2(mc)2
or
2 2
dΦ0 1 + e2(mc)
A (ct)
2
=c (6.22)
dt e2 A2 (Φ0 )
1 + 2(mc)2

Note that although the previous equation is ordinary, its solution Φ0 depends not only
on t but also on z through the initial condition Φ0 |t=tin = ctin − z. It can be easily
solved numerically and from the solution Φ0 , Z 0 can be obtained. For the case of X 0 the
non-trivial dependence is that on z which appears in the integral. With the notation
ct−Z
Z 0
1
IX = − dχeA(χ), (6.23)
mc
ct

IX obeys the differential equation


 
" # e2 A2 (ct)
eA Φ0 1 + 2(mc)2
 
dIX eA Φ0 dΦ0 eA (ct) eA (ct) 
=− −c = − − . (6.24)
dt m dt m m e A (Φ0 )
2 2
m
1 + 2(mc)2

Here also IX depends on z because Φ0 is z - dependent. In the numerical code we use two
one-dimensional vectors I X and Φ0 , of the same length as the grid along the Oz direction.
They are initialized at the moment tin (assumed very far in the past) as
Φ0,k = ctin − kh, IX,k = 0 (6.25)
and then updated at every time step ∆t according to
e2 A2 (ct)
1+ 2(mc)2
Φ0,i → Φ0,k + c ∆t, (6.26)
e2 A2 (Φ0,k )
1+ 2(mc)2
 
e2 A2 (ct)
1+

eA Φ0,k 2(mc)2 eA (ct) 
IX,k → IX,k −  − ∆t. (6.27)
m e 2 A2 (Φ0,k ) m
1+ 2(mc)2
84 Numerical method and results

The above equations are in fact equivalent to the first order Euler method for solving the
differential equations (6.22) and (6.24); although, obviously, more elaborate methods are
available, this proves to be accurate enough, since the time steps ∆t are however very
small. The effective potential in the point of coordinate r = (ihex + jhey + khez ) is then
calculated as
e2
Vi,j,k = − q (6.28)
2 2 2
X 0,i,k + Y 0,j + Z 0,k
with
X 0,i,k = ih + I0,k , Y 0,j = jh, Z 0,k = ct − Φ0,k (6.29)
In the origin of the reference frame the potential is singular; in the numerical code we use
a very simple cut-off procedure which consists in replacing the singularity in the origin
by a finite value, chosen such that the ground state energy of the Hydrogen atom is
reproduced.

6.1.3 The dual grid method


Even with all the optimization methods previously described the numerical algorithm is
still slow and requires very much memory. One of the reasons is that the spatial step size
must be small (at most 0.5 au) in order to obtain accurate results. On the other hand
the total length of the grid must be large enough to contain the wavefunction which, as
we saw in the previous Chapter moves very fast on very large distances (see Fig.5.2). An
elegant solution for this problem, proposed by H. G. Muller is described in this Section.
The method is based on the observation that the wave-function needs to be described
on a very fine grid only in a small region near the nucleus, where the atomic potential
changes abruptly, while at larger distances a coarse grid would be enough. Since using a
variable step spatial grid would complicate very much the Fourier transform algorithm,
the solution is to use two different grids. The inner grid, centered around the nucleus has
small spatial extension (∼ 10 − 20 au) and very small steps hinner ∼ 0.25 − 0.5 au; it is
embedded in the outer grid, with a coarse mesh (houter ∼ 1 − 2 au) such that their ratio
f = houter /hinner is a power of 2 (we remind here that also the total number of points in
every direction for both grids must be a power of two, as a requirement for applying the
FFT algorithm). A schematic two dimensional representation of the two grids is presented
in Fig.6.4 for the case houter /hinner = 2. The wavefunction Ψ is split between the two grids
Ψ(r, t) = Ψinner (r, t) + Ψouter (r, t); (6.30)
Ψinner (r, t) represented on the fine inner grid is located around nucleus, and Ψouter (r, t),
represented on the outer, coarser grid vanishes near the nucleus. Graphical representation
of the two components of the wavefunction in the plane y = 0 are presented in Fig.6.5.
Note the spherical shape of the inner grid which proves to be more appropriate; a rect-
angular shape (as in the schematic representation in Fig.6.4) would introduce very high
momenta in both part of the wavefunction.
Numerical method and results 85

h outer
h inner

Figure 6.4: Schematic two dimensional representation of the two grids, for the case houter =
2hinner .

For each of the two functions is applied the propagation algorithm as described before.
As the time passes the inner function tends to extend outside its grid, and the outer
function extends toward the center of the outer grid, “filling the hole” near the nucleus.
Periodically, the part of the inner function that moves off its grid is transfered on the
outer grid; similarly, the part of Ψouter coming near the nucleus in transfered on the inner
grid. The transfer from the outer to the inner grid requires interpolation of the transfered
part; similarly, the inverse transfer requires increasing of spatial step, i.e. the “inverse of
interpolation”.
Both these transfers can be done using a procedure based on the Fourier transforma-
tion. Given a function ψ(x), represented on a grid of length N with the step h, we want
to obtain its numerical representation on a grid of length f N , with step h/f , where the
ratio f is a power of 2. The first step is to calculate its Fourier transform, ψ̃(k), which
is copied in a vector of length f N , the positions corresponding to momenta higher than
2π/h being filled with zeroes. The resulting vector is transformed back to the position
representation, the result being the desired interpolation. Similarly, in order to obtain
a coarser representation, with length N/f and step f h, from the original Fourier trans-
formed function ψ̃(k) the momenta higher than 2π/(f h) are eliminated using a smooth
mask, and then the inverse Fourier transformation is performed. In the second case, the
procedure presented here is more accurate than just dropping part of the points in the
position representation, since it allows filtering out the high momenta which would intro-
duce numerical errors on the outer grid. In our code the components of high momenta
which are filtered are monitored; if they are non-negligible this is an indication that the
inner grid has too small spatial extension.
In the dual grid method the transfer between the two grids is the most difficult step.
It may easily lead to very large numerical errors if the two functions Ψinner and Ψouter are
too steep at the borders of the inner grid. However the procedure is very efficient, since a
given ratio f allows a reduction of the computation time and memory by a factor of f 3 .
86 Numerical method and results

1e-04

a 1e-06
1e-08
1e-10
1e-04
1e-06 1e-12
1e-08 1e-14
1e-10
1e-12 1e-16
1e-14 1e-18
1e-16 60
1e-18 60
40
40
20
20
0 0

-20 -20 z (au)


x (au)
-40 -40
-60

0.01
1e-04
b 1e-06
1e-08
1e-10
0.01 1e-12
1e-04 1e-14
1e-06
1e-08 1e-16
1e-10 1e-18
1e-12
1e-14 1e-20
1e-16
1e-18 1e-22
1e-20 60
1e-22 60
40
40
20
20
0 0

-20 -20 z (au)


x (au)
-40 -40
-60

Figure 6.5: Square modulus of the two parts of the wavefunction |Ψouter |2 (a) and |Ψinner |2
(b) in the plane y = 0. The snapshot was taken at the moment t = 1.5 T during the
interaction of the atom with a gaussian pulse with τp = 1 cycle, ω = 1 au and E0 = 50
au, (calculation done within the nonrelativistic dipole approximation).
Numerical method and results 87

6.2 Numerical results


We have numerically integrated Eqs.(6.1), (6.2), (6.4), for the Hydrogen atom initially in
the ground state
Ψ(r, tin ) = u100 (r) (6.31)
and the laser pulse
"  2 #
1.1774 φ ω 
A = ex A(ct − z), A(φ) = A0 exp − sin φ (6.32)
c τp c

with frequency ω = 1 au and duration τp = 1 cycle. For A0 we have chosen values between
10 and 100 a.u. We have calculated the total ionization probability defined as
X
Pion = 1 − Psurv = 1 − |hunlm |Ψ(tf )i|2 (6.33)
n,l,m

where tf is a moment of time after the end of the pulse. Note that for all cases consid-
ered here at the beginning and at the end of the pulse the Hamiltonian reduces to the
unperturbed atomic Hamiltonian
P2 e2
lim H = H0 ≡ − (6.34)
t→tin ,tf 2m r
whose eigenstates are analytically known.
In order to test the accuracy of out method we have run the NR-D and NR-ND cases
with our code (denoted here as R code) and we have compared the results with those
obtained from the non-relativistic code developed by H. G. Muller denoted as HGM code.
We mention that HGM code is extremely accurate, able to produce results with less that
0.1% relative error, which are considered “exact” for our comparison. The calculated
ionization probabilities are presented in the table 6.1. From comparison one can see that
our code is accurate within the limit of 2%, which we considered as acceptable. In the

A0 (au) NR-D NR-D NR-ND NR-ND


(HGM code) (R code) (HGM code) (R code)
10 0.63 0.62 0.64 0.63
25 0.65 0.64 0.74 0.74
50 0.72 0.71 0.97 0.96
100 0.77 0.77 0.99 0.98

Table 6.1: Comparison of the ionization probabilities obtained with the HGM code and
R code for a Gaussian pulse with τp = 1 cycle and ω = 1 au at several peak intensities;
the calculations were done within the NR-D and NR-ND approximations.

dipole approximation it is visible the atomic stabilization even at very high field intensity.
88 Numerical method and results

However, we must note that for large A0 the non-relativistic dipole approximation is not
valid; already at A0 = 25 au the retardation becomes important. For A0 larger than 50 au
its effect is the complete destabilization of the atom. The reason is the mismatch between
the electronic wavepacket driven by Aeff and the atomic nucleus at the end of the pulse
(see the discussion in the previous Chapter).
In the table 6.2 we present the numerical results for the case R, calculated using the
R code. (Note that the HGM code can not be used here since the effective potential

A0 (au) R (R code)
10 0.63
25 0.73
50 0.96
100 0.99

Table 6.2: Ionization probability within the R model for a Gaussian pulse with τp = 1
cycle and ω = 1 au at different peak intensities.

has no spherical symmetry.) By comparing the tables 6.1 and 6.2 one sees that the
differences between the NR-ND and R results are smaller than the precision limit of
our code. This is a confirmation of the general predictions presented in the previous
Chapter. Our conclusion is that is not possible to trace relativistic effects by calculating
ionization probabilities which are overwhelmingly influenced only by retardation2 . Finally,
we present a series of graphical representation of the probability density P(r) in the NR-
ND and R approximations for the laser pulse (6.32) and peak electric field intensity
E0 = 100 a.u. at several moments of time (t = −2T, −T, −T /2, −T /4, 0 in Fig.6.6 and
t = T /4, T /2, T, 2T, 3T in Fig.6.7). In both cases we have represented the probability
density in the plane Oxz
Z∞
P(x, z, t) = dy P(r, t). (6.35)
−∞

One can see that in both cases the wavefunction spreads and its maximum moves along
the trajectory hri(t) (5.60) for the NR-ND case and γ(t) (5.19) in the relativistic case R
(see also Fig.5.2). However, the spreading is rather small (note the logarithmic scale in
the figures) the electron being relativelly well localized at any moment of time. Note the
difference in the form of the density probability in the two approximations. While in the
NR-ND case the electronic wavepacket is almost sherically symmetric, in the relativistic
case it is periodically strecthed along the diagonals of the plane Oxz .
The figures also confirms that at the end of the pulse the two probability densities
are almost identical, i.e. the NR-ND model became “correct”, although it was “wrong”
during the pulse.
2
Still it might be possible to distinguish such effects in other quantities such as the harmonic spectrum,
which were not investigated in this work
Numerical method and results 89

200 0.1 200 0.1

0.01 0.01
150 150

0.001 0.001
100 100
z (au) 1e-04 z (au) 1e-04
50 50
1e-05 1e-05

0 0

t=-2T t=-2T
-50 -50
-100 -50 0 50 100 -100 -50 0 50 100
x (au) x (au)

200 0.1 200 0.1

0.01 0.01
150 150

0.001 0.001
100 100
z (au) 1e-04 z (au) 1e-04

50 50
1e-05 1e-05

0 0

t=-T t=-T
-50 -50
-100 -50 0 50 100 -100 -50 0 50 100
x (au) x (au)

200 0.1 200 0.1

0.01 0.01
150 150

0.001 0.001
100 100
z (au) 1e-04 z (au) 1e-04

50 50
1e-05 1e-05

0 0

t = - T/2 t = - T/2
-50 -50
-100 -50 0 50 100 -100 -50 0 50 100
x (au) x (au)

200 0.1 200 0.1

0.01 0.01
150 150

0.001 0.001
100 100
z (au) 1e-04 z (au) 1e-04

50 50
1e-05 1e-05

0 0

t = - T/4 t = - T/4
-50 -50
-100 -50 0 50 100 -100 -50 0 50 100
x (au) x (au)

200 0.1 200 0.1

0.01 0.01
150 150

0.001 0.001
100 100
z (au) 1e-04 z (au) 1e-04

50 50
1e-05 1e-05

0 0

t=0 t=0
-50 -50
-100 -50 0 50 100 -100 -50 0 50 100
x (au) x (au)

Figure 6.6: Snapshots of density probability P(x, z, t) at the moments of time t =


−2T, −T, −T /2, −T /4, 0. The calculations were done within the NR-ND approxima-
tion (left panel) and R approximation (right panel) for the pulse for a Gaussian pulse
with τp = 1 cycle and ω = 1 au at E0 = 100 au
90 Numerical method and results

200 0.001 200 0.001

150 1e-04 150 1e-04

100 100
1e-05 1e-05
z (au) z (au)

50 50
1e-06 1e-06

0 0

t = T/4 t = T/4
-50 -50
-100 -50 0 50 100 -100 -50 0 50 100
x (au) x (au)

200 0.001 200 0.001

150 1e-04 150 1e-04

100 100
1e-05 1e-05
z (au) z (au)

50 50
1e-06 1e-06

0 0

t = T/2 t = T/2
-50 -50
-100 -50 0 50 100 -100 -50 0 50 100
x (au) x (au)

200 0.001 200 0.001

150 1e-04 150 1e-04

100 100
1e-05 1e-05
z (au) z (au)

50 50
1e-06 1e-06

0 0

t=T t=T
-50 -50
-100 -50 0 50 100 -100 -50 0 50 100
x (au) x (au)

200 0.001 200 0.001

150 1e-04 150 1e-04

100 100
1e-05 1e-05
z (au) z (au)

50 50
1e-06 1e-06

0 0

t=2T t=2T
-50 -50
-100 -50 0 50 100 -100 -50 0 50 100
x (au) x (au)

200 0.001 200 0.001

150 1e-04 150 1e-04

100 100
1e-05 1e-05
z (au) z (au)

50 50
1e-06 1e-06

0 0

t=3T t=3T
-50 -50
-100 -50 0 50 100 -100 -50 0 50 100
x (au) x (au)

Figure 6.7: Snapshots of density probability P(x, z, t) at the moments of time t =


T /4, T /2, T, 2T, 3T . The calculations were done within the NR-ND approximation (left
panel) and R approximation (right panel) for the pulse for a Gaussian pulse with τp = 1
cycle and ω = 1 au at E0 = 100 au
Numerical method and results 91

6.3 References
[1] H. G. Muller, Laser Phys. 9, 138 (1999).

[2] H.G. Muller: Weakly relativistic stabilization : the effect of the magnetic field. In: Super-
Intense Laser-Atom Physics, edited by B. Piraux and K. Rzazewski. (Kluwer, 2001), p.
339-344.
Chapter 7

Outlook and conclusions

In this Thesis several theoretical aspects of the interaction of atoms with intense electro-
magnetic radiation are studied. Everywhere the semiclassical approximation is adopted
and only one-electron systems are considered.

The non-relativistic dipole approximation was considered in Chapter 2. In the case of


monochromatic radiation we presented basic elements of Floquet theory and a numerical
example for the double δ - function potential. We have numerically solved the high-
frequency limit of the Floquet problem and we have discussed the stationary atomic
stabilization and the polychotomic structure of the dressed bound states.
In the second part of Chapter 2 we presented a very systematic study of the dynamic
stabilization of ground state Hydrogen atom in a circularly polarized laser pulse. Us-
ing a very accurate numerical code developed by H. G. Muller we have integrated the
time-dependent Schrödinger equation and have studied the influence of the relevant laser
parameters (pulse shape, duration, frequency) on the dynamic stabilization. Our con-
clusion is that the stabilization of ground state Hydrogen atom in a circularly polarized
laser pulse does take place in a relatively large domain of pulse parameters which can be
obtained with the present experimental achievements. We have also tested the validity
of the adiabatic approximation, with the conclusion that it is applicable even for pulses
as short as a few cycles. Based of the adiabatic approximation we have proposed and
applied a method which allowed us to calculate with very good accuracy the ionization
rate of the Floquet state originating from the field-free ground state in a large domain of
intensities.

Chapter 3 is dedicated to the Volkov solutions (solutions of the Dirac equation for an
electron in an external plane-wave monochromatic field). We presented their derivation,
and also proofs based on direct calculation of the orthogonality (as a generalization of the
proof given by Filipowicz) and of completeness. We have also briefly presented a more
general expression of Volkov solutions introduced by Krstic and Mittleman.

93
94 Outlook and conclusions

In Chapter 4 we proposed a method to include retardation and relativistic effects in


the study of the one-electron atom - laser interaction. Our approach was to define a
relativistic generalization of the space translation method. Such generalization is obtained
by applying a unitary transformation T to the Dirac equation such that the new equation
does not contain the field explicitly, but only through a transformed atomic potential; we
also briefly discussed previous similar approaches presented in the literature.
In our approach, which is appropriate for the case of plane wave pulses, the operator T
is constructed in terms of Volkov solutions and solutions of the Dirac equation for the free
electron written in the Foldy-Wouthuysen representation. Its properties are discussed
in Section 4.2. As the exact equation obtained is very difficult to handle numerically
in Section 4.3 we developed the “atomic low momentum regime approximation” of the
operator T . It as applicable for light atoms, and it allows the relativistic descriptions of
the atomic processes which take place during the interaction with the laser (excitation,
ionization, harmonic generation), but does not include other relativistic effect as pair
creation, emission of high energy photons, whose description is not within the scope of
this thesis.

The resulting evolution equation was studied in Chapter 5; it is a Schrödinger-like


equation named “the generalized space-translated Schrödinger equation”. It contains a
modified potential (“the generalized space-translated potential”) which in some respect
is similar to the translated potential in the non-relativistic Kramers-Henneberger frame.
Its singularity moves along the trajectory of the relativistic classical electron, but with
a different speed and it is also distorted during the interaction with the field. Graphical
representation were presented for the case of a linearly polarized pulse.
In order to reach a more intuitive understanding of the phenomena which take place
we derived in Section 5.3 two unitarily equivalent forms of the the generalized space-
translated Schrödinger equation. The first one, named “the generalized Schrödinger
equation in the R picture” has a very similar form to the non-relativistic Schrödinger
equation with the first retardation correction included and allows to explicitly separate
retardation effects from the pure relativistic ones. The second equivalent form, “the gener-
alized Schrödinger equation in the L picture” explicitly contains the original undistorted
Coulomb potential. Qualitative predictions on the atom behavior were done using the
Ehrenfest theorems. These very simple considerations led us to the conclusion that the
electronic wavepacket moves along the relativistic trajectory of the classical electron, sim-
ilarly to the non-relativistic case. At the end of the pulse the electron will be relatively
well localized very far from the nucleus, the consequence being the atomic destabilization.
We have also qualitatively studied the results obtained in the non-dipole non-
relativistic approximation. Now the conclusion is that for high intensity fields this model is
not applicable, as it predicts superluminal velocities for the electron. However, despite this
contradiction, at the end of the pulse the wave-packet calculated in this non-relativistic
approximation is practically identical to result obtained in the relativistic approach.
Outlook and conclusions 95

In Chapter 6 we presented a numerical method developed in collaboration with H.


G. Muller for solving the generalized translated Schrödinger equation. We have solved
the evolution equation for several intensities of the laser pulse, for the three approxima-
tions considered in this thesis: non-relativistic dipole approximation (NR-D model), non-
relativistic approximation with the first retardation correction included (NR-ND model),
and the relativistic case (R model). The numerical results confirm the general predictions
discussed above. The dipole approximation lead to the atomic stabilization even at very
high intensities, while the inclusion of the retardation destroys the stabilization. The
ionization probabilities calculated within the NR-ND and R models are practically in-
distinguishable. Graphical representations of the electronic wavepacket calculated within
the NR-ND and R approximations at several moments of time during the interaction with
the laser are presented; they also confirm the general predictions discussed in Chapter 5.

A possible continuation of this work would be to include higher order terms in the
atomic approximation of the operator T allowing the study of heavier atoms. Another
possible continuation is extending the formalism for more general external fields than
the plane wave (for example two counter-propagating pulses). Also improvements of the
efficiency and accuracy of the numerical code are required.
Appendix A

Notations and conventions

This Appendix summarizes the general notations and conventions used in the thesis. The
SI units are used.

The electron rest mass and electric charge are denoted by m and e, respectively.

We shall denote a four vector by

a ≡ (a0 ; a1 , a2 , a3 ); (A.1)

the contravariant components bear an upper index, and the covariant ones a lower one. In
order to simplify the formulae, whenever is possible we shall use the alternative notation

{a1 , a2 , a3 } ≡ a ≡ {ax , ay , az }. (A.2)

The four product of two four vectors a and b is denoted by

a · b ≡ a0 b 0 − a · b (A.3)

and the four norm is defined by


2
|a|2 = a · a ≡ a0 − a2 . (A.4)

The coordinate four vector of a particle is

x = (ct; r) ≡ (ct; x, y, z), (A.5)

its four velocity is the derivative of the coordinate with respect to the proper time τ
dx
u≡ (A.6)

and the four momentum is

p · p = (mc)2 .

p = mu, (A.7)

97
98 Notations and conventions

With the usual notations of the Dirac matrices


   
0 σ I 0
α= , β= ≡ γ 0, γ = γ 0 α, γ = (γ 0 , γ) (A.8)
σ 0 0 −I
one defines
â ≡ a · γ = γ 0 a0 − γ · a. (A.9)

Given an unity vector n, the parallel and perpendicular component of an arbitrary


vector a with respect to it are defined by the relations
a = ak + a⊥ ≡ (a · n) n + n × (a × n). (A.10)
One also defines the four vector of zero norm
n = (1, n). (A.11)
and the parallel and perpendicular component of an arbitrary four-vector a with respect
to n
a = ak + a⊥ ≡ a0 ; ak + (0; a⊥ ) .

(A.12)
They obey the relations

n · ak = (n · a) , (n · a⊥ ) = 0 (A.13)
 
ak · a⊥ = 0, (a · a) = ak · ak + (a⊥ · a⊥ ) . (A.14)

An arbitrary electromagnetic plane wave propagating along the n direction can be


described in the Lorentz gauge by the four potential
A (x) ≡ (0, A(ct − n · r)) ≡ (0, A(φ)). (A.15)
where φ is the notation for
φ = n · x = ct − n · r. (A.16)
The Lorentz gauge condition ∂µ Aµ = 0 reads ∇ · A = 0 or n · A = 0.
In all cases considered in this thesis the coordinate system will be chosen such that
n = ez , the vector potential having non-zero components along Ox and Oy. In the case of
linear polarization we take Ox along the polarization direction (A = ex Ax ). The general
form of the vector potential is
 
φ ω 
Ax (φ) = A0 f sin φ . (A.17)
τ cT c
If in the previous equations one takes f ≡ 1 one obtains the case of monochromatic
radiation of frequency ω. In the more general case of a pulse f is a smooth envelope,
usually symmetric about t = 0, with a maximum fmax = 1. The frequency ω is named
Notations and conventions 99

“the pulse frequency” although a pulse is in fact a superposition of monochromatic plane


waves. τ is a parameter proportional to the width, measured in cycles (units of T = 2π ω
),
of the envelope. The full width at half maximum (FWHM) of the envelope will be denoted
by τp .
The Cartesian components of the vector potential for the case of circular polarization
are taken of the form 
φ ω 
Ax (φ) = A0 f sin φ (A.18)
τ cT c
 
cT
Ay (φ) = Ax φ ± (A.19)
4
The above definition of the circular polarization, slightly different from the usual one
 
φ ω 
Ax (φ) = A0 f sin φ (A.20)
τ cT c
 
φ ω 
Ay (φ) = ±A0 f cos φ , (A.21)
τ cT c
is more realistic. This can be easily understood as experimentally a circularly polarized
pulse is obtained starting from a linearly polarized one, passing through a quarter wave-
plate which introduces a delay along one of the axes. However, if the pulse is long enough
φ ω
such that the envelope f τ cT varies slowly with respect to sin c φ the equations (A.18,
A.19) and (A.20, A.21) are practically identical. A numerical example is presented in
Chapter 2.

The electric field which also depends on φ only is given by


∂A(φ) dA(φ)
E(φ) = − = −c . (A.22)
∂t dφ
The parameter E0 = ωA0 , used in order to characterize the laser intensity, it is usually
named “the electric field peak intensity” although for very short pulses it might not be
rigorous equal to the maximum value of the electric field (see the numerical example
presented in Section 2.2.1) .

In the dipole approximation the electromagnetic field depends only on time; the above
expressions of the vector potential reduce to
A(t) = Ax (t) ex (A.23)
 
t
Ax (t) = A0 f sin(ωt) (A.24)

for linear polarization and
A(t) = Ax (t) ex + Ay (t) ey , (A.25)
 
t
Ax (t) = A0 f sin(ωt), (A.26)

Ay (t) = Ax (t ∓ T /4) = ∓A0 f (t ∓ T /4) cos(ωt) (A.27)
100 Notations and conventions

for the circular case, and the electric field is given by


dA(t)
E(t) = − . (A.28)
dt
Appendix B

The Dirac equation for the free


electron

A set of linearly independent solutions of the free Dirac equation


∂u (x)
= c α · P + mc2 β u (x)

i~ (B.1)
∂t
is
E + c α · p + mc2
 
 1 i
ui x, p = p ζi exp − i p · x
(2π~)3/2 2E(E + mc2 ) ~
 
i
≡ ξi (p) exp − i p · x (B.2)
~
with

p 
i = 1, . . . , 4, p ∈ R3 , p≡ (mc)2 + p2 , p

and
  
  1   0
ξ  0  η  1 
ζ1 = ≡
 0 ,
 ζ2 = ≡
 0 ,

0 0
0 0
   
  0   0
0  0  0  0 
ζ3 = ≡
 1 ,
 ζ4 = ≡
 0 ,
 (B.3)
ξ η
0 1

+1, i = 1, 2
i = ; (B.4)
−1, i = 3, 4

101
102 The Dirac equation for the free electron

u1,2 are the positive energy solutions, and u3,4 are the negative energy solutions. They
obey the orthogonality and the completeness relations
Z
0 0
hui x, p |uj x, p i ≡ dr u+ = δ(p − p0 ) δij ,
   
i x, p · uj x, p (B.5)
R3

4 Z
X
dp ui t, r0 , p u+ 0 0
 
i t, r , p = δ(r − r ) I4 . (B.6)
i=1
R3

In Eq.(B.6) above the two coordinate four vectors x and x0 have been explicitly written
as x ≡ (ct, r) and x0 ≡ (ct, r0 ) in order to emphasize that they are taken at the same
moment of time.
The projectors on the positive/negative energy subspaces are
X Z  +  X Z
dp ξi p ξi+ p
 
P+ = dp ξi p ξi p , P− = (B.7)
i=1,2 i=3,4
R3 R3

The free particle solutions of the Dirac equation in the Foldy-Wouthuysen [1] represen-
tation
∂vi (x, p) √
i~ = β c2 P2 + m2 c4 vi (x, p) (B.8)
∂t
are
 
 1 i
vi x, p = ζi exp − i p · x . (B.9)
(2π~)3/2 ~
and they also form a complete and orthogonal basis. The solutions vi (x, p) can be obtained
from the solutions in the standard representation ui (x, p) using the unitary operator

EP + mc2 + β cα · P √
S= , EP = c2 P2 + mc2 (B.10)
[2EP (Ep + mc2 )]1/2

B.1 References
[1] L.L. Foldy and S.S. Wouthuysen, Phys. Rev. 78, 26 (1950).
Appendix C

Classical motion of an electron in a


laser pulse

The relativistic classical motion of a particle interacting with an electromagnetic plane


wave neglecting the radiation reaction is a textbook problem. A solution obtained in the
Hamilton-Jacobi formalism for the monochromatic case is given in the book by Landau
and Lifshitz [1]. Also a number of papers published in the last four decades considered this
problem, especially in connection with the problem of radiation scattering by electrons.
We mention here the papers by Brown and Kibble [2], Eberly and Sleeper [3], Sarachik
and Shappert [4] all of them treating the case of an arbitrary plane wave pulse in the
Hamilton-Jacobi formalism. Recently Salamin and Faisal [5] have presented the solution
for an arbitrary plane wave pulse and very general initial conditions. The Lagrange
formalism have been used by Hartemann et al [6] in relation with the problem of the
so-called ponderomotive scattering of electrons.
In this Chapter we use also the Lagrange formalism, and will present the equations
of motion for general initial conditions and arbitrary plane wave laser pulses; also, in
addition to the relativistic solution, the non-relativistic approximation will be studied.
Numerical examples will be given for two particular cases of interest. The notations are
those presented in the Appendix A.

C.1 Initial conditions


We shall assume that at a given moment of time t0 the particle is in the origin of the
reference frame, and it has the given velocity v0 ; in the relativistic case the corresponding
four-velocity will be denoted by u0 , and the four-momentum by p0 . The field propagates
along the n direction, chosen parallel to Oz, and is described by the potential A (A.15).
We denote by A0 the value of the potential A in the origin of the reference frame at the
initial moment t0

A0 ≡ (0, A0 ) = A(φ0 ) (C.1)

103
104 Classical motion of an electron in a laser pulse

with φ0 = ct0 . For the case of a pulse, the most natural choice for t0 is a moment
sufficiently far in the past, when the pulse has not reached the origin and the particle is
free (i.e. A0 = 0). The general equations of motion will be written for arbitrary values of
t0 and v0 , numerical results will be presented for two cases of interest:
Case 1: linearly polarized pulse; the particle is at rest in the origin at a moment t0
very far in the past, when the pulse has not yet reached the origin.
Case 2: linearly polarized monochromatic field; at the moment t0 when the field
vanishes in the origin, the particle is located in the origin, with an initial velocity v0
directed along the field propagation direction.

C.2 Non-relativistic case


The Lagrange function of a nonrelativistic particle of mass m and electric charge e moving
in the electromagnetic field (A.15)
mṙ2
L(r, ṙ, t) = + eṙ · A(φ) (C.2)
2
gives the equations of motion
d
(mṙ⊥ + eA(φ)) = 0, (C.3)
dt
d ∂A(φ) dA(φ)
(mż) = eṙ · = −eṙ · , (C.4)
dt ∂z dφ
where r⊥ is the component of r orthogonal on the propagation direction n ≡ ez . From
the first of the two equations above, and taking into account the initial conditions one
obtains
e
ṙ⊥ = v0⊥ − (A(φ) − A0 ). (C.5)
m
Using this result in Eq.(C.4) one gets
 2  2 
d d e m
ż = A(φ) − A0 − v0⊥ (C.6)
dt dφ 2m2 e
It is convenient to use φ as the independent variable. From the relation
 
dφ ż
=c 1− (C.7)
dt c
one can see that this change of variable only makes sense if ż < c; the restriction is
however not a problem, since we are not interested, anyway, to find a solution with ż ≥ c.
The above equation in the new variable φ writes as
d ż 1 ż 2 e2 
   2 
d m
− = A(φ) − A0 − v0⊥ . (C.8)
dφ c 2 c2 dφ 2(mc)2 e
Classical motion of an electron in a laser pulse 105

Taking into account the initial conditions one obtains the equation
ż 1 ż 2 e2 A2 (φ) v03 1 v03 2
− 2 = + − , (C.9)
c 2c 2(mc)2 c 2 c2
with
2mv0⊥
A2 (φ) = (A(φ) − A0 )2 − · (A(φ) − A0 ) (C.10)
e
whose solution is
s
ż  v03 2 e2 A2 (φ)
=1− 1− − . (C.11)
c c (mc)2
One notices that the solution ż is defined only if
e2 A2 (φ)  v03 2
< 1− (C.12)
2(mc)2 c
and that ż cannot become larger than c. Again, using φ as the independent variable, one
gets
dz 1 dz dr⊥ 1 dr⊥
= dz
, = (C.13)
dφ c − dt dt dφ c − dzdt
dt
or
r
2 e2 A2 (χ)
ct−z
Z 1− 1 − v03
c
− (mc)2
z= dχ r , φ0 = ct0 (C.14)
2
1 − v03 − e A (χ)
 2 2
φ
0
c (mc)2
ct−z
A(χ) − A0 − mve ⊥
Z
e
r⊥ = − dχ r . (C.15)
mc 2 2
2
A (χ)
1 − v03 − e (mc)

φ0 2
c

The first of the above equations must be solved for z, then its solution used in the second
one, to get r⊥ . In the following we shall discuss the two particular cases presented in
Section C.1.

Case 1 is described by the conditions: φ0 = −∞, A0 = 0, v0 = 0, the corresponding


trajectory being:
q
ct−z 2 A2 (χ) ct−z
Z 1 − 1 − e (mc) 2 e
Z
A(χ)
z= dχ q , r⊥ = − dχ q . (C.16)
1 −
2 2
e A (χ) mc 1 − e2 A2 (χ)
−∞ (mc)2 φ0 (mc)2

The solution is defined for e2 A2 (φ) < (mc)2 ; in fact, the non-relativistic approximation
is valid only in the limit e2 A2 (φ)  (mc)2 . In this case the previous solution becomes
ct−z ct−z
e2 A2 (χ)
Z Z
e
z= dχ , r⊥ = − dχA(χ). (C.17)
2(mc)2 mc
−∞ φ0
106 Classical motion of an electron in a laser pulse

In the case 2 we have φ0 = 0, A0 = 0, v0 = v03 n, and the corresponding solution


q 2 e2 A2 (χ)
ct−z ct−z
Z 1− 1 − v03
c
− (mc)2 e
Z
A(χ)
z= dχ q , r ⊥ = − dχ q .
1− cv03 2
 e2 A2 (χ)
− (mc)2 mc 1− cv03 2
 e2 A2 (χ)
− (mc)2
0 0

(C.18)
In this case the conditions of validity of the non-relativistic approximation are e2 A2 (φ) 
(mc)2 and v03  c, which leads to the equations of motion
ct−z ct−z
v03 e2 A2 (χ)
Z   Z
e
z= dχ + , r⊥ = − dχA(χ). (C.19)
c 2(mc)2 mc
−∞ φ0

C.3 Relativistic case


In the relativistic case the equations of motions are
dpµ dx
= eF µν uν , u= , p = mu, (C.20)
dτ dτ
where τ is the proper time and F̂ is the electromagnetic four-tensor of the field intensities:
dA1 (φ) dA2 (φ)
0 dφ dφ
0

1 1
− dAdφ(φ) 0 0 − dAdφ(φ)
F µν = ∂ µ Aν − ∂ ν Aµ ≡ . (C.21)
2 2
− dAdφ(φ) 0 0 − dAdφ(φ)

dA1 (φ) dA2 (φ)


0 dφ dφ
0
With the above definitions the equations of motion become
dp0 dp3 dA(φ)
= = −e · u⊥ , (C.22)
dτ dτ dφ
dp⊥ dA(φ) 0
u − u3 .

= −e (C.23)
dτ dφ
From the first of the above formulae one obtains
d(p0 − p3 )
=0 (C.24)

i.e. p0 − p3 = m(u0 − u3 ) is a constant of motion; taking into account the initial conditions
we have p0 − p3 = m(u00 − u30 ). Further, noticing that
d  dφ
u0 − u3 = const = x 0 − x3 = (C.25)
dτ dτ
Classical motion of an electron in a laser pulse 107

Eq.(C.23) becomes

dp⊥ dA dφ
= −e . (C.26)
dτ dφ dτ

It is again convenient to look for the solutions of the above equations as functions of φ.
Using the initial conditions one obtains
 p0⊥ 
p⊥ (φ) = −e A(φ) − A0 + , (C.27)
e

and, then, from Eqs.(C.22) and (C.27)

dp3 (φ) e2 d p0⊥ 2


= A(φ) − A0 + (C.28)
dφ 2m(u00 − u30 ) dφ e

or

3 e2 A2 (φ)
p (φ) = + p03 , (C.29)
2m(u00 − u30 )

with

2p0⊥
A2 (φ) = (A(φ) − A0 )2 − · (A(φ) − A0 ) . (C.30)
e

The differential equations in φ obeyed by the coordinates are

dr 1 dr p(φ)
= 0 3
= (C.31)
dφ (u0 − u0 ) dτ m(u00 − u30 )

with the solution


ct−z
Z
e h p0⊥ i
r⊥ = − A(χ) − A0 − dχ (C.32)
m(u0 − u30 )
0
e
φ0
ct−z
e2 A2 (χ) u30
Z  
z= dχ + . (C.33)
2m2 (u00 − u30 )2 u00 − u30
φ0

The above equations are equivalent to those presented by Salamin and Faisal ([5], Eq.27).
As in the non-relativistic case, the above equations can be solved only numerically; still,
it is possible to find some properties of the solutions for simple pulses; we shall study the
same particular cases as in the previous Section.
108 Classical motion of an electron in a laser pulse

In the particular case 1 the relativistic solution reduces to


ct−z
Z
e
r⊥ = − A(χ)dχ, (C.34)
mc
−∞
ct−z
Z
e
z= A2 (χ)dχ. (C.35)
2(mc)2
−∞

In the low intensity limit the above equation become identical with the non-relativistic
result (C.17). Another observation is that the total displacement in the polarization plane
vanishes (see the discussion in Chapter 2), so that at the end of the pulse the particle
is left at rest in a point along the Oz axis. If the envelope is simple enough it is also
possible to calculate the total displacement along the propagation direction; for example,
for a Gaussian pulse of amplitude A0 , frequency ω and FWHM τp one gets
r " " 2 ##
2 2

e A0 π T τp 1 2πτp
∆z = 2
1 − exp − (C.36)
4m c 2 1.1774 2 1.1774

For all realistic cases the second term in the previous equation is negligible, and the total
displacement becomes
r
e2 A20 π T τp
∆z = . (C.37)
4m2 c 2 1.1774
We remind here that the advance along the propagation direction does not appear in the
dipole approximation, being a consequence of the retardation.

In the case 2 the solution is


ct−z
Z
e
r⊥ = − dχA(χ) (C.38)
m(u0 − u30 )
0
0
ct−z
e2 A2 (χ) u30
Z  
z= dχ + . (C.39)
2m2 (u00 − u30 )2 u00 − u30
0

Again, in the non-relativistic limit u00 = c, u30 = v03 , v03  c the previous solution
becomes identical with the nonrelativistic one (C.19). If the initial velocity u0 vanishes the
trajectory has a “Z” like shape - an oscillation along the polarization direction, composed
with an advance along the propagation direction. It is possible to find a particular “initial”
velocity v0 such that the total displacement along the Oz axis during one optical period
cancels. The condition writes as
e2 A20 (φ) u30
+ =0 (C.40)
4m2 (u00 − u30 )2 (u00 − u30 )
Classical motion of an electron in a laser pulse 109

and it is satisfied for the initial velocity


e2 A20
v0 = −nc ≡ nV0 (C.41)
4(mc)2 + e2 A20
With this initial condition, and using the notation
e2 A20
m∗2 = m2 + (C.42)
2c2
the solution becomes
ct−z ct−z
e2 A20
Z Z  
e 2
r⊥ = − ∗ dχA(χ), z= dχ A (χ) − . (C.43)
mc 2(m∗ c)2 2
0 0

In this case the trajectory has the well known “figure 8” shape in the plane Oxz. Is
important to notice that, unlike in the case 1 the amplitude of the oscillation is limited
along both Ox and Oz directions when A0 tends to infinity; the limits are
√ c c
lim ∆x = 2 , lim ∆z = . (C.44)
A0 →∞ ω A0 →∞ 4ω
The quantity m∗ defined in Eq.(C.42) is the so-called “dressed mass”; it is an average
effective mass of the electron interacting with the monochromatic field. One can prove
[5] that it satisfy the average energy-momentum relation
hE 2 i = (m∗ c2 )2 + c2 hPcan
2
i (C.45)
where Pcan is the canonical momentum of the electron.

C.4 Numerical example


In this Section numerical examples are presented for several cases of interest. We shall
consider a linearly polarized Gaussian pulse with ω = 1 and FWHM τp = 2. The value of
the vector potential in the origin of the reference frame A(0, t) is represented in Fig.C.1
as a function of time.
For this pulse, we have represented in Fig.C.2 the relativistic and non-relativistic
trajectories for three values of A0 : A0 = 10, A0 = 40, A0 = 100 and initial conditions
corresponding to the case 1. At the lowest intensity the two calculations give almost
identical results, but as A0 increases the relativistic corrections become more and more
important. The total displacement along the propagation direction, (which does not exist
in the dipole approximation) is significant even all relatively small intensities; the values
are in agreement with the formula (C.37). In Figs.C.3 and C.4 are presented results for
the monochromatic case, described by Eqs.(C.38), (C.39), for two initial velocities. The
former contains the “Z” like trajectory obtained for zero initial velocity v0 = 0, for a
linearly polarized monochromatic field of frequency ω = 1 au and amplitude A0 = 10 au.
The initial condition (C.41) was considered in Fig.C.4 for the same frequency and several
field amplitudes from A0 = 50 au up to A0 = 5000 au. It is clearly visible the tendency
of the spatial amplitudes of the motion to level off at the values given by Eq.(C.44).
110 Classical motion of an electron in a laser pulse

τp = 2

0.5
A(0,t) / A0

-0.5

-1
-6 -4 -2 0 2 4 6
t/T

Figure C.1: The vector potential A in the origin of the reference frame, for the case of a
linearly polarized Gaussian pulse of τp = 2, ω = 1 au.

350
2.5
A0 = 10 A0 = 40 A0 = 100
40
300

250
30

1.5 200
z

20
150
1

100

relativistic 10 relativistic relativistic


non-relativistic non-relativistic non-relativistic
0.5
50

0 0 0
-10 -5 0 5 10 -40 -20 0 20 40 -100 0 100
x x x

Figure C.2: Case 1: the relativistic (in black) and non-relativistic (in red) trajectories
for a lineraly polarized pulse with τp = 2 cycles, ω = 1 au and for three values of the
amplitude A0 .
Classical motion of an electron in a laser pulse 111

A0 = 10 au
z (au)

0
-10 -5 0 5 10
x (au)

Figure C.3: Case 2, initial condition v0 = 0: the relativistic trajectory for monochromatic
linearly polarized field of amplitude A0 = 10 au.

40
A0 = 5000 au

A0 = 1000 au
20

A0 = 100 au
z (au)

A0 = 50 au
-20

-40
-200 -100 0 100 200
x (au)

Figure C.4: Case 2, initial condition v0 = V0 n: the relativistic trajectories for monochro-
matic linearly polarized field and four values of the amplitude A0 .
112 Classical motion of an electron in a laser pulse

C.5 References
[1] L. D. Landau and L. M. Lifshitz, Classical theory of fields, Addison-Wesley, Reading,
Mass, 1951.

[2] L. S. Brown and T. W. B. Kibble, Phys. Rev. 133, A705 (1964).

[3] J. H. Eberly and A. Sleeper, Phys. Rev. 176, 1570 (1968).

[4] E. S. Sarachik and G. T. Shappert, Phys. Rev. D 1 2738 (1970).

[5] Y. I. Salamin and F. H. M. Faisal, Phys. Rev. A 54, 4383 (1996).

[6] F. V. Hartemann, S. N. Fochs, G. P. Le Sage, N. C. Luchmann, Jr, J. G. Woodworth,


M. D. Perry, Y. J. Chen and A. K. Kerman, Phys. Rev. E 51, 4833 (1995).
Appendix D

Summary of relativistic and


non-relativistic approximations
discussed

In this Appendix we list the evolution equations and the expressions of the observables
position (R), velocity (V) and probability density with respect to the laboratory frame
(P(r)) used in the Thesis. The external field is described by the vector potential A(φ) as
discussed in Chapter 4.

D.1 Non-relativistic dipole approximation


D.1.1 Laboratory frame, velocity gauge

∂ψN R−D (r, t)


i~ = HN R−D ψN R−D (r, t) (D.1)
∂t
P2 e e2
HN R−D = − P · A(ct) − (D.2)
2m m r
RN R−D = r (D.3)
P − eA(ct)
VN R−D = (D.4)
m
PN R−D (r) = | ψN R−D (r, t) | 2 . (D.5)

D.1.2 Kramers-Henneberger frame

∂ψKH (r, t)
i~ = HKH ψKH (r, t) (D.6)
∂t
Zct
P2 e2 e
HKH = − , α(t) = − A(χ)dχ (D.7)
2m |r + α(t)| mc

113
114 Summary of approximations

RKH = r + α(t) (D.8)


P − eA(ct)
VKH = (D.9)
m
PKH (r) = | ψKH (r − α(t), t) | 2 . (D.10)

D.2 Non-relativistic approximation, first order re-


tardation corection included
D.2.1 Laboratory frame, velocity gauge

∂ψN R−N D (r, t)


i~ = HN R−N D ψN R−N D (r, t) (D.11)
∂t
P2 e e2 eA2 (φ)
HN R−N D = − P · Aeff (ct) − , Aeff (φ) = A(φ) − n (D.12)
2m m r 2mc
RN R−N D = r (D.13)
P − eAeff (ct)
VN R−N D = (D.14)
m
PN R−N D (r) = | ψN R−N D (r, t) | 2 . (D.15)

D.3 Relativistic approach


D.3.1 Generalized Kramers-Henneberger picture

∂ψT (r, t)
i~ = HT ψT (r, t) (D.16)
∂t
P2 e2
HT = − ,
2m |R−1 + (r)|
Zφ Zφ
1 n
R+ (r) = r + dχeA(χ) − dχe2 A2 (χ) (D.17)
mc 2(mc)2
−∞ −∞

RT = R−1
+ (r) (D.18)
" #
Px eAx (Φ+ ) 1 1 ax (Φ+ ) ax (Φ+ )
VT,x = − a 2 (Φ ) + Pz a 2 (Φ ) + 2 Pz (D.19)
m m 1+ 2 + 2m 1+ 2 +
1 + a (Φ2
+)

" #
Py eAy (Φ+ ) 1 1 ay (Φ+ ) ay (Φ+ )
VT,y = − a2 (Φ ) + Pz a2 (Φ ) + 2 Pz (D.20)
m m 1+ 2 + 2m 1+ 2 +
1 + a (Φ2
+)

" #
2 2
e A (Φ+ ) 1 1 1 1
VT,z = 2 a2 (Φ ) + Pz a 2 (Φ ) + 2 Pz (D.21)
2m c 1 + +
2
2m 1+ +
2
1 + (Φ+ )
a
2
Summary of approximations 115

Φ+ = ct − n · R−1
+ (r) (D.22)

a2 (φ)
 
PT (r) = 1 + | ψT (R+ (r), t) | 2 . (D.23)
2

D.3.2 The “R” picture

∂ψR (r, t)
i~ = HR ψR (r, t) (D.24)
∂t
P2 e e2 eA2 (t)
HR = − P · Aeff (ct) − −1 , Aeff (t) = A(t) − n (D.25)
2m m |R+ (r)| 2mc
Zφ Zφ
1 n
R+ (r) = r + dχeA(χ) − dχe2 A2 (χ)
mc 2(mc)2
ct ct
−1
RR = R+ (r) (D.26)
" #
Px eAx (Φ+ ) 1 1 ax (Φ+ ) ax (Φ+ )
VR,x = − 2 (Φ )
+ Pz 2 (Φ )
+ 2
Pz (D.27)
m m 1+ 2+ a 2m 1+ 2+ a
1 + a (Φ +)
2
" #
Py eAy (Φ+ ) 1 1 ay (Φ+ ) ay (Φ+ )
VR,y = − 2
+ Pz a2 (Φ+ )
+ P
a2 (Φ+ ) z
(D.28)
m m 1 + a (Φ +) 2m 1 + 1 +
2 2 2
" #
2 2
e A (Φ+ ) 1 1 1 1
VR,z = + Pz + Pz (D.29)
2m2 c 1 + a2 (Φ+ ) 2m 2
1 + a (Φ+ )
2
1 + a (Φ+ )
2 2 2
−1
Φ+ = ct − n · R+ (r) (D.30)

a2 (φ)
 
PR (r) = 1 + | ψR (R+ (r), t) | 2 . (D.31)
2

D.3.3 The “L” picture

∂ψL (r, t)
i~ = HL ψL (r, t) (D.32)
∂t 
!2 !2 
1  eA(φ) 1 1 1 1 a(φ) · P⊥
HL = P⊥ − 2 + Pz 2 (φ) + 2 (φ) Pz + 2

2m 1 + a 2(φ) 2 1+ a
2
21+ a
2
1 + a 2(φ)

e2 e2 A2 (φ) 1 e2 A2 (φ) 1 e2 A2 (φ) 1


− + Pz + P z − 2
r 4m2 c 1 + a2 (φ) 4m2 c 1 + a2 (φ) 2m

a2 (φ)
2 2 1+ 2

(D.33)
116 Summary of approximations

RL = r (D.34)
 
Px  a2 (φ)
VL,x = 1 + 2  − (D.35)

m
 
a2 (φ)
1+ 2
 
eAx (φ) 1 1  ax (φ) ax (φ)
− 2 (φ) + Pz  2 +   2 Pz 

m 1+ a 2m 2 2
2 1 + a 2(φ) 1 + a 2(φ)
 
Py  a2 (φ)
VL,y = 1 + 2  − (D.36)

m
  2
1 + a 2(φ)
 
eAy (φ) 1 1  ay (φ) ay (φ)
− 2 (φ) + Pz  + Pz 

a 2 2
m 1+ 2m 2
 2

2 1 + a 2(φ) 1 + a 2(φ)
 
2 2
e A (φ) 1 1 P · a(φ) 1  1 1
VL,z = + + Pz  2 +  2 P z 

2 2
2m2 c 1 + a (φ) m 1 + a (φ) 2m a2 (φ) a2 (φ)
2 2 1+ 2
1+ 2

(D.37)

PL (r) = | ψL (r, t) | 2 . (D.38)

You might also like