11
This content has been downloaded from IOPscience. Please scroll down to see the full text. Download details: IP Address: 130.235.188.129 This content was downloaded on 14/12/2014 at 13:19 Please note that terms and conditions apply. Measurements of relative photoemission time delays in noble gas atoms View the table of contents for this issue, or go to the journal homepage for more 2014 J. Phys. B: At. Mol. Opt. Phys. 47 245602 (http://iopscience.iop.org/0953-4075/47/24/245602) Home Search Collections Journals About Contact us My IOPscience

Measurements of relative photoemission time delays in noble gas atoms

Embed Size (px)

Citation preview

This content has been downloaded from IOPscience. Please scroll down to see the full text.

Download details:

IP Address: 130.235.188.129

This content was downloaded on 14/12/2014 at 13:19

Please note that terms and conditions apply.

Measurements of relative photoemission time delays in noble gas atoms

View the table of contents for this issue, or go to the journal homepage for more

2014 J. Phys. B: At. Mol. Opt. Phys. 47 245602

(http://iopscience.iop.org/0953-4075/47/24/245602)

Home Search Collections Journals About Contact us My IOPscience

Measurements of relative photoemissiontime delays in noble gas atoms*

D Guénot1, D Kroon1, E Balogh2, E W Larsen1, M Kotur1, M Miranda1,T Fordell1, P Johnsson1, J Mauritsson1, M Gisselbrecht1, K Varjù2,C L Arnold1, T Carette3,6, A S Kheifets7, E Lindroth5, A LʼHuillier1 andJ M Dahlström3,4,5

1Department of Physics, Lund University, PO Box 118, SE-22100 Lund, Sweden2Department of Optics and Quantum Electronics, University of Szeged, Dóm tér 9, 6720 Szeged, Hungary3Department of Physics, Stockholm University, AlbaNova University Center, SE-106 91 Stockholm,Sweden4Center for Free-Electron Laser Science, Luruper Chaussee 149, D-22761 Hamburg, Germany5Max Planck Institute for the Physics of Complex Systems, Noethnitzerstr. 38, D-01187 Dresden, Germany6 Laboratoire de Chimie quantique et photophysique, CP160/09, Université Libre de Bruxelles, B 1050Brussels, Belgium7Research School of Physical Sciences, The Australian National University, Canberra ACT 0200, Australia

E-mail: [email protected]

Received 12 May 2014, revised 19 October 2014Accepted for publication 24 October 2014Published 3 December 2014

AbstractWe determine relative photoemission time delays between valence electrons in different noblegas atoms (Ar, Ne and He) in an energy range between 31 and 37 eV. The atoms are ionized byan attosecond pulse train synchronized with an infrared laser field and the delays are measuredusing an interferometric technique. We compare our results with calculations using the randomphase approximation with exchange and multi-configurational Hartree–Fock. We alsoinvestigate the influence of the different ionization angular channels.

Keywords: photoemission time delay, attosecond, RABITT, MCHF, RPAE

(Some figures may appear in colour only in the online journal)

1. Introduction

The photoionization of atoms and molecules has beenextensively studied during the last decades using various lightsources, such as lasers and synchrotron radiation, towardincreasing precision and completeness (see e.g. [1] andreferences therein). One of the goals has been in particular toperform ‘complete’ experiments, which means that the partialamplitudes and the relative phases between the ionizationchannels are determined [2–4]. The dynamics between

different channels of the photoionization process is thencaptured. Such experiments often require alignment of theatoms or molecules prior to photoionization, as well asangular resolution and/or light polarization control.

An important application of attosecond pulses [5, 6] thathas arisen during the last few years is the measurement of thephotoemission time delay, equal to the derivative of the phaseof the ionization transition matrix element with respect to theenergy [7–9]. This phase is not the same as the relative phasebetween the different channels. It is a new experimentalquantity which can be (and should be) compared with pre-dictions from theoretical models. The photoionization timedelay can be interpreted as the group delay of the outgoingionized electronic wave packet created by the absorption of anattosecond pulse, as it propagates in the atomic potential[10, 11]. In the case of a single outgoing angular momentumchannel, the phase of the photoemission amplitude is identical

Journal of Physics B: Atomic, Molecular and Optical Physics

J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 245602 (10pp) doi:10.1088/0953-4075/47/24/245602

* Intended contribution to the ‘Special issue on ultrafast electron andmolecular dynamics’ available at http://iopscience.iop.org/0953-4075/47/12.

Content from this work may be used under the terms of theCreative Commons Attribution 3.0 licence. Any further

distribution of this work must maintain attribution to the author(s) and thetitle of the work, journal citation and DOI.

0953-4075/14/245602+10$33.00 © 2014 IOP Publishing Ltd Printed in the UK1

to the scattering phase [12, 13], while in a more general case,where two or more angular channels are present, it carriesinformation on the different channels involved in the photo-emission. Measurements of the photoionization time delaysallow us to get insight and understanding about the temporaldynamics of photoionization from any subshell and shouldultimately be combined with other types of measurements toprovide a complete picture.

One of the first time-resolved photoemission measure-ments was that of the difference in photoemission delays fromthe 2s and 2p shells in neon [8]. Using the streaking techniqueand single attosecond pulses centered around 80 eV, the dif-ference was found to be 21 ± 5 as. The interpretation of thisresult triggered an intense theoretical activity [8, 14–19] (seealso earlier theoretical work [20, 21]). Other experiments,based on interferometry with attosecond pulse trains (APT)centered at ≈30 eV, studied resonance effects in N2 [22, 23]and He [24] or compared photoemission time delays betweenthe 3s and 3p shells in Ar [9, 25]. The latter results alsostimulated several theoretical studies which aimed at under-standing the effect of the probe laser field on the measurement[11, 26] as well as accurately describing the time delays in anenergy region where strong 3s–3p intershell correlationeffects take place [26–30]. Recently, this approach has beenextended to the case of double photoemission in xenon [31],using coincidence measurements.

In this work, we present measurements of the differencesin photoionization time delays between outer shell electronsin helium, neon and argon. The time delays are extracted withan interferometric technique, using a frequency comb of high-order harmonics (with photon energies from 29 to 39 eV) andan infrared (IR) laser pulse. We make use of an activelystabilized interferometer, which allows us to precisely controlthe length difference between its two arms, and thus the delaybetween the IR and extreme ultraviolet (XUV) pulses. Theexperimental results are compared with calculations usingrandom phase approximation with exchange (RPAE) andmulti-configurational Hartree–Fock (MCHF) approaches andincluding contributions from several possible angular chan-nels. In section 2, we present the experimental setup and the

results. In section 3, we describe our calculations and com-pare the results with the experiment in section 4.

2. Experiment

2.1. Experimental setup

A scheme of the experimental setup is shown in figure 1(a).We use a Ti:saphire femtosecond laser system delivering20 fs, 5 mJ pulses centered at 800 nm, with 1 kHz repetitionrate. The laser is sent in two arms of a Mach–Zendherinterferometer. In one arm, the beam is focused in a gas cellwhere an XUV APT is created via high-order harmonicgeneration. A 200 nm Al filter is used to remove the funda-mental radiation. In the second arm the pulse is delayedrelative to the APT by a piezoelectric-controlled delay stage,and then is recombined with the APT by a reflection on aconvex holey mirror. The APT and IR pulses are focused witha toroidal mirror into a magnetic bottle electron spectrometerwhere photo-electrons are collected in a solid angle of π2steradians. Both XUV and IR light fields are polarized ver-tically, i.e. perpendicularly to the electron detection axiswhich is horizonal. Although this choice can lead to a col-lection anisotropy whose effect is angular channel-dependent,the resulting change in the measured delays is within ourerror bars.

2.2. Interferometric measurements

Following ionization with an APT, electrons are created atdiscrete energies corresponding to absorption of odd harmo-nics, separated by ω2 , where ω is the IR frequency and ℏ thereduced Planck constant. The addition of a weak IR fieldallows further absorption or emission of one IR photonleading to sideband (SB) peaks in the electron spectra (seefigure 2(a)). Each SB can be created through two differentpathways so that the electron signal varies as [33, 34]

ωτ Δϕ Δϕ= + − −( )S A B cos 2 , (1)atto ion

where τ is the delay between the IR and the APT, Δϕatto is thephase difference between consecutive harmonics, and Δϕion is

Figure 1. (a) Scheme of the experimental setup used for the interferometric measurement. We also indicate the elements needed for theinterferometer stabilization, which is based upon the position of interference fringes observed on the camera. The cross indicates thepolarization axis of the IR and XUV fields. (b) Typical photoelectron spectrogram in Ar.

2

J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 245602 D Guénot et al

the difference in phase of the two-photon ionization matrixelements involved in the creation of a SB. As explained insection 3, Δϕion is related to the photoemission time delay.Since τ and Δϕatto are generally unknown, we cannot extractΔϕion directly from the measured oscillation. Previously, wehave been able to extract the difference in Δϕion corre-sponding to ionization from different subshells [9, 25] byusing different energy regions of the same photo-electronspectrogram. Here, the active stabilization of our spectrometerallows us to record several spectrograms, in different gases,while keeping control of the optical delay of the inter-ferometer. Since τ and Δϕatto are kept constant, we are able todetermine the change in Δϕion between two gases (labeled 1and 2) at the same excitation energy and thus the difference intwo photon emission time delay (τ(2) ):

τ τΔϕ Δϕ

ω− =

2. (2)1

(2)2(2) ion,1 ion,2

Figure 2(b)–(d) schematically shows the involved (uncorre-lated) ionization paths from the ground state to the differentfinal states of the ion and an outgoing electron, which arecoupled to the total angular momentum 1Se or

1De. The finalstates are np5 2P, where n = 3 in Ar, and n = 2 in Ne, (b)–(c)or 2s S1 2 in He (d). At large distances from the ion the dif-ferent angular momentum channels for the photoelectron canbe unambiguously defined since the coupling between themgoes to zero when → ∞r . For shorter distances the channelsdo couple, which is properly accounted for with the many-body methods discussed below (see [35], equations (5)–(7)for more details).

Some of the pathways interfere, [ → →p d p and→ →p s p (m = 0)] since two angular momentum paths

reach the same final state. In contrast, final f states can onlybe reached by a single angular momentum path [ → →p d f( = ±m 0, 1)] and consequently no interference is expected.The different final states are added incoherently in an angle-integrated experiment, which is the case in the present work.

The influence of the different channels on the SB oscil-lation will be analyzed in section 3. In all the measurementspresented in this work, (see figure 1(b)), the SB oscillations

exhibit a good contrast, of the order of 90%, allowing us toeasily extract the phase by Fourier transform.

2.3. Interferometer stabilization and measurement procedure

In order to stabilize the Mach-Zehnder interferometer (seefigure 1(a)), we split off a small fraction of the beam in eacharm. Part of the pump beam is picked off after the generationcell. The fraction of the probe beam used for stabilization isthat transmitted through the recombination mirror. The twobeams are overlapped on a camera chip at a small angleleading to an interference fringe pattern. The fringe position,which depends on the relative delay between the pump andprobe pulses, is used as the error signal for the stabilizationand is fed back to the controller of the piezoelectric delaystage in the probe arm. The active stabilization allows us tokeep the delay fluctuations to below 50 as over 5 min. A slow,thermally-induced drift of approximately 1 fs per hour,remains. It can be minimized by keeping both pump andprobe pulse energies constant during long periods of time, andby stabilizing the pressure in the generating gas cell [36].Stabilization schemes, which do not show this thermal drifthave been used by other groups [37–39]. These techniques,however, require the use of an additional, co-propagatinglaser beam.

Figure 3 shows consecutive measurements of the phaseof SB22 in argon and in neon. Rather than plotting phasesalong the vertical axis, we show delays, i.e. phase (or phasedifference) divided by the period of the process, π ω = 1.3 fs.Each scan, resulting in a point on the graph, takes about fiveminutes, so that the total measurement time is about 1 h. Thestatistical error on the delay determination for each scan isestimated to be approximately 25 as. The drift can be mea-sured by repeating the scans, alternating between the twogases. It is found to be approximately linear with time, whichallows us to determine the difference in delay by taking thedifference between the experimental measurement in one gas(Ar) with the line fitted to the measurements in the other gasat the same time, as shown in the inset. The error bars wereestimated by taking the standard deviation of the statisticalseries (with a 68% confidence interval). They include Poisson

Figure 2. Energy diagrams. (a) Interferometric measurement principle. The blue arrows correspond to harmonic photons and the red to IRphotons. (b)–(d) Angular channels involved in two photon ionization of noble gas atoms from different subshells: (b) np6 (m = 0) (c) np6

( = ±m 1) (d) ns2. We also indicate the corresponding angular coefficients [32].

3

J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 245602 D Guénot et al

statistics and fluctuations due to the drift. Data taken when thethermal drift is perfectly linear with time, gives better statis-tics, like in the case of He/Ne.

2.4. Experimental results

The main result of this article is shown in figure 4. The delaydifferences (Ar, He), (Ar, Ne) and (Ne, He) are plotted as afunction of SB order. Three different SBs are used: 20, 22 and24 corresponding to excitation energies of 31–37.2 eV. Thedelay between argon and the other two gases is of the order of

70 as while the delay between neon and helium is measured tobe around 15 as. A test of the accuracy of the measurement isto compare τ τ−Ne

(2)He(2) to the delay difference that can be

reconstructed from τ τ−Ar(2)

Ne(2) and τ τ−Ar

(2)He(2) . The dotted

green line in figure 4 shows this comparison, which confirmsthe consistency of our measurements. Another test is to per-form the same measurement at different days (i.e. with dif-ferent drift), the results are similar within the error bars. In thefollowing, we present the interpretation of our experimentalresults, based upon lowest-order perturbation theory for theinteraction between atoms and the XUV and IRfields [26, 27].

3. Theory

3.1. Background

In the single active electron approximation, the matrix ele-ment describing ionization by absorption of two photons is

∫∑ ν νϵ Ω ϵ ε

= −+ − +ω Ω

ε ν ν→ +M k E E

k d d i( ) i lim

i. (3)a

i

(2, )

0

Atomic units are used throughout. We use, for simplicity,notations corresponding to a single active electron approx-imation, though our approach, as explained in more detailsbelow, includes channel coupling and correlation effects [35].We choose the quantization axis to be the (common) polar-ization vector of the two fields. The complex amplitudes ofthe laser and the harmonic fields are denoted by ωE and ΩE ,with photon energies ω and Ω, respectively (in the case oftwo-photon ionization, we assume absorption of one IRphoton). The dipole operator is represented by d ( = z for laserpolarization along the z-axis). The initial state is denotedi l m( , , )i i and the final state k L M( , , ). The energies of theinitial and intermediate states are ϵi and ϵν, respectively. Thesum in the expression for M (2) is performed over all possibleintermediate states ν λ μ( , , ), both discrete and continuum. Werefer the reader to [26] for detailed manipulations of the two-photon matrix element.

The next step consists in separating the angular and radialparts of the wavefunctions. We obtain

λμ

λμ

= −

× − −

ω Ωλμ

η

λ

= ±=

⎜ ⎟⎛⎝

⎞⎠

⎛⎝⎜

⎞⎠⎟

( )M k E E Y k

LM

l

mT k

( ) i e i ˆ

10

10

( ), (4)

LMa

limi

k LLM

i

iL

a

(2, )

1

i ( )

(2, )

L

where η k( )L is the scattering phase and YLM a spherical har-

monic. The reduced matrix element, λT k( )L(2) can be written as

λ λ λ ρ=λ κλ⎛⎝⎜

⎞⎠⎟( )T k L l L l

R r( ) ˆ ˆ ˆ 10 0 0

10 0 0

, (5)La

ii

kL(2, ) 2

with = +L Lˆ 2 1 and where ρ ⟩κλ| is the so-called radial

Figure 3. Drift of SB 22 for consecutive scans, performedalternatively in argon (blue crosses) and neon (red crosses). Thearrow shows the average phase difference between argon and neon.The lines (argon blue solid, neon red dashed) are linear fits to thedata, and the vertical lengths of the crosses represents the statisticalerror. The inset shows how the difference in delay is extracted forevery set of scans argon–neon–argon.

Figure 4.Delay differences, τ τ−Ar(2)

He(2) (blue, solid), τ τ−Ar

(2)Ne(2) (red,

dashed), τ τ−Ne(2)

He(2) (green, dashed–dotted) as a function of photon

energy. The open symbols and green dotted curve are reconstructeddelay differences τ τ−Ne

(2)He(2) obtained by taking the difference

between the blue and red delays. The measurements are plotted withartificially shifted energies for the sake of visibility.

4

J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 245602 D Guénot et al

perturbed wavefunction, defined as

∫∑ρϵ Ω ϵ ε

=+ − +κλ

ε ν

νλ νλ

ν→ +

R R r Rlim

i. (6)

n l

i0

i i

The momentum κ is the on-shell momentum, which corre-sponds to a pole in the integrand in equation (6) and λ is theangular momentum of the perturbed wave function. Toevaluate the phase of this quantity, as explained in more detailin [26], we approximate RkL(r) and ρκλ by their asymptoticvalues and evaluate analytically the integral ρ⟨ ⟩κλR r| |kL inequation (5). We obtain

λ π η κ η ϕ κ≈ − + − +λ λ⎡⎣ ⎤⎦T k L k karg ( ) ( )

2( ) ( ) ( , ),L

aL

(2, )cc

where η κλ ( ) and η k( )L are the scattering phases of the per-turbed wavefunction and final state respectively (see figure 5),and where ϕ κk( , )cc is an additional phase associated with thecontinuum–continuum radiative transition, resulting from theinteraction with the IR field in the presence of the Coulombpotential. For a given angular channel λl Li , and omitting the

contributions from the fundamental and harmonic fields[9, 26] for simplicity, we have

λπ η κ ϕ κ≈ − + +λ⎡⎣ ⎤⎦M karg

2( ) ( , ). (7)LM

a(2, )cc

The matrix element in the case of the emission of the IRphoton M k( )e(2, ) has a similar expression. The perturbedwave function now lies above the final state. The difference ofphase which is measured in the experiment does not dependon η k( )L which cancels out, and is given by

Δϕ η κ η κ ϕ κ ϕ κ= − + −λ λ> < > <( ) ( )k k( ) ( ) , , (8)ion cc cc

where κ>, κ< are the momenta corresponding to the perturbedwave function in case of IR emission and absorption (seefigure 2(a)). This equation shows that when the intermediatestep, corresponding to one XUV-photon absorption, onlyinvolves a single angular momentum channel (figures 2(c),(d)), the phase difference does not depend on the final state.(It depends however, on the energy of the final state throughthe cc contribution.) Dividing equation (8) by ω2 , we have

Figure 5. (a) Cross section (blue) and phase (red) of different ionization channels in Ar, Ne, He; (a), (b) 3p ϵ→ d (cross section: solid line,phase: dashed–dotted, line); 3p ϵ→ s (cross section: dashed line, phase: dotted line); (a) RPAE; (b) MCHF; (c) Ne, RPAE, 2p ϵ→ d (crosssection: solid line, phase: dashed–dotted line); 2p ϵ→ s (cross section: dashed line, phase: dotted line); (d) He, RPAE: 1s ϵ→ p (cross section:solid line, phase: dashed line).

5

J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 245602 D Guénot et al

τΔϕ

ωτ τ= = +k k k( )

2( ) ( ), (9)(2) ion (1)

cc

where

τη κ η κ

ωΔη

ω=

−=λ λ λ> <

k( )( ) ( )

2 2, (10)(1)

is a finite difference approximation to the Wigner time delayη ϵλd d [13] and thus reflects the properties of the electronicwave packet ionized by one-photon absorption into theangular channel λ. τ(2) also includes a contribution from theIR field which is independent of the angular momentum

τϕ κ ϕ κ

ω=

−> <( ) ( )k

k k( )

, ,

2. (11)cc

cc cc

3.2. One-photon ionization phases and cross sections

The phase and amplitude of the photoionization transitionmatrix element can be determined using the RPAE (for detailsof the theory, see [30]). The dipole matrices in equation (3)are replaced by screened dipole interactions which include thedynamical correlations among all electrons in the two outer-most occupied atomic orbitals. The atomic orbitals are cal-culated using the Hartree–Fock method. Figure 5 shows thecalculated phases and cross sections for all the channelsinvolved in the photoionization of argon (a), neon (c) andhelium (d) between 20 and 40 eV (25 and 40 for Ne and He)[30]. An important feature is that, the channels going from theground state to excited d states in Ne and Ar are the dominantchannels in the region investigated experimentally as pre-dicted by the Fano propensity rule [40]. This is not the case inargon at high energy (48 eV) due to the presence of a Cooperminimum [41].

In figure 5(b), we present a different calculation in Arusing a MCHF approach [28] convoluted in energy by thewidth of the XUV field (0.1 eV). The average behavior ofboth phases and amplitudes is quite similar to those obtainedwith RPAE. MCHF also accounts for the influence of doublyexcited states, which leads to multiple resonance structures,superposed on a smooth background.

3.3. Influence of multiple channels

The photoionization phases and cross sections presented infigure 5 strongly depend on the final angular momentum. Inthis subsection, we examine how different angular momentaboth in the intermediate and final steps of the two-photonionization process (figure 2) contribute to the SB oscillationand how it is possible to deduce experimental photoionizationtime delays from the measurements. Including both quantumpaths (with absorption and emission of the IR photon), thetotal signal is equal to

∑= +I M M . (12)LM

LMa

LMe2, 2, 2

We consider first the case of two channels with the sameintermediate step like in He (see figure 2(d)). The total signal

is:

= + + +I M M M M . (13)a e a e202,

202, 2

002,

002, 2

The interferometric traces corresponding to different finalstates oscillate in phase, since the phase of these oscillationsdoes not depend on the final state. They can be added inco-herently without any blurring of the SB oscillations and themeasured phase difference corrected by the influence of theIR field can be interpreted unambiguously as the →s pionization delay.

In the case of Ar and Ne, however, the situation is morecomplex. The total signal is

= + + +

+ + + +

± ± ± ±I M M M M

M M M M

2 2

. (14)

a e a e

a e a e

3 12,

3 12, 2

1 12,

1 12, 2

302,

302, 2

102,

102, 2

The first three terms oscillate in phase, because the inter-mediate state is uniquely a d-wave. The fourth term containscontributions from both intermediate s and d-waves andrequires extra attention. The channels → →p s p and

→ →p d p (m = 0) interfere coherently. The measurement ofthe ionization delay corresponding to the channel →p drequires that the →p s channel is weak, or alternatively thatits phase variation remains close to that of the →p d channel.

To estimate the influence of multiple channels, we cal-culate the phase corresponding to the interfering →p d and

→p s channels in Ar. The coherent addition of two inter-fering channels with amplitudes Apsp and Apdp can be writtenas

= +η η ηA A Ae e e . (15)iint

ipsp

ipdpint 0 2

Assuming that IR absorption (or emission) contributes withthe same amplitude, and accounting for the angular coeffi-cients for the different steps, we have

σ σ≃ −A A 8 25pdp psp 2 0 where σ0,2 are the differentialphoto-absorption cross sections with final angular momentum0, 2 respectively. The resulting phase is given by the fol-lowing equation

η ηη η

η η≃ +

− − σσ

⎢⎢⎢

⎥⎥⎥arctan

sin ( )

cos ( ). (16)int 2

0 2

0 28

252

0

Figure 6(a) compares the derivatives of η2, η0 and ηint. Tomimic the experiment, we represent finite approximations tothese derivatives, τ Δη ω= 2i i

(1) . The ‘interference’ delay,

τint(1) , deviates slightly from that of the →p d channel espe-

cially toward high photon energy, due to the increasing →p scross section.

We also examine the influence of the incoherent channelson our interferometric measurement. The total interferometric

6

J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 245602 D Guénot et al

signal (equation 14) can be expressed as

ω τ τ τ

ω τ τ τ

ω τ τ τ

= + − −

+ − −

= + − −

⎡⎣ ⎤⎦⎡⎣ ⎤⎦

⎡⎣ ⎤⎦

( )( )

( )

I I I

I

I I

cos 2

cos 2

cos 2 , (17)

0 2 2(1)

cc

int int(1)

cc

0 tot tot(1)

cc

where I0, I2, Iint and Itot are intensities and where the totalsignal oscillates with a phase equal to ω τ τ+2 ( )tot

(1)cc , where

τ τ

ω

ω τ τ

ω τ τ

=

−−

+ −

⎡⎣ ⎤⎦⎡⎣ ⎤⎦

( )( )

I

I I

1

2arctan

sin 2

cos 2. (18)

tot(1)

2(1)

int 2(1)

int(1)

2 int 2(1)

int(1)

The results presented in figure 6(a) show that τtot(1) is very

close to τ2(1) so that, in these conditions, the →p s channel

has little influence on the measured delay. Vice versa, thedelay that we measure in our experiment (corrected for theinfluence of the IR field) can be interpreted as the group delay

of the →p d ionizing wave packet, with a small offset of theorder of 10 as [26]. This conclusion might change in othersituations where two interfering channels have comparableamplitudes.

3.4. Influence of the atomic structure

In figure 6(b), we present a different calculation in Ar using aMCHF approach [28], including, as above, all of the chan-nels. This approach accounts for the influence of doublyexcited states as well as 3s-excitation channels, which lead tomultiple resonance structures. The MCHF photoionizationdelay looks quite noisy. This is due to the influence of theresonances on the phase (see figure 5(b)) in combination withthe finite approximation of the phase derivative. The energyof the structures can be calculated only within an accuracy of≃0.5 eV. This affects the determination of the time delays,leading to a possible error when comparing to the experiment,of the order of the observed structures (at most ±20 as).

Figure 6. Theoretical one-photon delays. (a) Investigation of the influence of multiple channels in Ar using RPAE; τ2(1) (red solid), τ0

(1)

(yellow dashed), τint(1) (green dotted) and τtot

(1) (blue dashed–dotted) (see text); (b) one-photon delay in Ar with MCHF (magenta solid), RPAE(blue dashed–dotted) and for a pure Coulomb potential (red dashed); (c) one-photon delay in Ne, RPAE (blue solid) and Coulomb (reddashed); (d) one-photon delay in He, RPAE (blue solid) and Coulomb (red dashed).

7

J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 245602 D Guénot et al

Finally, we present in figure 6 one-photon delays in Ne(c) and He (d) calculated using RPAE. These delays are quitesimilar, since in this region, correlation effects are weak. Thecalculated delay mostly reflects the variation of the Coulombphase in this energy region [26], as also shown by comparingto the one-photon delays for a Coulomb potential (red). Incontrast, the difference obtained in Ar emphasizes the sensi-tivity of the delay to the atomic structure in general.

3.5. Two-photon ionization delays

The two-photon ionization delay can be approximated byadding the one photon delay and the continuum–continuumdelay according to equation (8). Figure 7(a) shows the τcc

delay for the three gases investigated (Ar, Ne, He) as afunction of the harmonic order. The difference between atomsis simply due to the difference in kinetic energy of the out-going electron.

The results for the three gases (RPAE + CC) are shownin figure 7(b) and table 1. We also performed directly a two-photon calculation, within the framework of the RPAE [27].The results are indicated by the symbols for the three raregases. The agreement between the two calculations is excel-lent in He and Ar, which shows the validity of the approx-imation consisting in calculating the phase induced by theabsorption or emission of an IR photon by using theasymptotic form of the wave functions (‘cc-approximation’).Differences of 10–15 as are observed in Ne, which possiblyindicates that the cc-approximation, especially for the path

→ →p s p, is not as good in this case.

4. Comparison with experiment

In figure 8, we compare theoretical (RPAE + CC, two-photonRPAE) and experimental results for the delay differences. Wealso present calculations where τAr is calculated with theMCHF approach (+CC) and τNe or τHe with RPAE + CC. Theagreement is excellent for the differences in delay between Neand He, which can be accurately calculated, since the influ-ence of correlation is small. There is a larger differencebetween theory and experiment in the case of Ar.

The imprecision in determination of the energy of eachresonance in the MCHF method implies that a better repre-sentation of the MCHF results would be a smooth curveaveraging over the resonances combined with a ‘theoretical’error bar of ±20 as. This average curve (which the reader canvisualize by eye) increases from 31 to 35.5 eV and thendecreases until 37 eV. It fits the experimental observationbetter for both SBs 22 and 24 than RPAE (figures 8(a) and(b)). This might indicate that the theoretical description of theAr delays requires the inclusion of double excitations in thisenergy range.

We summarize our experimental and theoretical results intable 2.

Figure 7. (a) τcc delay in Ar (solid line), Ne (dashed line), He (dotted line); (b) two-photon ionization delays: τAr (RPAE + CC, blue, solidline; two-photon RPAE black squares; MCHF, magenta); τNe (RPAE + CC, blue, dashed line; two-photon RPAE, empty circles) and τHe

(RPAE + CC, blue, dotted line; two-photon RPAE, full circles).

Table 1. Continuum–continuum and two photon delays. All delaysgiven in as.

Atom Sideband 20 22 24

Photon energy (eV) 31 34.1 37.2

Ar τcc −47 −38 −31.6RPAE + CC 8.9 12.6 −0.2RPAE 2 photons 7.9 10.3 7.3MCHF + CC 8.9 7.5 11

Ne τcc −82 −59 −45.5RPAE + CC −37.9 −29.9 −23.8RPAE 2 photons −51.7 −40.7 −33.2

He τcc −120 −77.1 −56.3RPAE + CC −62.5 −47.4 −38RPAE 2 photons −64 −49.1 −37.8

8

J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 245602 D Guénot et al

5. Conclusion and outlook

In conclusion, the new experimental scheme presented hereallows an accurate measurement of the photo-emission timedelay between different atomic systems, providing more datato test the different theoretical approaches. The small dis-crepancy between experiment and theoretical calculationsindicates that both more accurate measurements and calcu-lations should be performed in particular in the case of argon.This technique can potentially allow the measurement of thedelays in molecules, relative to those in a reference atom. Ourresults highlight the importance of including doubly excitedstates as well as all angular channels in the theoretical cal-culations. Using an angularly-resolved detection technique,such as a velocity map imaging spectrometer, would helpdisentangle the different channels.

Figure 8. Comparison between experiment and theory. (a) τ τ−Ar Ne (b) τ τ−Ar He and (c) τ τ−Ne He. The red crosses correspond to theexperimental measurement, the blue curves to one photon RPAE + CC, the black symbols to two-photon RPAE. The magenta curve has beenobtained by using MCHF + CC in Ar.

Table 2. Time delay measurements and calculations. All delaysgiven in as.

Sideband 20 22 24

Photon energy (eV) 31 34.1 37.2τ τ−(Ar) (Ne) 68 ± 15 70 ± 12 52 ± 25RPAE + CC 46 38 24RPAE 2 photons 59.6 51 40MCHF(Ar)-RPAE(Ne) + CC 46 42 34τ τ−(Ar) (He) 82 ± 15 83 ± 22 71 ± 21RPAE + CC 71 55 38RPAE 2 photons 71.9 59.4 45.1MCHF(Ar)- RPAE(He) + CC 71 60 49τ τ−(Ne) (He) 23 ± 4 12 ± 4 10 ± 8RPAE + CC 23 15 14RPAE 2 photons 12.3 8.4 4.6

9

J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 245602 D Guénot et al

Acknowledgments

This research was supported by the European ResearchCouncil (ALMA, PALP), the ATTOFEL network, the Knutand Alice Wallenberg foundation, the Swedish Foundationfor Strategic Research, the Swedish Research Council and theAustralian Research Council. KV acknowledges support fromthe Bolyai Grant of the Hungarian Academy of Sciences EBis financed by the Hungarian Scientific Research Fund(OTKA project NN 107235).

References

[1] Schmidt V 1992 Rep. Prog. Phys. 55 1483[2] Reid K L 2003 Annu. Rev. Phys. Chem. 54 397[3] Geßner O et al 2002 Phys. Rev. Lett. 88 193002[4] Strand M P, Hansen J, Chien R-L and Berry R S 1978 Chem.

Phys. Lett. 59 205[5] Corkum P and Krausz F 2007 Nat. Phys. 3 381[6] Krausz F and Ivanov M 2009 Rev. Mod. Phys. 81 163[7] Cavalieri A L et al 2007 Nature 449 1029[8] Schultze M et al 2010 Science 328 1658[9] Klünder K et al 2011 Phys. Rev. Lett. 106 143002[10] Ivanov M and Smirnova O 2011 Phys. Rev. Lett. 107 213605[11] Dahlström J, LHuillier A and Maquet A 2012 J. Phys. B: At.

Mol. Opt. Phys. 45 183001[12] de Carvalho C A and Nussenzveig H M 2002 Phys. Rep.

364 83[13] Wigner E P 1955 Phys. Rev. 98 145[14] Yakovlev V S, Gagnon J, Karpowicz N and Krausz F 2010

Phys. Rev. Lett. 105 073001

[15] Baggesen J C and Madsen L B 2010 Phys. Rev. Lett. 104043602

[16] Zhang C-H and Thumm U 2010 Phys. Rev. A 82 043405[17] Kheifets A S and Ivanov I A 2010 Phys. Rev. Lett. 105 233002[18] Nagele S et al 2011 J. Phys. B: At. Mol. Opt. Phys. 44 081001[19] Moore L et al 2011 Phys. Rev. A 84 061404[20] Nicolaides C A, Mercouris T and Komninos Y 2002 J. Phys.

B: At. Mol. Opt. Phys. 35 L271[21] Mercouris T, Komninos Y and Nicolaides C A 2007 Phys. Rev.

A 76 033417[22] Haessler S et al 2009 Phys. Rev. A 80 011404[23] Caillat J et al 2011 Phys. Rev. Lett. 106 093002[24] Swoboda M et al 2010 Phys. Rev. Lett. 104 103003[25] Guénot D et al 2012 Phys. Rev. A 85 053424[26] Dahlström J M et al 2013 Chem. Phys. 414 53[27] Dahlström J, Carette T and Lindroth E 2012 Phys. Rev. A 86

061402[28] Carette T, Dahlström J, Argenti L and Lindroth E 2013 Phys.

Rev. A 87 023420[29] Dixit G, Chakraborty H S and Madjet M E-A 2013 Phys. Rev.

Lett. 111 203003[30] Kheifets A 2013 Phys. Rev. A 87 063404[31] Månsson E P et al 2014 Nat. Phys. 10 207–11[32] Toma E S and Muller H G 2002 J. Phys. B: At. Mol. Opt. Phys.

35 3435[33] Paul P M et al 2001 Science 292 1689[34] Mairesse Y et al 2003 Science 302 1540[35] Dahlström J M and Lindroth E 2014 J. Phys. B: At. Mol. Opt.

Phys. 47 124012[36] Kroon D et al 2014 Opt. lett. 39 2218[37] Böttcher F et al 2008 Appl. Phys. B 91 287[38] Chini M et al 2009 Opt. express 17 21459[39] Fieß M et al 2010 Rev. Sci. Instrum. 81 093103[40] Fano U 1985 Phys. Rev. A 32 617[41] Kennedy D J and Manson S T 1972 Phys. Rev. A 5 227

10

J. Phys. B: At. Mol. Opt. Phys. 47 (2014) 245602 D Guénot et al