Author(s): Niikura, Hiromichi; Wörner, Hans Jakob; Villeneuve, David M.; Corkum, Paul B.

Similar documents
SUPPLEMENTARY INFORMATION

Revival Structures of Linear Molecules in a Field-Free Alignment Condition as Probed by High-Order Harmonic Generation

Looking into the ultrafast dynamics of electrons

High-contrast pump-probe spectroscopy with high-order harmonics

HHG Sub-cycle dynamics

Supporting Online Material for

Photoelectron Spectroscopy using High Order Harmonic Generation

SUPPLEMENTARY INFORMATION

attosecond laser pulse

ATTOSECOND AND ANGSTROM SCIENCE

Ideal laser waveform construction for the generation of super-bright attosecond pulses

C. D. Lin Kansas State U.

Attosecond laser systems and applications

WP-3: HHG and ultrafast electron imaging

Wavelength scaling of high-order harmonic yield from an optically prepared excited state atom

Probing collective multi-electron dynamics in xenon with high-harmonic spectroscopy

Strongly Dispersive Transient Bragg Grating for High Harmonics

High-energy collision processes involving intense laser fields

Two-pulse alignment of molecules

Nonlinear Optics (WiSe 2015/16) Lecture 12: January 15, 2016

Supplementary Material for In situ frequency gating and beam splitting of vacuum- and extreme-ultraviolet pulses

Analysis of recombination in high-order harmonic generation in molecules

arxiv: v2 [physics.atom-ph] 9 Dec 2014

Attosecond-correlated dynamics of two electrons in argon

Multislit interference patterns in high-order harmonic generation in C 60

Recollision physics. feature. Paul B. Corkum

Models for Time-Dependent Phenomena

1 Mathematical description of ultrashort laser pulses

stabilized 10-fs lasers and their application to laser-based electron acceleration

SUPPLEMENTARY INFORMATION

Overview: Attosecond optical technology based on recollision and gating

Coherent Electron Scattering Captured by an Attosecond Quantum Stroboscope

arxiv: v1 [physics.atom-ph] 18 May 2012

High Harmonic Generation of Coherent EUV/SXR Radiation. David Attwood University of California, Berkeley

Generation and Applications of High Harmonics

Quantum interference and multielectron effects in highharmonic spectra of polar molecules

Lecture on: Multiphoton Physics. Carsten Müller

High-Harmonic Generation II

XUV attosecond pulses

Tunneling time, what is its meaning?

High-order harmonics with fully tunable polarization by attosecond synchronization of electron recollisions

Harmonic Generation for Photoionization Experiments Christian J. Kornelis Physics REU Kansas State University

Models for Time-Dependent Phenomena

Molecular alignment, wavepacket interference and Isotope separation

Generation of a single attosecond pulse from an overdense plasma surface driven by a laser pulse with time-dependent polarization

Interferometric time delay correction for Fourier transform spectroscopy in the extreme ultraviolet

An extreme ultraviolet interferometer using high order harmonic generation

Ultrafast Laser Physics!

Electron spins in nonmagnetic semiconductors

Multiphoton transitions for delay-zero calibration in attosecond spectroscopy arxiv: v1 [physics.atom-ph] 12 Jun 2014

arxiv: v1 [quant-ph] 24 Sep 2009

4. High-harmonic generation

AMO physics with LCLS

Tunneling Time in Ultrafast Science is Real and Probabilistic

High Harmonic Phase in Molecular Nitrogen. Abstract

Coherent interaction of femtosecond extreme-uv light with He atoms

Schemes to generate entangled photon pairs via spontaneous parametric down conversion

Effects of driving laser jitter on the attosecond streaking measurement

(002)(110) (004)(220) (222) (112) (211) (202) (200) * * 2θ (degree)

Ultrafast XUV Sources and Applications

Chemistry 311: Instrumentation Analysis Topic 2: Atomic Spectroscopy. Chemistry 311: Instrumentation Analysis Topic 2: Atomic Spectroscopy

High order harmonic generation and applications

Efficient isolated attosecond pulse generation from a multi-cycle two-color laser field

Theory of high-order harmonic generation from molecules by intense laser pulses

CHINESE JOURNAL OF PHYSICS VOL. 52, NO. 1-II February Intense Few-Cycle Infrared Laser Pulses at the Advanced Laser Light Source

Attosecond-streaking time delays: Finite-range property and comparison of classical and quantum approaches

Chapter 13. High Harmonic Generation

Dark pulses for resonant two-photon transitions

Models for Time-Dependent Phenomena. I. Laser-matter interaction: atoms II. Laser-matter interaction: molecules III. Model systems and TDDFT

Simple strategy for enhancing terahertz emission from coherent longitudinal optical phonons using undoped GaAs/n-type GaAs epitaxial layer structures

Attosecond science PROGRESS ARTICLE P. B. CORKUM 1 AND FERENC KRAUSZ 2,3 1

SUPPLEMENTARY INFORMATION

Phys 2310 Mon. Dec. 11, 2014 Today s Topics. Begin Chapter 9: Interference Reading for Next Time

arxiv: v1 [physics.optics] 15 Dec 2011

Supplementary Materials

Introduction to intense laser-matter interaction

High-order harmonic generation in fullerenes with icosahedral symmetry

LETTER. Linking high harmonics from gases and solids

SUPPLEMENTARY INFORMATION

Ultrashort Phase Locked Laser Pulses for Asymmetric Electric Field Studies of Molecular Dynamics

Time delay in atomic photoionization with circularly polarized light

Atoms, Molecules and Solids. From Last Time Superposition of quantum states Philosophy of quantum mechanics Interpretation of the wave function:

Control and Characterization of Intramolecular Dynamics with Chirped Femtosecond Three-Pulse Four-Wave Mixing

Dynamics of Laser Excitation, Ionization and Harmonic Conversion in Inert Gas Atoms

Extending the strong-field approximation of high-order harmonic generation to polar molecules: gating mechanisms and extension of the harmonic cutoff

Generation of ultrashort XUV femtosecond to attosecond pulses Katalin Varjú ELI-ALPS. 2nd MOLIM Training School 6 10 March, 2017 Paris-Saclay

Instantaneous Multiphoton Ionization Rate and Initial Distribution of Electron Momenta. Abstract

Role of nuclear spin in photoionization: Hyperfine-resolved photoionization of Xe and multichannel quantum defect theory analysis

Abstract... I. Acknowledgements... III. Table of Content... V. List of Tables... VIII. List of Figures... IX

PHYS 4 CONCEPT PACKET Complete

Generating intense attosecond x-ray pulses using ultraviolet-laser-induced microbunching in electron beams. Abstract

An Investigation of Benzene Using Ultrafast Laser Spectroscopy. Ryan Barnett. The Ohio State University

Time-frequency characterization of femtosecond extreme ultraviolet pulses.

Supplemental material for Bound electron nonlinearity beyond the ionization threshold

Laserphysik. Prof. Yong Lei & Dr. Yang Xu. Fachgebiet Angewandte Nanophysik, Institut für Physik

Construction of an extreme ultraviolet polarimeter based on highorder harmonic generation

37. 3rd order nonlinearities

PIs: Louis DiMauro & Pierre Agostini

37. 3rd order nonlinearities

Extension of Harmonic Cutoff and Generation of Isolated Sub-30 as Pulse in a Two-Color Chirped Laser Field

Transcription:

Research Collection Journal Article Probing the Spatial Structure of a Molecular Attosecond Electron Wave Packet Using Shaped Recollision Trajectories Author(s): Niikura, Hiromichi; Wörner, Hans Jakob; Villeneuve, David M.; Corkum, Paul B. Publication Date: 2011-08-26 Permanent Link: https://doi.org/10.3929/ethz-a-010783488 Originally published in: Physical Review Letters 107(9), http://doi.org/10.1103/physrevlett.107.093004 Rights / License: In Copyright - Non-Commercial Use Permitted This page was generated automatically upon download from the ETH Zurich Research Collection. For more information please consult the Terms of use. ETH Library

Probing time-dependent structure of the attosecond electron wave packet using shaped re-collision trajectories Hiromichi Niikura 1,2,3, Hans Jakob Wöner 1,4, D. M. Villeneuve 1 and P. B. Corkum 1,5 1 National Research Council of Canada, 100 Sussex Drive, Ottawa, Ontario, Canada K1A0R6 2 PRESTO, Japan Science and Technology Agency, Sanbancho building, 5-Sanbancho, Chiyodaku, Tokyo 102-0075, Japan 3 Department of Applied Physics, Waseda University, Okubo3-4-1, Shinjyuku, Tokyo 169-8555, Japan 4 Laboratorium für physikalische Chemie, ETH Zürich, Wolfgang-Pauli-Strasse 10, 8093 Zürich, Switzerland 5 Department of Physics, University of Ottawa, 150 Louis Pasteur, Ottawa, Ontario, Canada K1N 6N5 Corresponding author: niikura@waseda.jp Revised on Mar.30, 2011. Using orthogonally polarized, 800 nm and 400 nm laser pulses, we generated high harmonics from C 2 H 6 and measured the spectra as a function of the delay between the two laser pulses. We observed that the intensity of each harmonics modulates with the delay. The modulation period of the 14 th and 16 th harmonics is twice as the period of the modulation of other harmonics. By comparing theoretical calculation, we indentify that the double periodicity is a result of the electron wave packet motion in the valence shell of C 2 H 6 with attosecond time-scale. Our method can be an approach to measure internal electron dynamics and does not require molecular alignment, making it applicable to complex molecules. PACS numbers: 33.80Rv, 42.50Hz 1

Attosecond electron wave packets moving in a valence shell maybe an unexplored but common feature of chemical reactions, photoionization and charge transfer in molecules [1-5]. When we shine intense, infrared laser pulses to a molecule whose electronic state is close to each other, tunnel ionization will occur from the several electronic states [6-8]. In 2009, it was suggested that this process can create an electron wave packet or electron-hole dynamics in the valence shell of a molecule by utilizing the high harmonic generation from aligned CO 2 [6]. However, there are several difficulties to apply this method directly to other molecules. First, although they identified that the intensity minimum observed on the high harmonic spectra is caused as a result of the interference between the two ionization channels, the interference minimum will appear if two harmonics are emitted from independent two species [9]. It doesn t guarantee existing of the internal electron wave packet directly. Second, because one linearly polarized laser pulse is responsible for both tunnel ionization and electron re-collision processes, the spatial structure of the internal and continuum electron wave packets are influenced by the molecular alignment angle with respect to the filed direction. That makes observation to be complex and requires extensive theoretical calculation for analysis. Third, it is difficult to trace dynamical change in the electron wave packet as a function of time. Using C 2 H 6, we demonstrate that electron wave packet motion in a valence shell is observed with attosecond time-resolution as it changes. To measure the dynamics, we employ a two-color laser field approach with unaligned molecules [10-11]. This approach allows us to indentify the symmetry of an electron wavefunction (or orbital) in a molecule which is responsible for the high harmonic generation process. It has advantageous over the alignment method employed in Ref. [6] in terms of measuring electron wave packet motion. First, tunnelling and valence electron wave packets produced at a time of ionization should be essentially unchanged for the parameter of measurement, such as the delay between the two laser pulses, except that the tunnelling wave packet will initially move along the direction given by the field. This can be possible because this approach permits disentangling the ionization process from the probing process with re-collision by combing two-color laser fields. That minimizes complexity of the analysis. Second, we can determine key times when the structure as seen by the continuum electron changes drastically since dynamical changes in the wave packet can be explored directly from the measurement by observing the changes in the polarization direction of harmonics as a function of the re-collision angle. 2

Fig. 1a shows a scheme of the measurement. We combine an 800 nm laser pulse with its second harmonic spatially and temporally. The polarization direction of the 800 nm is set to orthogonal to that of the second harmonic. At the peak of the combined laser fields, tunnel ionization of molecules can form the correlated electron wave packet pair. One moves in the valence and the other moves in the ionization continuum. With unaligned molecules, if tunnel ionization probability is directional to a specific range of angles with respect to the molecular principal axis, the ionized molecular ensemble is distributed in space directionally to the specific angle range. Within one optical period after the ionization, the continuum electron re-collides with the other electron wave packet, leading high harmonic emission. When we change the relative delay between the two laser pulses, the amplitude and the direction of sum of the laser fields change instantaneously. It controls continuum electron trajectory and re-collision angle, θ c, with respect to the molecular axis. As θ c changes, the direction of the induced dipole moment, i.e., the polarization angle of harmonics (φ HHG ) varies. The variation of φ HHG with θ c depends on characteristically the alignment and the spatial symmetry of the valence state electron wave packet [10]. Therefore, if dynamical motion occurs in the valence state, then the dependence of φ HHG on θ c follows its motion. Because the harmonic number corresponds to the re-collision time, t c, the dynamical motion is traced along the harmonic number when we choose an appropriate phase matching condition to yield only a short trajectory [12]. We obtain the relative value of φ HHG from the measured high harmonic generation spectra by comparing the intensity ratio between the adjacent odd and even harmonics [10-11]. This is based on the fact that even (odd) harmonics are polarized parallel to the 400 nm (800 nm) polarization axis, respectively. We utilize H 2 as a reference molecule to obtain θ c as a function of the harmonic number. With unaligned molecules, φ HHG of H 2 is approximately equal to θ c because the angular dependence of the tunnel ionization probability is approximately spherical [13]. Therefore, comparison of φ HHG of H 2 with that of C 2 H 6 tells us how the polarization angle changes as a function of the re-collision angle. In Fig. 1b, we present the structure of the several electronic states of C 2 H 6 obtained by ab-initio calculation and the vertical ionization energy [14]. We also plot the calculated angular dependence of the tunnel ionization probability for the E g and A g state at the total laser intensity of 2.0 x 10 14 W/cm 2. The dotted line shows the incoherent sum of both tunnel ionization probabilities. It has maximum values at around angles of ~ 45 ± 3

90 and 225 degrees. The ionization occurs preferentially from these angles. The mechanism to prepare internal electron wave packet replies on ref [6]. The electron wavefunction after the tunnel ionization is given by Ψ ( t) Ψ Ψ, b c Ψ = a( 1) Ψ ( E ) + a(2) Ψ ( A )exp( iϕ( t)) (1) b b g b g where a(1) and a(2) is the coefficient and ϕ(t) is the phase difference between the two states. Ψb(Eg) and Ψb(Ag) are the electron wavefunctions of the singly-charged ion or Dyson orbital of each ionization channel, where one electron is removed from the most and the second most loosely bound state of C 2 H 6, respectively. The coherent superposition between the two wavefunctions generates dynamical motion of the electron vacancy, i.e., hole dynamics or electron wave packet motion in the molecule. Phase matching requires that the continuum electron is returned to its original ground state, forcing the electron to recombine to a new location and a changed hole distribution. We generate an intense, 35 fs pulse duration, 800 nm laser pulse by a Ti:Sapphire laser system (KMlabs). We double the frequency of the 800 nm pulse by a 300 micrometer thickness β-bab 2 O 4 crystal. We pass both laser beams coaxially through a 0.65 mm calcite plate. By rotating the calcite plate to accumulate an extra optical path, we adjust the phase delay between the two laser pulses [10]. We focus the laser pulses to downstream of a pulsed gas jet with a 50 cm focal length spherical mirror. The generated harmonics are dispersed by a flat-field grating (Hitachi, 001-226). The dispersed spectrum is imaged onto a two-dimensional microchannel plate and a phosphor screen. We capture the image by a CCD camera and transfer to a computer. We integrate the signal counts vertical to the dispersed axis. In Fig. 2, we show the measured high harmonic emission spectra of C 2 H 6 vs relative delay between the two laser pulses at the total laser intensity of 2.0 x 10 14 W/cm 2. For reference, we measure the spectra of high harmonics generated from H 2 and N 2 as well using the same laser field (not shown). Except for 14 th and 16 th harmonics, the intensity of each harmonic emission of C 2 H 6 modulates with half of the 400 nm optical period, ~ 0.66 fs. This modulation period is caused by the fact that the same amplitude of the laser field is repeated with the period [10, 11]. For 14 th and 16 th, two intensity peaks are observed at every half of the 400 nm optical period. One peak appears at approximately the same delay as the adjacent odd harmonic has a peak (~0.33 and ~0.99 4

fs), while the other peak shifts by π in the delay (~0.66 and ~1.33 fs). These two peaks converges to one peak in the harmonic range higher than 18 th. In case of H 2 and N 2, all harmonics are modulated with half of the 400 nm optical period. In order to clarify the dynamics, in Fig. 3 a and b we plot φ HHG of H 2 and C 2 H 6 obtained from the measured high harmonic spectra. One φ HHG peak is found in H 2 while two φ HHG peaks are found in C 2 H 6 at every half of the 400 nm period. In the range from 13 th to 15 th at the delay of ~0.33 fs and ~1.00 fs, φ HHG of C 2 H 6 is approximately proportional to φ HHG ( = θ c ) of H 2. This indicates that dipole moment responsible for high harmonic emission is induced approximately parallel to the re-collision angle. It is consistent with σ symmetry character [10]. In the range from the 16 th to 21 st at the delay of ~0, 0.66 and 1.33 fs, φ HHG of C 2 H 6 has a peak up ~45 degrees while φ HHG of H 2 is ~0 degree. This indicates that the dipole moment is induced to the angle different from the re-collision angle. Thus, the spatial structure of Ψ b must be changed during from the re-collision time which yields the low harmonics (~13 th ) to the time which yields the high harmonics (~18 th ). To exhibit the relationship of the re-collision time with the harmonic number, we plot the re-collision time obtained by classical electron trajectory calculation [16] in the right axis of Fig. 3. The range between 13 th and 18 th corresponds to the re-collision time from 0.8 fs to 1.2 fs. We have repeated the same measurement at four different driving laser intensities in the range from 1.1 x 10 14 W/cm 2 to 1.7 x 10 14 W/cm 2. From the observed spectra, we obtain the φ HHG distribution for each laser intensity as shown in the online material [15]. The highest harmonic number where the double intensity peak appears shifts from 16 th to 14 th and the φ HHG distribution pattern shifts to low harmonic number as the laser intensity decreases. Thus, the possibilities of interference specific to the molecular structure and the electron wavefunction can be excluded. This is because in case of the structural interference, the characteristic intensity minimum must be independent of the driving laser intensity [6, 17]. When we convert the harmonic number to the re-collision time at each laser intensity using classical electron trajectory calculation, we found that the φ HHG peak appearing at the delay of 0, ~0.66, and ~1.33 fs stays at the same re-collision time of ~1.2 fs regardless of the laser intensity. Therefore, the experimental result is consistent with the prediction that the tunnel ionization prepares the electron wave packet moving in the molecule [6]. Next, we identify the dynamics of Ψ b by simulating φ HHG as a function of the high 5

harmonic number and the two-color laser delay using a semi-classical, strong-field approximation [16]. First, we calculate electron trajectories and θ c under the two-color laser fields. Then we calculate φ HHG from the high harmonic intensity induced parallel to the 800 nm axis and the 400 nm axis [14]. We use Eq. 1 to generate the valence state wave packet, Ψ b. We neglect the Stark shift between the two states and non-adiabatic population transfer because the transition dipole moment between them is zero since they have the same parity. In this case, the phase evolution in Eq. 1 is given by ϕ ( t ) ΔE /! τ + ϕ(0) (2) = c where ΔE is the energy separation between the E g and Α g states, and ϕ(0) is the initial phase shift which is determined at the time of tunnel ionization. For characterizing the wave packet, ΔE and ϕ(0) are the parameters to be adjusted. We rely on ΔE ~ 2eV from a result of ab-initio calculation and measured photoelectron spectrum [18]. At fixed ΔE = 2 ev, we adjusts ϕ(0) over 2π range to fit the observed φ HHG with the calculate φ HHG distribution. We found that ϕ(0) = 0.4 π is a well-fitted value. In Fig. 4 a and b, we show the calculated φ HHG at ϕ(0) = -0.4 π and ϕ(0) = 0.4 π, respectively. We choose ϕ(0) = -0.4 π to demonstrate how the φ HHG distribution is sensitive to the initial phase difference. As we change ϕ(0) from -0.4 π to 0.4 π, the two φ HHG peaks shift from small to large re-collision time. In order to show the relationship between φ HHG and Ψ b 2, we present the calculated Ψ b 2 at selected re-collision times of t c = 0.8 fs, 1.0 fs and 1.2 fs in case of φ(0) = 0.4 π in Fig. 4. To confirm the double φ HHG structure is due to the dynamics, we calculate φ HHG vs the harmonic number and the delay using the single E g or A g state, respectively. The sum of φ HHG peaks calculated from only the E g or A g state doesn t fit the observed φ HHG distribution. We also calculate the angular distribution of the tunnel ionization probability from the coherent superposition of the two states at ϕ(0) of 0.4. The distribution is approximately the same as the incoherent sum case, shown in Fig. 1b. Before concluding, we present posibilities how the phase shift between two states, ϕ(0), can be caused at the time of tunnel ionization. The re-colliding electron wave packet (Ψ c ) generated from different electronic states must constructively interfere with each other in order to lead to efficient high harmonic generation. This requirement may select ensembles of the bound state electron wavefunctions with a particular phase difference so that the Ψ c can constructively interfere. Other possibilities include the effect of the 6

laser fields which can change the time-evolution of the wavefunction, or the effect of different electrostatic potentials of the two channels on the outgoing electron. In summary, we have demonstrated that dynamical motion of the valence electron wave packet can be mapped on the high harmonic emission spectra generated by the two-color laser pulses. For futher applications, our approach can be extended to measuring dynamics associated with changes in the dipole moment such as internal electron motion induced by intense laser fields, core-level ionization, chemical reaction, external atom or molecular migration, and so on [19-20]. If we use an infrared laser pulse such as 2000 nm and its second harmonic, then we can increase θ c over 90 degrees and extend the harmonic number because the continuum electron achieves a higher re-collision energy. If the amplitude of the dipole moment vector as well as φ HHG can be obtained, then we expect to re-construct a 3D orbital image as it changes by combining with molecular orientation technique [21]. References [1] F. Krausz and M. Ivanov, Rev. Mod. Phys. 81, 163 (2009). [2] F. Remacle and R. D. Levine, Z. Phys. Chem. 221, 647 (2007). [4] G. Sansone et al., Nature 465, 763 (2010). [5] E. Goulielmakis et al., Nature 466, 739 (2010). [6] O. Smirnova et al. Nature 460, 972 (2009). [7] S. Haessler et al,. Nature Phys. 6, 200 (2010). [8] H. Akagi et al., Science 325, 5946 (2009). [9] T. Kanai, E. Takahashi, Y. Nabekawa and K. Midorikawa, Phys. Rev. Lett. 98, 153904 (2007). [10] H. Niikura, N. Dudovich, D. Villeneuve and P. B. Corkum, Phys. Rev. Lett. 105, 053003 (2010). [11] D. Shafir, et al., Nature Phys 5, 412 (2009). [12] Y. Mairesse et al., Science 302, 1540 (2003). [13] A. Staudte et al., Phys. Rev. Lett. 102, 033004 (2009). [14] M.W. Schmidt et al., J. Comput. Chem. 14, 1347 (1993). [15] See the supplementary material for the data and analysis. [16] P. Corkum, Phys. Rev. Lett. 71, 1994 (1993). [17] H. J. Wörner, H. Niikura, J. B. Bertrand, P. B. Corkum and D. M. Villeneuve, Phys. Rev. Lett. 102, 103901 (2009). [18] A. Richartz, R. J. Buenker P. J. Bruna and S. D. Peyerimhoff, Mol. Phys. 33, 1345 7

(1977). [19] H. J. Wörner et al., Nature 466, 604 (2010). [20] W. Li et al., Science 322, 1081 (2008). [21] J. Itatani et al., Nature 432, 867 (2004). 8

Figure caption Figure 1 a. Tunnel ionization creates wave packet pairs in the ionization continuum (Ψ c ) and in the valence state (Ψ b ). Upon re-collision between two wave packets, the high harmonics are emitted. At each re-collision time, we scan the re-collision angle (θ c ) with respect to the molecule by scanning the phase delay between an 800 nm and a 400 nm laser pulses. The time of re-collision is determined by the observed photon energy of the high harmonic emission. We read the structure of the valence electron wave packet from the relationship between the polarization angle of the high harmonics and the re-collision angle (see text). b. (left) Schematic diagram of the calculated, selected valence-shell electronic structure of C 2 H 6. (right) The calculated angular distribution of the tunnel ionization probability for the most loosely bound energy level (E g state, magenta line) and the second most loosely bound energy level (A g, red line) of C 2 H 6 at the laser intensity of 2.0 x 10 14 W/cm 2. We show only one of the E g degenerated wave function. The dotted line is the incoherent sum of these two lines. The ionization probability of the other states can be neglected. Figure 2 High-harmonic generation spectra from C 2 H 6 vs the delay between the two laser pulses. We choose the zero-delay at an arbitrary phase. The total harmonic intensity modulates with the period of ~0.66 fs while the intensity of 14 th and 16 th harmonics modulates with the period of ~ 0.33 fs. Figure 3 φ HHG vs the delay between the two laser pulses for a H 2 and b C 2 H 6. We interpolate φ HHG along the two axis. In the right axis, we plot the re-collision time which is calculated from the harmonic number. Figure 4 Calculated φ HHG (degrees) vs the harmonic number and the phase delay at an initial phase (ϕ(0)) of (a) -0.4 π and (b) 0.4 π. We use the energy difference between the most and the second most loosely bound state of C 2 H 6, ΔE = 2.0 ev. The right panel shows the calculated Ψ b 2 and the total phase ϕ in Eq. 2 at selected re-collision time. 9

Figure 1 (Niikura et al.) a Ψ c Ψ b θc θ c t = 0 t c t c b Ionization energy E g A g E u Ionization rate x 10 14 4 135 2 0 180 2 225 4 90 270 45 0 315 10

Figure 2 (Niikura et al.) Harmonic number 23 22 21 20 19 18 17 16 15 4E3 7.3E3 1.3E4 2.4E4 4.4E4 8E4 14 13 0.0 0.5 1.0 1.5 2.0 0 0.3 0.66 0.99 1.32 Relative two-color delay Two-color delay / fs 11

Figure 3 (Niikura et al.) Harmonic number 21 19 17 15 13 21 19 17 15 a b 13 0.0 0.2 0.4 0.6 0.8 1.0 1.2 Two-color delay / fs 1.4 1.2 1.0 0.8 1.4 1.2 1.0 0.8 φ HHG Time / fs 6.0 11 16 21 26 30 15 20 25 30 35 40 43 b 12

Figure 4 (Niikura et al.) Ψ b 2 ϕ (a) Harmonic number (b) Harmonic number φ 0 10 20 30 40 HHG 21 1.4 19 1.3 17 1.2 15 1.1 1.0 13 0.9 11 0.8 21 1.4 19 1.3 17 1.2 15 1.1 1.0 13 0.9 11 0.8 0.0 0.2 0.4 0.6 0.8 1.0 1.2 Two-color delay / fs Time / fs 3.0 π 2.0 π 1.5 π 1.0 π 13