Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Relative phase effect of nonsequential double ionization of molecules by counter-rotating two-color circularly polarized fields

Open Access Open Access

Abstract

Relative phase effect of nonsequential double ionization (NSDI) of aligned molecules by counter-rotating two-color circularly polarized (TCCP) fields is investigated with a three-dimensional classical ensemble model. Numerical results show that NSDI yield in counter-rotating TCCP fields sensitively depends on the relative phase of the two components, which exhibits a sin-like behavior with the period of π/2. NSDI yield achieves its maximum at the relative phase π/8 and minimum at 3π/8. Back analysis indicates the recollision time and the return angle of the electron strongly depend on the relative phase of the two components, which results in the dominant emission direction of the electrons, is different for different relative phases. This indicates that the recollision process can be steered by changing the relative phase of the two components in counter-rotating TCCP laser fields. Meantime, it provides an avenue to obtain information about the recollision time and the return angle in the recollision process from the electron momentum distribution.

© 2020 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

The two electrons from nonsequential double ionization (NSDI) in strong laser fields are highly correlated by a recollision process between the return electron and the parent ion [1,2]. The correlation behavior attracted much attention and is deeply studied experimentally [311] and theoretically [1221] in the past decades. Two typical ionization channels, direct recollision-impact ionization (RII) [5,22] and recollision-induced excitation with subsequent ionization (RESI) [2227], have been well demonstrated.

In recent years, much attention is concentrated on two-color circularly polarized (TCCP) laser fields, which consists of two circularly polarized pulses with different frequencies. The wave shape of the combined electric field can flexibly be tailored by changing the parameters of those components constituting the combined electric fields [2830]. It provides an effective means for steering the electron dynamics in two dimensional (2D) space. Recently, the TCCP laser field is used to generate high-brightness circularly polarized harmonics in the extreme ultraviolet and soft x-ray regions [31]. A subcycle interference in the light propagation direction [32] and photoelectron holographic interferences [33,34] have been shown for ionization of atoms by counter-rotating TCCP laser fields. An attoclock photoelectron interferometry for probing the phase and the amplitude of emitting wave packets [35] is achieved by co-rotating TCCP fields. The nonadiabatic offset of the initial electron momentum distribution in strong field tunneling ionization of argon [36] has been demonstrated by TCCP fields.

The studies of NSDI in TCCP laser fields indicate that NSDI yields and recollision trajectories strongly depend on the field amplitude ratios of two components [3741]. Double-recollision trajectory and its intensity dependence are demonstrated in NSDI by counter-rotating TCCP laser fields [42,43]. Very recently, the angular distributions of the correlated electrons in NSDI of Mg in counter-rotating TCCP laser fields are discussed [44]. Electron recollision assisted enhanced ionization is also experimentally observed for N$_2$ molecule in counter-rotating TCCP laser fields [45]. In these studies on NSDI above, the phase difference between two components constituting the combined electric fields is fixed. As we know, a change of the relative phase between the two components can result in a corresponding rotation of the wave shape of the combined electric field in the 2D space, which will affect the directions of the emission and the returning of the electron. For atoms the change of the relative phase will not affect the NSDI yield and the electron momentum distributions because atoms are spherically symmetric. However, for aligned molecules, the spherical symmetry is broken. The change of the relative phase will significantly affect the ultrafast electron dynamics and the occurrence of NSDI of aligned molecules. Now the dependence of NSDI of aligned molecules on the relative phase in counter-rotating TCCP fields is still unclear.

In this paper, we focus on the relative phase effect of NSDI of aligned molecules by counter-rotating TCCP fields. Our numerical results show that NSDI yield in counter-rotating TCCP fields sensitively depends on the relative phase of the two components, which achieves its maximum at the relative phase $\pi$/8 and minimum at 3$\pi$/8. Back analysis indicates that the recollision time and the return angle of the electron strongly depend on the relative phase of the two components. It results in that the dominant emission direction of the electrons is different for different relative phases. This provides an avenue to obtain information about the recollision time and the return angle in the recolision process from the electron momentum distribution.

2. Classical ensemble model

In this paper, we employ the three-dimensional (3D) classical ensemble model with a soft-core potential for the Coulomb interactions proposed by Eberly and coworkers [46,47], which has been widely recognized as a reliable and useful approach for interpretation and prediction of strong-field double ionization phenomena [16,17,4144,48,49]. In this model, the evolution of the two-electron system is determined by the Newton’s equations of motion (atomic units are used throughout unless stated otherwise):

$$\frac{d^{2}\textbf{r}_{i}}{dt^{2}}={-}\nabla [V_{ne}(\textbf{r}_{i})+V_{ee}(\textbf{r}_{1},\textbf{r}_{2})]-\textbf{E}(t),$$
where the subscript $i$=1, 2 is the label of the two electrons and r$_i$ is the coordinate of the i$_{th}$ electron. In this work the target is a model diatomic molecule aligned along the x axis. The two cores are fixed at (-$R$/2,0,0) and ($R$/2,0,0) respectively. Here, the internuclear distance is set to be $R$=2.0 a.u. The interaction between electron and the parent ion can be modeled by the two-center soft-core Coulomb potential $V_{ne}(\textbf {r}_{i})=-1/\sqrt {(x_i-R/2)^2+y_i^2+z_i^2+a^2}-1/\sqrt {(x_i+R/2)^2+y_i^2+z_i^2+a^2}$. The softening parameter a=1.25 is introduced here to avoid unphysical autoionization. The electron-electron interaction is $V_{ee}(\textbf {r}_{1},\textbf {r}_{2})=1/\sqrt {(\textbf {r}_{1}-\textbf {r}_{2})^2+b^2}$. The softening parameter b is included primarily for numerical stability and here is set to be 0.05.

The electric field of the laser pulse is given by

$$\textbf{E}(t)=E_{1600}f(t)[cos(\omega t +\phi)\widehat{x}-sin(\omega t +\phi)\widehat{y}]+E_{800}f(t)[cos(2\omega t)\widehat{x}+sin(2\omega t)\widehat{y}],$$
where $\widehat {x}$ and $\widehat {y}$ are the unit vectors along the x and y directions, respectively. $E_{1600}$ and $E_{800}$ are the electric field amplitudes of the 1600-nm and 800-nm pulses, respectively. Here the 1600-nm field rotates clockwise and the 800-nm field rotates anticlockwise. Their intensities are set to be the same value 5$\times$10$^{13}$ W/cm$^2$. $\phi$ is the relative phase between the two circularly polarized fields. Here the period of the 1600-nm pulse is defined as T. Because the period of the time evolution of the two-color combined field is the same as the period of the 1600-nm pulse, T also represents the period of the two-color combined field. $f(t)$ is the envelope of the laser pulse which has a trapezoidal shape with 2T turn on, 6T at full strength, and 2T turn off.

To obtain the initial conditions for Eq. (1), the ensemble is populated starting from a classically allowed position for the energy of -1.67 a.u., corresponding to the sum of the first and second ionization potentials of H$_2$. The available kinetic energy is distributed between the two electrons randomly, and the directions of the momentum vectors of the two electrons are also randomly assigned. Then the two-electron system is allowed to evolve a sufficient long time (400 a.u.) in the absence of the laser field to obtain stable position and momentum distributions. Once the initial ensemble is obtained, the laser field is turned on and all trajectories are evolved in the combined Coulomb and laser fields. We check the energies of the two electrons at the end of the laser pulse, and a double ionization event is determined if both electrons achieve positive energies, where the energy of each electron contains the kinetic energy, potential energy of the electron-ion interaction, and half electron-electron repulsion energy.

3. Results and discussions

Figure  shows NSDI probabilities as a function of the relative phase between the two components constituting the combined electric fields. The NSDI probability sensitively depends on the relative phase. It exhibits a sin-like behavior with the relative phase increasing. The period is $\pi$/2. The NSDI probability achieves its maximum at $\phi$=$\pi$/8 and 5$\pi$/8 and its minimum at $\phi$=3$\pi$/8 and 7$\pi$/8.

 figure: Fig. 1.

Fig. 1. NSDI probabilities as a function of the relative phase between the two components.

Download Full Size | PDF

In order to understand the periodicity of the NSDI probability vs the relative phase, we check the variation of the space structure of the combined electric field with the relative phase. As shown in Fig. 2, the electric field and negative vector potential trace out a trefoil pattern and a triangle, respectively. A lobe of the electric field corresponds to a side of the triangle of the negative vector potential. Each lobe is marked by different colors. The negative vector potentials corresponding to the electric field maxima are located in the middle of each side of the triangle of the negative vector potential, which are indicated by the square blocks in Fig. 2. The arrows indicate the time evolution direction. The electric field recursively evolves from lobe 1 (blue), lobe 2 (red) to lobe 3 (green) and the negative vector potential evolves from side 1 (blue), side 2 (red) to side 3 (green). In the polarization plane the angle between any two lobes is 120$^{\circ }$. As the relative phase increases the trefoil-shape electric field and the triangle-shape negative vector potential both rotate clockwise. Moreover, for each $\pi$/8 increase of the relative phase, the trefoil-shape electric field rotates 15$^{\circ }$. From Fig. 2 one can see that when the relative phase increases $\pi$ the space structure of the combined field is completely the same. When the relative phase increases $\pi$/2, the wave shapes of the two combined fields are symmetrical with respect to the origin. In the two combined fields both angels the electron is emitted from and returns to the parent ion are the same with respect to the molecular axis. Thus NSDI probabilities are equal for those combined fields with the difference $\pi$/2 of the relative phase. This is the reason that the period of the NSDI probability vs the relative phase is $\pi$/2.

 figure: Fig. 2.

Fig. 2. The combined laser electric field E(t) (dashed) and the corresponding negative vector potential -A(t) (solid) for different relative phases. The arrows indicate the time evolution direction. The electric field and negative vector potential trace out a trefoil pattern and a triangle, respectively. A lobe of the electric field corresponds to a side of the triangle of the negative vector potential. The square blocks mark the field maxima and their negative vector potential. The black dumbbell represents the diatomic molecule aligned along x axis.

Download Full Size | PDF

To obtain a deep understanding for ultrafast dynamics of NSDI in counter-rotating TCCP fields with different relative phases, we trace the classical NSDI trajectories. Figures 3(a)-3(c) show a sample trajectory from NSDI events at $\phi$=0. Figure 3(a) shows electron distances from the parent ion versus time. One electron is ionized at t=6.05T and then returns and collides with the parent ion at t=6.32T. The electron travels for 0.27T from single ionization to the recollision. It is obvious that prominent energy transfer from the returning electron to the bound electron takes place during the recollision, as shown in Fig. 3(b). The electrons’ 2D path in the field plane for the trajectory is shown in Fig. 3(c). The electron ionizes at lobe 3 and returns at lobe 1. So firstly the electron moves to -x direction and then returns from the third quadrant. We define the angle between the direction of the momentum of the return electron at the instant 3 a.u. before the recollision and the +x direction as the return angle. For the trajectory in Fig. 3(c), the return direction is opposite to the electric field vector and the return angle of the electron is 41$^{\circ }$. Furthermore, we do statistical analysis on the traveling time (the time interval between single ionization and the recollision) and find its distributions for different relative phases are similar. They all show a single-peak structure and the center of the peak is around 0.28T. The distribution of the traveling time for $\phi$=0 is shown in Fig. 3(d). It indicates that the recollision always occurs at the next electric field lobe after single ionization. Figures 3(e) and 3(f) show the electrons’ path in the field plane for another two typical sample NSDI trajectories for $\phi$=0. In Fig. 3(e) the electron is ionized at lobe 2 and returns at lobe 3 with the return angle of 160$^{\circ }$. For the trajectory in Fig. 3(f) the electron is ionized at lobe 1 and return at lobe 2 with the return angle of 283$^{\circ }$. This indicates that the electron can return to and collide with the parent ion from different angles via counter-rotating TCCP fields.

 figure: Fig. 3.

Fig. 3. Planes (a), (b) and (c) show electron distances from the parent ion versus time, electron energies versus time and the electrons’ path in the field plane for a sample NSDI trajectory. The arrows indicate the time evolution direction. The dotted line marks the return direction of the free electron. (d) Distribution of the traveling time for $\phi$=0. Planes (e) and (f) show the electrons’ path in the field plane for another two sample NSDI trajectories with different return angle. These trajectories are from those NSDI events for $\phi$=0.

Download Full Size | PDF

Based on the statistical analysis of those classical NSDI trajectories, we obtain the distributions of single ionization time t$_{SI}$ and the recollision time t$_R$, as shown in the first and second rows of Fig. 4 for $\phi$=0 (the first column), $\pi$/8 (the second column), $\pi$/4 (the third column) and 3$\pi$/8 (the fourth column). By comparing these distributions from different relative phases, one can see that the differences of NSDI yields triggered by the single ionization at lobe 1 and lobe 2 between different relative phases are small. The main difference of NSDI yields between different relative phases originates from those NSDI events induced by the single ionization at lobe 3. For the electric field with a specific amplitude, the smaller the angle between the electric field and the molecular axis is, the larger the probability of the single ionization is. From Figs. 2(a)–2(d), one can see that the angle between the lobe 3 and the molecular axis increases gradually with the relative phase increasing. So in the range of the relative phase (0, $\pi$/2) the probability of single ionization from lobe 3 decreases gradually as the relative phase increases. Those electrons ionized at lobe 3 return to and recollide with the parent ion at lobe 1. After recollision the occurrence of the double ionization is significantly affected by the height of the suppressed potential barrier by the lobe 1. Previous study has indicated that the smaller the angle between the electric field and molecular axis is, the lower the suppressed potential barrier is [50]. From Figs. 2(a)–2(d), one can see that the angle between the lobe 1 and molecular axis decreases gradually with the relative phase increasing. That is to say, in the range of the relative phase (0, $\pi$/2), as the relative phase increases the probability of single ionization from lobe 3 decreases and the suppressed potential barrier at the recollision lowers gradually. The former is unfavorable and the latter is favorable to occurrence of NSDI. The competition between the two factors results in that the NSDI yield achieves its maximum at $\phi$=$\pi$/8 and minimum at $\phi$=3$\pi$/8.

 figure: Fig. 4.

Fig. 4. Single ionization time (the first row), recollision time (the second row) and return angle distribution (the third row) of the electron are shown from up to down for $\phi$=0 (the first column), $\pi$/8 (the second column), $\pi$/4 (the third column) and 3$\pi$/8 (the fourth column). In order to more clearly show the laser phase at single ionization and recollision, the single ionization time and recollision time are transferred into the interval of [(k+$\frac {\pi -\phi }{6\pi }$)T, (k+$\frac {\pi -\phi }{6\pi }$+1)T], in which the combined electric field evolves from lobe 1 to lobe 3. The laser electric field in arbitrary units is shown in planes (a)-(h).

Download Full Size | PDF

Figures 4(i)–4(l) show the return angle distributions of the free electron for $\phi$=0, $\pi$/8, $\pi$/4 and 3$\pi$/8 respectively. It is obvious that the free electron always returns to the parent ion from three directions for all relative phases, which correspond to the three single ionization bursts from three field lobes. The return direction of the electron is approximately opposite to the electric field direction at the recollision. So the specific return angle is strongly dependent on the relative phase. Similar to the trefoil-shape electric field, with the relative phase increasing the return angle distribution rotates clockwise. For $\phi$=0, $\pi$/8 and $\pi$/4, there is a return direction in which the return electrons result in more NSDI events. The particular return angles are 42$^{\circ }$, 25$^{\circ }$ and 11$^{\circ }$ respectively. For $\phi$=3$\pi$/8, those NSDI events returning from 0$^{\circ }$ and 240$^{\circ }$ are comparable. The result indicates that the return angle of the electron can be well controlled by changing the relative phase of the two components in counter-rotating TCCP fields.

Figure 5 shows the momentum distributions of the two electrons in the field plane for eight relative phases. The corresponding negative vector potentials -A(t) are also shown. According to simple-man model, where the initial momentum at the ionization instant and the effect of the Coulomb potential on the ionized electron are ignored, the final momentum of the ionized electron is equal to the negative vector potential -A(t) at the ionization instant. So the electrons mainly distribute along the negative vector potential curve. But these electrons do not uniformly distribute on the three sides of the negative vector potential curve. For $\phi$=0, $\pi$/8, $\pi$/4, more electrons cluster on the side 1 of the negative vector potential triangle. For $\phi$=3$\pi$/8, the electrons on the side 1 and side 2 are comparable and more than those on the side 3. Because the negative vector potential curve rotates clockwise with the relative phase increasing, the emission direction of the electrons changes with the relative phase. In addition, the electron momentum distributions for $\phi$=0, $\pi$/8, $\pi$/4, $\phi$=3$\pi$/8 and $\phi$=$\pi$/2, 5$\pi$/8, 3$\pi$/4, $\phi$=7$\pi$/8 are symmetrical with respect to the origin. It is because that their electric fields are symmetrical with respect to the origin.

 figure: Fig. 5.

Fig. 5. Electron momentum distributions in the field plane for $\phi$=0 (a), $\pi$/8 (b), $\pi$/4 (c), 3$\pi$/8 (d), $\pi$/2 (e), 5$\pi$/8 (f), 3$\pi$/4 (g) and 7$\pi$/8 (h). Both electrons are included. The negative vector potentials -A(t) are shown.

Download Full Size | PDF

Based on the final ionization order after recollision the two electrons are defined as the first and the second electron. We separately present the momentum distributions of the first electrons (the first row) and the second electrons (the second row) for $\phi$=0, $\pi$/8, $\pi$/4 and $\phi$=3$\pi$/8 in Fig. 6. For all relative phases, the distributions of the second electrons on the three sides of the negative vector potential are similar. The distributions of the first electrons determine the shape of the total momentum distribution of two electrons on the three sides. Experimentally, it is not possible to distinguish the two electrons involved in NSDI as we do. But the results above indicate that the measured total momentum distribution of two electrons in experiment can reflect the dominant emission direction of the first electrons. For the laser parameters considered in this work, after recollision the first electron is freed immediately. The second electrons are ionized with a prominent time delay with respect to the recollision as shown in Fig. 7. It indicates that most of NSDI proceed by RESI channel. If the initial momentum at the ionization instant and the effect of the Coulomb potential on the ionized electron are ignored, finally the first electron is most likely emitted with the momentum -A(t$_R$). This establishes a connection between the final momentum of the first electron and the recollision time. From Figs. 4(e), 4(f) and 4(g), one can see that more electrons return to and recollide with the parent ion at lobe 1 for $\phi$=0, $\pi$/8 and $\pi$/4. After recollision the first electron is ionized immediately and thus more first electrons distribute on the side 1. For $\phi$=3$\pi$/8, those NSDI events recolliding at lobe 1 and lobe 2 are comparable. Thus the first electrons mostly distribute on side 1 and side 2 of the negative vector potential triangle. So for different relative phases, the dominant NSDI events originate from different recollision times, different return angles and finally the electrons are emitted to different directions. Conversely, the electron momentum distribution can be regarded as the map of the recollision time and the return angle. The information about the recollision time and the return angle can be derived from the electron momentum distribution. In addition, our results also demonstrate that the recollision process can be steered by changing the relative phase of the two components in counter-rotating TCCP laser fields.

 figure: Fig. 6.

Fig. 6. Momentum distributions of the first and second electrons in the field plane for $\phi$=0, $\pi$/8, $\pi$/4 and 3$\pi$/8.

Download Full Size | PDF

 figure: Fig. 7.

Fig. 7. Distributions of the time separation between the double ionization and the recollision for $\phi$=0 (blue), $\pi$/8 (red), $\pi$/4 (green) and 3$\pi$/8 (magenta).

Download Full Size | PDF

4. Conclusion

In conclusion, we have investigated NSDI of aligned molecules by counter-rotating TCCP fields for different relative phases. Numerical results indicate that NSDI yield in counter-rotating TCCP fields sensitively depends on the relative phase of the two components, which achieves its maximum at the relative phase $\pi$/8 and minimum at 3$\pi$/8. Back analysis shows that the recollision time and the return angle of the electron can be well steered by changing the relative phase of the two components in counter-rotating TCCP laser fields. Correspondingly, it also results in the difference of the dominant emission direction of the electrons between different relative phases. The information about the recollision time and the return angle can be derived from the electron momentum distribution. This provides an avenue to access the subcycle dynamics of the recollision process.

Funding

National Natural Science Foundation of China (11504302, 61475127); Fundamental Research Funds for the Central Universities (XDJK2019C110).

Disclosures

The authors declare no conflicts of interest.

References

1. P. B. Corkum, “Plasma perspective on strong-field multiphoton ionization,” Phys. Rev. Lett. 71(13), 1994–1997 (1993). [CrossRef]  

2. K. C. Kulander, J. Cooper, and K. J. Schafer, “Laser-assisted inelastic rescattering during above-threshold ionization,” Phys. Rev. A 51(1), 561–568 (1995). [CrossRef]  

3. D. N. Fittingoff, P. R. Bolton, B. Chang, and K. C. Kulander, “Observation of nonsequential double ionization of helium with optical tunneling,” Phys. Rev. Lett. 69(18), 2642–2645 (1992). [CrossRef]  

4. B. Walker, B. Sheehy, L. F. DiMauro, P. Agostini, K. J. Schafer, and K. C. Kulander, “Precision measurement of strong field double ionization of helium,” Phys. Rev. Lett. 73(9), 1227–1230 (1994). [CrossRef]  

5. Th. Weber, H. Giessen, M. Weckenbrock, G. Urbasch, A. Staudte, L. Spielberger, O. Jagutzki, V. Mergel, M. Vollmer, and R. Dörner, “Correlated electron emmision in multiphoton double ionization,” Nature 405(6787), 658–661 (2000). [CrossRef]  

6. A. Staudte, C. Ruiz, M. Schöffler, S. Schössler, D. Zeidler, Th. Weber, M. Meckel, D. M. Villeneuve, P. B. Corkum, A. Becker, and R. Dörner, “Binary and recoil collisions in strong field double ionization of Helium,” Phys. Rev. Lett. 99(26), 263002 (2007). [CrossRef]  

7. A. Rudenko, V. L. B. de Jesus, Th. Ergler, K. Zrost, B. Feuerstein, C. D. Schröter, R. Moshammer, and J. Ullrich, “Correlated two-electron momentum spectra for strong-field nonsequential double ionization of He at 800 nm,” Phys. Rev. Lett. 99(26), 263003 (2007). [CrossRef]  

8. B. Bergues, M. Kübel, N. G. Johnson, B. Fischer, N. Camus, K. J. Betsch, O. Herrwerth, A. Senftleben, A. M. Sayler, T. Rathje, T. Pfeifer, I. Ben-Itzhak, R. R. Jones, G. G. Paulus, F. Krausz, R. Moshammer, J. Ullrich, and M. F. Kling, “Attosecond tracing of correlated electron-emission in non-sequential double ionization,” Nat. Commun. 3(1), 813 (2012). [CrossRef]  

9. B. Wolter, M. G. Pullen, M. Baudisch, M. Sclafani, M. Hemmer, A. Senftleben, C. D. Schröter, J. Ullrich, R. Moshammer, and J. Biegert, “Strong-field physics with mid-IR fields,” Phys. Rev. X 5(2), 021034 (2015). [CrossRef]  

10. Y. Wang, S. Xu, W. Quan, C. Gong, X. Lai, S. Hu, M. Liu, J. Chen, and X. Liu, “Recoil-ion momentum distribution for nonsequential double ionization of Xe in intense midinfrared laser fields,” Phys. Rev. A 94(5), 053412 (2016). [CrossRef]  

11. T. Shaaran, N. Camus, J. Dura, L. Fechner, A. Thai, A. Britz, M. Baudisch, T. Steinle, A. Senftleben, C. D. Schröter, J. Ullrich, T. Pfeifer, C. H. Keitel, J. Biegert, K. Z. Hatsagortsyan, and R. Moshammer, “Role of high ponderomotive energy in laser-induced nonsequential double ionization,” Phys. Rev. A 99(2), 023421 (2019). [CrossRef]  

12. C. Figueira de Morisson Faria and X. Liu, “Electron-electron correlation in strong laser fields,” J. Mod. Opt. 58(13), 1076–1131 (2011). [CrossRef]  

13. W. Becker, X. Liu, P. Jo Ho, and J. H. Eberly, “Theories of photoelectron correlation in laser-driven multiple atomic ionization,” Rev. Mod. Phys. 84(3), 1011–1043 (2012). [CrossRef]  

14. M. Lein, E. K. U. Gross, and V. Engel, “Intense-field double ionization of helium: identifying the mechanism,” Phys. Rev. Lett. 85(22), 4707–4710 (2000). [CrossRef]  

15. J. S. Parker, B. J. S. Doherty, K. T. Taylor, K. D. Schultz, C. I. Blaga, and L. F. DiMauro, “High-energy cutoff in the spectrum of strong-field nonsequential double ionization,” Phys. Rev. Lett. 96(13), 133001 (2006). [CrossRef]  

16. X. Wang and J. H. Eberly, “Elliptical polarization and probability of double ionization,” Phys. Rev. Lett. 105(8), 083001 (2010). [CrossRef]  

17. F. Mauger, C. Chandre, and T. Uzer, “Strong field double ionization: the phase space perspective,” Phys. Rev. Lett. 102(17), 173002 (2009). [CrossRef]  

18. X. Hao, J. Chen, W. Li, B. Wang, X. Wang, and W. Becker, “Quantum effects in double ionization of argon below the threshold intensity,” Phys. Rev. Lett. 112(7), 073002 (2014). [CrossRef]  

19. A. S. Maxwell and C. Figueira de Morisson Faria, “Controlling below-threshold nonsequential double ionization via quantum interference,” Phys. Rev. Lett. 116(14), 143001 (2016). [CrossRef]  

20. Z. Chen, Y. Liang, and C. D. Lin, “Quantum theory of recollisional (e, 2e) process in strong field nonsequential double ionization of helium,” Phys. Rev. Lett. 104(25), 253201 (2010). [CrossRef]  

21. D. Ye, M. Li, L. Fu, J. Liu, Q. Gong, Y. Liu, and J. Ullrich, “Scaling laws of the two-electron sum-energy spectrum in strong-field double ionization,” Phys. Rev. Lett. 115(12), 123001 (2015). [CrossRef]  

22. B. Feuerstein, R. Moshammer, D. Fischer, A. Dorn, C. D. Schröter, J. Deipenwisch, J. R. Crespo Lopez-Urrutia, C. Höhr, P. Neumayer, J. Ullrich, H. Rottke, C. Trump, M. Wittmann, G. Korn, and W. Sandner, “Separation of recollision mechanisms in nonsequential strong field double ionization of Ar: the role of excitation tunneling,” Phys. Rev. Lett. 87(4), 043003 (2001). [CrossRef]  

23. N. Camus, B. Fischer, M. Kremer, V. Sharma, A. Rudenko, B. Bergues, M. Kübel, N. G. Johnson, M. F. Kling, T. Pfeifer, J. Ullrich, and R. Moshammer, “Attosecond correlated dynamics of two electrons passing through a transition state,” Phys. Rev. Lett. 108(7), 073003 (2012). [CrossRef]  

24. Y. Chen, Y. Zhou, Y. Li, M. Li, P. Lan, and P. Lu, “Rabi oscillation in few-photon double ionization through doubly excited states,” Phys. Rev. A 97(1), 013428 (2018). [CrossRef]  

25. Y. Liu, L. Fu, D. Ye, J. Liu, M. Li, C. Wu, Q. Gong, R. Moshammer, and J. Ullrich, “Strong-field double ionization through sequential release from double excitation with subsequent coulomb scattering,” Phys. Rev. Lett. 112(1), 013003 (2014). [CrossRef]  

26. C. Huang, W. Guo, Y. Zhou, and Z. Wu, “Role of coulomb repulsion in correlated-electron emission from a doubly excited state in nonsequential double ionization of molecules,” Phys. Rev. A 93(1), 013416 (2016). [CrossRef]  

27. Q. Liao, A. H. Winney, S. K. Lee, Y. Lin, P. Adhikari, and W. Li, “Coulomb-repulsion-assisted double ionization from doubly excited states of argon,” Phys. Rev. A 96(2), 023401 (2017). [CrossRef]  

28. C. A. Mancuso, D. D. Hickstein, P. Grychtol, R. Knut, O. Kfir, X. Tong, F. Dollar, D. Zusin, M. Gopalakrishnan, C. Gentry, E. Turgut, J. L. Ellis, M. Chen, A. Fleischer, O. Cohen, H. C. Kapteyn, and M. M. Murnane, “Strong-field ionization with two-color circularly polarized laser fields,” Phys. Rev. A 91(3), 031402 (2015). [CrossRef]  

29. E. Hasovic, W. Becker, and D. B. Milǒsević, “electron rescattering in a bicircular laser field,” Opt. Express 24(6), 6413–6424 (2016). [CrossRef]  

30. C. A. Mancuso, D. D. Hickstein, K. M. Dorney, J. L. Ellis, E. Hasović, R. Knut, P. Grychtol, C. Gentry, M. Gopalakrishnan, D. Zusin, F. J. Dollar, X. Tong, D. B. Milǒsević, W. Becker, H. C. Kapteyn, and M. M. Murnane, “Controlling electron-ion rescattering in two-color circularly polarized femtosecond laser fields,” Phys. Rev. A 93(5), 053406 (2016). [CrossRef]  

31. A. Fleischer, O. Kfir, T. Diskin, P. Sidorenko, and O. Cohen, “Spin angular momentum and tunable polarization in high-harmonic generation,” Nat. Photonics 8(7), 543–549 (2014). [CrossRef]  

32. S. Eckart, M. Kunitski, I. Ivanov, M. Richter, K. Fehre, A. Hartung, J. Rist, K. Henrichs, D. Trabert, N. Schlott, L. Ph. H. Schmidt, T. Jahnke, M. S. Schöffler, A. Kheifets, and R. Dörner, “Subcycle interference upon tunnel ionization by counter-rotating two-color fields,” Phys. Rev. A 97(4), 041402 (2018). [CrossRef]  

33. M. Li, W. Jiang, H. Xie, S. Luo, Y. Zhou, and P. Lu, “Strong-field photoelectron holography of atoms by bicircular two-color laser pulses,” Phys. Rev. A 97(2), 023415 (2018). [CrossRef]  

34. Q. Ke, Y. Zhou, J. Tan, M. He, J. Liang, Y. Zhao, M. Li, and P. Lu, “Two-dimensional photoelectron holography in strong-field tunneling ionization by counter rotating two-color circularly polarized laser pulses,” Opt. Express 27(22), 32193–32209 (2019). [CrossRef]  

35. M. Han, P. Ge, Y. Shao, Q. Gong, and Y. Liu, “Attoclock photoelectron interferometry with two-color corotating circular fields to probe the phase and the amplitude of emitting wave packets,” Phys. Rev. Lett. 120(7), 073202 (2018). [CrossRef]  

36. S. Eckart, K. Fehre, N. Eicke, A. Hartung, J. Rist, D. Trabert, N. Strenger, A. Pier, L. Ph. H. Schmidt, T. Jahnke, M. S. Schöffler, M. Lein, M. Kunitski, and R. Dörner, “Direct experimental access to the nonadiabatic initial momentum offset upon tunnel ionization,” Phys. Rev. Lett. 121(16), 163202 (2018). [CrossRef]  

37. J. L. Chaloupka and D. D. Hickstein, “Dynamics of strong-field double ionization in two-color counterrotating fields,” Phys. Rev. Lett. 116(14), 143005 (2016). [CrossRef]  

38. C. A. Mancuso, K. M. Dorney, D. D. Hickstein, J. L. Chaloupka, J. L. Ellis, F. J. Dollar, R. Knut, P. Grychtol, D. Zusin, C. Gentry, M. Gopalakrishnan, H. C. Kapteyn, and M. M. Murnane, “Controlling nonsequential double ionization in two-color circularly polarized femtosecond laser fields,” Phys. Rev. Lett. 117(13), 133201 (2016). [CrossRef]  

39. S. Eckart, M. Richter, M. Kunitski, A. Hartung, J. Rist, K. Henrichs, N. Schlott, H. Kang, T. Bauer, H. Sann, L. Ph. H. Schmidt, M. Schöffler, T. Jahnke, and R. Döner, “Nonsequential double ionization by counterrotating circularly polarized two-color laser fields,” Phys. Rev. Lett. 117(13), 133202 (2016). [CrossRef]  

40. C. Huang, M. Zhong, and Z. Wu, “Nonsequential double ionization by co-rotating two-color circularly polarized laser fields,” Opt. Express 27(5), 7616–7626 (2019). [CrossRef]  

41. B. Li, X. Yang, X Ren, and J. Zhang, “Enhanced double ionization rate from O2 molecules driven by counter-rotating circularly polarized two-color laser fields,” Opt. Express 27(22), 32700–32708 (2019). [CrossRef]  

42. T. Xu, Q. Zhu, J. Chen, S. Ben, J. Zhang, and X. Liu, “Multiple recollisions in nonsequential double ionization by counter-rotating two-color circularly polarized laser fields,” Opt. Express 26(2), 1645–1654 (2018). [CrossRef]  

43. C. Huang, M. Zhong, and Z. Wu, “Intensity-dependent two-electron emission dynamics in nonsequential double ionization by counter-rotating two-color circularly polarized laser fields,” Opt. Express 26(20), 26045–26056 (2018). [CrossRef]  

44. X. Ma, Y. Zhou, Y. Chen, M. Li, Y. Li, Q. Zhang, and P. Lu, “Timing the release of the correlated electrons in strong-field nonsequential double ionization by circularly polarized two-color laser fields,” Opt. Express 27(3), 1825–1837 (2019). [CrossRef]  

45. K. Lin, X. Jia, Z. Yu, F. He, J. Ma, H. Li, X. Gong, Q. Song, Q. Ji, W. Zhang, H. Li, P. Lu, H. Zeng, J. Chen, and J. Wu, “Comparison study of strong-field ionization of molecules and atoms by bicircular two-color femtosecond laser pulses,” Phys. Rev. Lett. 119(20), 203202 (2017). [CrossRef]  

46. R. Panfili, J. H. Eberly, and S. L. Haan, “Comparing classical and quantum simulations of strong-field double-ionization,” Opt. Express 8(7), 431–435 (2001). [CrossRef]  

47. S. L. Haan, L. Breen, A. Karim, and J. H. Eberly, “Variable time lag and backward ejection in full-dimensional analysis of strong-field double ionization,” Phys. Rev. Lett. 97(10), 103008 (2006). [CrossRef]  

48. S. Luo, X. Ma, H. Xie, M. Li, Y. Zhou, W. Cao, and P. Lu, “Controlling nonsequential double ionization of Ne with parallel-polarized two-color laser pulses,” Opt. Express 26(10), 13666–13676 (2018). [CrossRef]  

49. A. Tong, Q. Li, X. Ma, Y. Zhou, and P. Lu, “Internal collision induced strong-field nonsequential double ionization in molecules,” Opt. Express 27(5), 6415–6425 (2019). [CrossRef]  

50. C. Huang, Y. Zhou, A. Tong, Q. Liao, W. Hong, and P. Lu, “The effect of molecular alignment on correlated electron dynamics in nonsequential double ionization,” Opt. Express 19(6), 5627–5634 (2011). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (7)

Fig. 1.
Fig. 1. NSDI probabilities as a function of the relative phase between the two components.
Fig. 2.
Fig. 2. The combined laser electric field E(t) (dashed) and the corresponding negative vector potential -A(t) (solid) for different relative phases. The arrows indicate the time evolution direction. The electric field and negative vector potential trace out a trefoil pattern and a triangle, respectively. A lobe of the electric field corresponds to a side of the triangle of the negative vector potential. The square blocks mark the field maxima and their negative vector potential. The black dumbbell represents the diatomic molecule aligned along x axis.
Fig. 3.
Fig. 3. Planes (a), (b) and (c) show electron distances from the parent ion versus time, electron energies versus time and the electrons’ path in the field plane for a sample NSDI trajectory. The arrows indicate the time evolution direction. The dotted line marks the return direction of the free electron. (d) Distribution of the traveling time for $\phi$=0. Planes (e) and (f) show the electrons’ path in the field plane for another two sample NSDI trajectories with different return angle. These trajectories are from those NSDI events for $\phi$=0.
Fig. 4.
Fig. 4. Single ionization time (the first row), recollision time (the second row) and return angle distribution (the third row) of the electron are shown from up to down for $\phi$=0 (the first column), $\pi$/8 (the second column), $\pi$/4 (the third column) and 3$\pi$/8 (the fourth column). In order to more clearly show the laser phase at single ionization and recollision, the single ionization time and recollision time are transferred into the interval of [(k+$\frac {\pi -\phi }{6\pi }$)T, (k+$\frac {\pi -\phi }{6\pi }$+1)T], in which the combined electric field evolves from lobe 1 to lobe 3. The laser electric field in arbitrary units is shown in planes (a)-(h).
Fig. 5.
Fig. 5. Electron momentum distributions in the field plane for $\phi$=0 (a), $\pi$/8 (b), $\pi$/4 (c), 3$\pi$/8 (d), $\pi$/2 (e), 5$\pi$/8 (f), 3$\pi$/4 (g) and 7$\pi$/8 (h). Both electrons are included. The negative vector potentials -A(t) are shown.
Fig. 6.
Fig. 6. Momentum distributions of the first and second electrons in the field plane for $\phi$=0, $\pi$/8, $\pi$/4 and 3$\pi$/8.
Fig. 7.
Fig. 7. Distributions of the time separation between the double ionization and the recollision for $\phi$=0 (blue), $\pi$/8 (red), $\pi$/4 (green) and 3$\pi$/8 (magenta).

Equations (2)

Equations on this page are rendered with MathJax. Learn more.

d 2 r i d t 2 = [ V n e ( r i ) + V e e ( r 1 , r 2 ) ] E ( t ) ,
E ( t ) = E 1600 f ( t ) [ c o s ( ω t + ϕ ) x ^ s i n ( ω t + ϕ ) y ^ ] + E 800 f ( t ) [ c o s ( 2 ω t ) x ^ + s i n ( 2 ω t ) y ^ ] ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.