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

Role of excited states in molecular alignment-dependent ionization

Open Access Open Access

Abstract

We introduce an ab initio approach and the modified strong-field approximation to investigate the alignment-dependent ionization of H2+(1πu) exposed to different few-cycle laser fields. The ab initio calculations are performed by the B-splines one-center method and the Crank-Nicolson method in spherical coordinates. It is shown that the peak ionization probabilities appear around alignment angles 50° and 40° at the laser intensities 3×1013 W/cm2 and 5×1013 W/cm2, respectively, and the above distinct features come from the resonant excitation of the molecular ion, which is confirmed by calculation including and excluding the state 2σg in the basis expansion. Furthermore, the results obtained by including the state 2σg in the ab initio simulations can be qualitatively reproduced by the modified molecular length gauge strong-field approximation (SFA) taking account of the 1πu and 2σg states simultaneously. Analysis indicates that a part of electron is directly emitted from the 1πu orbital and another portion of electron is released from 2σg orbital and other excited state after the single-photon resonant transition between 1πu and 2σg orbitals.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Investigations of excited states have attracted considerable attentions in intense-field physics recently, because excited states play important roles in the generation of fascinating strong-field phenomena. One typical example is the effect of excited state on ionization behavior. It is found that charge-resonance-enhanced ionization occurs for extended molecules due to the strong coupling of charge states (1σg and 1σu) for H2+ and H2 [1, 2], and the above distinct feature has been observed experimentally [3]. For asymmetric molecule HeH2+, the coupling of the ground state and the excited state shows dependence on the carrier-envelope phase of few-cycle pulses, which has an important impact on the enhanced ionization process [4]. Recently, it is shown that the excited states are related to the very low or near-zero energy structure in the photoelectron spectrum in intense laser fields [5, 6]. For high-energy structures in photoelectron spectra, it is also found that the excited states have an important impact on the resonance-like enhancement in high-order above threshold ionization [7, 8]. Another example is the role of excited states in high-order harmonic generation (HHG). Atoms (molecules) in excited state are prepared to increase HHG yields and extend the cutoff frequency of the harmonic spectrum compared with systems in ground state [9–11]. The excited orbital has a great impact on the dip and the cutoff of the harmonic spectra for stretched H2+ in intense laser fields [12, 13]. Recently, it is predicted that the resonant population of excited states during the laser pulse plays a role in fractional-order harmonics and spectral redshift [14, 15], and the aforementioned effect of ionization from excited states and recombination to the ground state has been verified experimentally [16].

Due to improvements in molecular orientation and alignment technology [17], study of the orientation effect on strong-field ionization has attracted significant attention lately. The molecular tunneling ionization probability shows dependence on the alignment of the molecular axis with regard to the laser field for small internuclear distance, and the alignment-dependent ionization probability can be used to reconstruct the contour of the spatial electron distribution of the active orbital, which stimulates a number of experimental studies on molecular angular-dependent ionization [18, 19]. However, for CO2, the narrow ionization distribution and its maximum ionization at alignment angle 45° are observed in experiments, while the maximum ionization shows up at alignment angle near 25° for the theoretical result [19]. For OCS, the experimental observation of the peak ionization probability appears at alignment angle 90°, while the theory predicts at 45° [20]. Considerable efforts have been devoted to addressing the above problems, but the underlying mechanisms behind their anomalous alignment-dependent ionization are still under debate [21–25].

In theory, the methods of time-dependent Schrödinger equation (TDSE) and strong-field approximation (SFA) are widely used to investigate angular-dependent ionization of molecules with small internuclear distances [26–30]. For TDSE, although it is time-consuming and requires large computation resources, the roles of coulomb potential and excited states are taken into account. For SFA, the contribution of excited states to the evolution of the system is neglected, and the length-gauge SFA is more appropriate to study molecular alignment-dependent ionization with respect to velocity-gauge SFA [31, 32]. Recently, oscillation of the ionization yields is shown for H2+ as a function of internuclear distances in a high-frequency laser field parallel to the molecular axis, and the above peculiar feature disappears for the polarization direction of the laser field perpendicular to the molecular axis, which is traced to the high-energy electron two-center interference [33]. Afterwards, it is demonstrated that the molecular ionization probability shows oscillatory behavior with respect to the alignment angle for large internuclear separation due to the low-energy electron interference [34].

As mentioned above, excited states have a significant impact on the ionization dynamics and HHG, while a detailed investigation of the effect of excited states on the molecular angular-dependent ionization still lacks, which is in need for further understanding on the experimental findings, for instance, imaging electron molecular orbitals via ionization and HHG [18, 19, 35]. In this paper, we study the alignment-dependent ionization of H2+ by a three-dimensional TDSE and SFA, and our calculations show that the excited states play an important role in the orientation-dependent ionization of H2+. Atomic units (a.u.) are used unless otherwise indicated.

2. Theoretical methods

2.1. I. Time-dependent Schrödinger equation

Within the dipole approximation, the full-dimensional TDSE of H2+ in the length gauge is given in the following form

itΨ(r,t)=[H1(r)rE(t)]Ψ(r,t).

For a given magnetic quantum number m, the detailed expression of H1(r) is given by [36, 37]

H1(r)=12r2rr2r+12r2[ξ(1ξ2)ξm21ξ2]λ=0r<λr>λ+1Pλ(ξ)λ=0(1)λr<λr>λ+1Pλ(ξ).

Here r> (r<) denotes the bigger(smaller) one of (r,R2), and ξ = cos θ(θ is the angle between r and the internuclear distance R). Pλ(ξ) are the Legendre polynomials, and E(t) indicates the time-varying electric field. We assume that the polarization vector of the field lies in the plane x-z (the z axis is parallel to the molecular axis), and the laser-molecule interaction term is written as

Ht=rE(t)(ξcosβ+1ξ2sinβcosφ),
where β is the alignment angle between the polarization direction of the laser field and the molecular axis, and the field-free Schrödinger equation of H2+ is
H1(r)ψj(r,ξ,φ)=ϵjψj(r,ξ,φ).

The wave function is expanded by B-spline as

ψj(r,ξ,φ)=12πμvCμvjBμk(r)rBvk(ξ)(1ξ2)|m|2eimφ,
where k=7, and the details of B-splines can be found in Refs. [36, 38]. Diagonalizing the Hamiltonian in Eq. (4), we obtain the eigenvalues ϵj and the corresponding eigenfunction ψj , and the time-varying wave function is expanded by eigenfunctions
Ψ(r,t)=jCj(t)ψj(r,ξ,φ).

Substituting the above time-dependent wave function into Eq. (1), we propagate it by Crank-Nicolson method [36–38]. A cos1/8 absorber function is used near the boundary to reduce spurious reflection. At the end of the pulse, the ionization probability is defined as Pion = 1 − Σj |Cj |2 with ϵj < 0. To study the effect of a specific excited state on the ionization dynamics, we include or exclude it in the basis expansion artificially, which has been used to analyse the contribution of 1σu state to the enhanced ionization of H2+ subjected to intense laser fields recently [39].

In the present work, the vector potential is A(t)=E0ωεsin2(πt/tmax)cosωt with unit vector ε, 0 < t < tmax , and the time-dependent electric field is defined via E(t)=A(t)t. E0 is the peak electric field, and tmax and ω are the duration and the frequency of the laser pulse, respectively. In the following calculation, H2+ is exposed to a three-cycle laser pulse of frequency ω = 0.057 a.u. (λ = 800 nm), I0=1014 W/cm2 is used as the unit of laser intensity, and the time step is 0.1 a.u. The internuclear distance R=2 a.u. is assumed, and the radial box is truncated at rmax =90 a.u. 100 radial B-splines and 20 angular B-splines are adopted, the magnetic quantum numbers are kept from m=−6 to m=6, and convergence is reached with the above settings. The initial state is 1πu unless otherwise indicated.

2.2. II. Modified strong-field approximation

In this work, we also study the alignment-dependent ionization of H2+ adopting strong-field approximation (SFA) [31, 40–45]. Within the modified SFA, the corresponding wave function in the present paper can be assumed to be [43]

Ψ(r,t)=eiIpt[a0(t)|0>+a1(t)|1>+dPb(P,t)|P>],
where |0> and |1> denote ψ1πu and ψ2σg, respectively.
|0>=c1[ϕ2px(r+R/2)+ϕ2px(rR/2)+0.1(ϕ2px(r+R/2)+ϕ2px(rR/2))],|1>=c2[ϕ2s(r+R/2)+ϕ2s(rR/2)].

Here c1 and c2 are the normalization factors. Then the equations of a0(t) and a1(t) are given by

ia˙0(t)=E(t)[a0(t)d00+a1(t)d01],ia˙1(t)=E(t)[a0(t)d10+a1(t)d11]+(Ip1Ip2)a1(t).

Here the time-varying electric field is defined as E(t)=A(t)t, and the vector potential is the same as that aforementioned in TDSE calculations. dij (i, j=0, 1) represent the transition matrices between the bound states, and Ip1 and Ip2 indicate the ionization potentials of 1πu and 2σg states, respectively.

For the SFA theory in the length gauge, the single ionization rate of a molecule in the state |s> (s=0, 1) under the linearly polarized laser field is written as

Ws=2πNen=n0|Tpns|2δ(Ep+Ipnω)dP,
with
Tpns=1T0Tdtexp[i(E028ω3sin2ωt+E0ePω2sinωt)]×exp[i(P22+Ip+Up)t]E0sinωtϕ~[P+A(t)]as(t).
e is the unit vector, Ne indicates the number of equivalent electrons in the valence orbital, and n0 is the minimum number of photons needed to release the electron. Ep =Up + P2/2 indicates the quasienergy, where Up(=E02/4ω2) represents the ponderomotive energy, and P2/2=UpIp indicates the kinetic energy of the emitted electron after absorbing n photons for the ionization potential Ip . In the following calculations, we employ ϕ~[P+A(t)]=2cos[(P+A(t))R/2]dj[P+A(t)] for undressed modified molecular SFA [31, 43, 44], and dj [P+A(t)] (j = 2px , 2px or 2s) indicate the atomic dipole moment of the 2px , 2px, and 2s orbitals for the initial states of 1πu and 2σg, respectively. The total ionization probability is defined as W = ΣWs (s=0, 1). The 2px orbital is given by ϕ2px(r)=k52rπexp(kr)sinθcosφ with k=1.52Ip1 and Ip1=0.428 a.u. (ϕ2px(r)=k52rπexp(kr)sinθcosφ with k=0.42Ip1), and the 2s orbital is written as ϕ2s(r)=1πk52(1kr)exp(kr) with k=1.352Ip2 and Ip2=0.36 a.u. In this work, Eqs. (9) and (10) are calculated numerically [34, 45].

3. Results and discussion

We present the energy of the bound states for various values of magnetic quantum numbers m=0 and 1 in Table I, which is in a good accordance with those in Ref. [26]. We also show the contour plot of the electronic wave function for 1σg, 1σu, 2σg and 1πu states in Fig. 1. It is demonstrated that the charge distribution of σ orbital is parallel to the molecular axis, while the electron cloud of 1πu orbital is distributed perpendicularly to the internuclear axis.

Tables Icon

Table 1. The eigenvalue (a.u.) of the bound states for magnetic quantum numbers m=0 and 1, and the internuclear separation R=2 a.u.

 figure: Fig. 1

Fig. 1 Two-dimensional distribution of the wave function for different orbitals. (a): 1σg; (b): 1σu; (c): 2σg; (d): 1πu .

Download Full Size | PDF

In order to visualize orientation effect, the ionization probability is normalized to its maximum value unless otherwise indicated. Figure 2 displays the angular dependence of ionization yields for H2+(1πu) calculated by different theoretical methods. Fig. 2(a) shows the results obtained by TDSE, which has been checked by our previous numerical procedure [34]. For 0.3I0, it can be seen that the ionization probability increases from 0° to 50°, while a quick decrease of the ionization yield is present between 60° and 90°. For 0.5I0, the peak ionization yield shows up at the alignment angle 40°. In Fig. 2(b), we depict the alignment-dependent ionization of H2+ obtained by SFA calculation (2σg state is not taken into account in Eq. (7)), and it is found that the ionization yields increase with ascending alignment angle for 0.3I0 and 0.5I0 due to the spacial distribution of 1πu perpendicular to the molecular axis in Fig. 1(d). The normalized ionization probability of 0.3I0 is close to those of 0.5I0 due to the small difference of the above two laser intensities. Clearly the location of peak ionization probability in Fig. 2(b) is different from that in Fig. 2(a). It is well known that the effect of excited states is not taken into account for SFA, which maybe account for the discrepancy of the alignment-dependent ionization obtained by TDSE and SFA calculations in Fig. 2.

 figure: Fig. 2

Fig. 2 Alignment-dependent ionization probabilities of H2+ for the initial state 1πu as a function of the laser intensity calculated by different methods. (a): TDSE; (b): SFA without considering 2σg state in Eq. 7 (see text). I0=1014W/cm2.

Download Full Size | PDF

Now let’s turn to the maximum ionization probability, which occurs at different alignment angles for TDSE and SFA in Fig. 2. In Table 1, the energy gap between 2σg and 1πu states is about 0.068 a.u., which is close to the photon energy 0.057 a.u. (λ = 800 nm) of the laser field, so a strong coupling of single photon transition may occur between the 2σg and 1πu states. Fig. 3 depicts the time-dependent populations of 1πu , 2σg and other excited states as a function of time with different alignment angles for 0.3I0 and 0.5I0, respectively. In general, the populations of the 2σg state grow with increasing alignment angles, since the coupling between 1πu state and 2σg state becomes stronger with increasing alignment angle. Consequently, 2σg state plays an important role in the ionization of H2+(1πu) when the alignment angle is large. In addition, it can be found in Fig. 3 that the total populations of other excited states are even considerably larger than that of the 2σg state and also increase when the alignment angle raises from 30° to 90° for a specific laser intensity. In Fig. 4, we also present the time-dependent populations of 2σu and 3σg states with increasing alignment angle for different laser intensities. It is found that the time-varying probabilities of 2σu and 3σg decrease as a function of alignment angle for a specific laser intensity, because the coupling of 2σg state and other σ states becomes weak with increasing alignment angle. It should be noted that the population of 2σu state is negligible in Figs. 4(c) and (f), since the coupling of 2σg and 2σu states is zero for β=90°. In Figs. 4(c) and 4(f), the time-dependent probabilities of 3σg states are much larger than those of 2σu states due to the coupling between 1πu and 3σg states.

 figure: Fig. 3

Fig. 3 Time-varying probabilities of 1πu and 2σg, and other excited states calculated by TDSE with increasing alignment angle β for different laser intensities. (a)–(c): 0.3I0, (d)–(f): 0.5I0. T = 2π/ω and I0=1014W/cm2.

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 Time-varying probabilities of 2σu and 3σg states obtained by TDSE as a function of alignment angles β for different laser intensities. (a)–(c): 0.3I0, (d)–(f): 0.5I0. T = 2π/ω and I0=1014W/cm2.

Download Full Size | PDF

To see the significance of the 2σg state in the alignment-dependent ionization of H2+ more clearly, we presents the ionization yields obtained by removing the contribution of the 2σg state artificially from the TDSE simulation in Fig. 5 for 0.3I0 and 0.5I0. Because there is no coupling between 2σg and 1πu states at alignment angle β = 0°, the alignment-dependent ionization yields are studied from β = 10° to β = 90° in Fig. 5(a). The ionization probabilities increase monotonously as a function of the alignment angle, which is qualitatively consistent with the result of the SFA calculation shown in Fig. 2(b) but different from those in Fig. 2(a). In Figs. 5(b) and 5(c), we also depict the time-dependent probabilities of other excited states as a function of the alignment angles for 0.3I0 and 0.5I0, respectively. It is shown that the probabilities in Fig. 5(b) and 5(c) are much smaller with respect to those shown in Fig. 3, which indicates that the other excited states are populated mainly via transition from the 2σg state but not directly from the 1πu state. Therefore, the 2σg state has an important impact on the alignment dependence of the ionization for H2+ exposed to different laser fields.

 figure: Fig. 5

Fig. 5 Simulations of H2+ via TDSE by excluding 2σg state in the basis expansion with the initial state of 1πu. (a): Alignment-dependent ionization probability. (b) and (c): Time-varying probabilities of other excited states with increasing alignment angles for different laser intensities. T = 2π/ω and I0=1014W/cm2.

Download Full Size | PDF

To shed more physical insights into the ionization dynamics of H2+ in different laser fields, we show the alignment-dependent ionization probability obtained by the SFA theory with initial state 1πu and 2σg state taken into account (see Eq. (7)) in Fig. 6. In Figs. 6(a) and 6(b), the alignment-dependent ionization probabilities of 1πu and 2σg states are displayed, respectively. In Fig. 6(a), the maximal ionization yield from 1πu appears at alignment angle β=50° and 30° for laser intensities of 0.3I0 and 0.5I0, respectively, which is different from that in Fig. 2(b). The distribution of the electron cloud of 1πu is perpendicular to the molecular axis (see Fig. 1(d)) and the depletion of 1πu state is not taken into account, therefore the ionization yield increases with ascending alignment angles as shown in Fig. 2(b). However, the coupling of 1πu and 2σg states becomes stronger when the alignment angle increases, hence the population of the 1πu state drops with increasing alignment angle (see Fig. 7). So the ionization yield does not increase monotonically like that in Fig. 2(b). In addition, since the coupling between the 1πu and 2σg states becomes stronger with increasing laser intensity for a specific alignment angle, the population of the 1πu state decreases faster for 0.5I0 than that of 0.3I0 (see Fig. 7), so the alignment angle corresponding to the peak ionization yield shifts to lower angle when the laser intensity increases. Moreover, the ionization yield of the 2σg also increases as a function of the alignment angle (see Fig. 6(b)), though its wavefunction distributes along the molecular axis as shown in Fig. 1(c). This can be attributed to that the coupling between the 1πu and 2σg states becomes stronger when the alignment angle increases, leading to fast increasing population of the 2σg state with ascending alignment angle (see Fig. 7).

 figure: Fig. 6

Fig. 6 Alignment-dependent ionization probabilities calculated by SFA with increasing laser intensities for the initial state 1πu (2σg state is taken into account in Eq. (7)). (a): The ionization yields of 1πu state. (b): The ionization probabilities of 2σg state. (c): Sum of the ionization probabilities of 1πu and 2σg states. I0=1014W/cm2.

Download Full Size | PDF

 figure: Fig. 7

Fig. 7 Time-dependent probabilities of 1πu and 2σg states calculated by SFA with increasing alignment angles for different laser intensities (2σg state is taken into account in Eq. (7)). (a)–(c): 0.3I0; (d)–(f): 0.5I0. T = 2π/ω and I0=1014W/cm2.

Download Full Size | PDF

In Fig. 6(c), the total alignment-dependent ionization yield of the 1πu and 2σg states are present, and it can be seen that the peak ionization probability shifts from β=60° to 40° with increasing laser intensity, which is in a reasonable agreement with the TDSE simulation in Fig. 2(a). It is worthwhile mentioning that the coupling of the 1πu and 2σg states are much stronger than that obtained by the TDSE calculation (see Fig. 3). Since the other excited states are not taken into account in the modified SFA, the 2σg state can be considered to play the role of all the excited states (comparing Fig. 7 with Fig. 3), which gives rise to the qualitative agreement on the positions of the peaks in the alignment-dependent ionization probabilities calculated by the TDSE and modified SFA theory. However, it should be noted that the ionization probability calculated by the SFA at angle larger than 60° is noticeably higher than that of the TDSE calculation. This discrepancy can be attributed to that, since only the 2σg excited state is included, the population of the 1πu state will be retrieved after two optical cycles via Rabi oscillation between these two states in the SFA calculation (e. g., see Figs. 7(b) and (c)) while in the TDSE with all the excited states considered, large portion of the population of the electron will be pumped to the other excited state but not transfers back to the 1πu state after it is pumped to the 2σg state (e. g., see Figs. 3(b) and (c)). Consequently, the ionization channel of 1πu → 2σg → other excited states (such as 2σu and 3σg states in Fig. 4)→ionization gives rise to the decrease of ionization after β = 50° and β = 40° for 0.3I0 and 0.5I0 in Fig. 2(a), respectively.

4. Conclusions

In summary, we investigate the angular-dependent ionization of H+ 2 subjected to few-cycle laser fields using a three-dimensional TDSE based on single-center method, B-splines in spherical coordinates and Crank-Nicolson propagation method. It is found that the peak ionization probability occurs at alignment angles of β=50° and 40° for intensities of 0.3I0 and 0.5I0 (I0=1014W/cm2) for the initial state 1πu, respectively. However, the ionization yield increases monotonically as a function of the alignment angle if the 2σg state is excluded in the basis expansion, which is qualitatively in agreement with the calculation of SFA for the initial state 1πu without considering the contribution of the 2σg state. When the 2σg state is taken into account in the modified SFA, it gives peak ionization probability at β=60° and 40° for intensities of 0.3I0 and 0.5I0, in qualitative agreement with the complete TDSE calculation. Our analysis indicates that a part of electron is directly released from the 1πu orbital by the laser field, and another portion of the electron firstly transits from 1πu to the 2σg state via single-photon transition and transits to the other excited states thereafter, then the electron is ionized by the laser field from the excited states. The maximal ionization yields at alignment angles β=50° and 40° for 0.3I0 and 0.5I0 can be attributed to the sum of contributions from different ionization channels. Our work demonstrates that the excited states have an important impact on molecular alignment-dependent ionization in the laser field, which may be related to the peculiar angular dependence of ionization dynamics for CO2 and OCS molecules in the strong laser pulses [19, 20].

Funding

National Key Research and Development program of China (2016YFA0401100); National Natural Science Foundation of China (NSFC) (11774361, 11804405, 11425414, 11504215); Fundamental Research Fund of Sun Yat-Sen University (20187100031610008).

Acknowledgments

This work was supported by the National Key R&D program of China (No. 2016YFA0401100) and the NSFC of China (Grants No. 11774361, No. 11334009, No. 11425414, No. 11374197, No. 11804405, and No. 11504215), and the Fundamental Research Fund of SUN YAT-SEN UNIVERSITY (No. 20187100031610008).

References

1. T. Zuo and A. D. Bandrauk, “Charge-resonance-enhanced ionization of diatomic molecular ions by intense lasers,” Phys. Rev. A 52, R2511 (1995). [CrossRef]   [PubMed]  

2. A. Saenz, “Enhanced ionization of molecular hydrogen in very strong fields,” Phys. Rev. A 61, 051402 (2000). [CrossRef]  

3. E. Constant, H. Stapelfeldt, and P. B. Corkum, “Observation of Enhanced Ionization of Molecular Ions in Intense Laser Fields,” Phys. Rev. Lett . 76, 4140 (1996). [CrossRef]   [PubMed]  

4. G. L. Kamta and A. D. Bandrauk, “Phase Dependence of Enhanced Ionization in Asymmetric Molecules,” Phys. Rev. Lett . 94, 203003 (2005). [CrossRef]   [PubMed]  

5. B. Wolter, C. Lemell, M. Baudisch, M. G. Pullen, X. M. Tong, M. Hemmer, A. Senftleben, C. D. Schröter, J. Ullrich, R. Moshammer, J. Biegert, and J. Burgdörfer, “Formation of very-low-energy states crossing the ionization threshold of argon atoms in strong mid-infrared fields,” Phys. Rev. A 90, 063424 (2014). [CrossRef]  

6. L. Zhao, J. W. Dong, H. Lv, T. X. Yang, Y. Lian, M. X. Jin, H. F. Xu, D. J. Ding, S. L. Hu, and J. Chen, “Ellipticity dependence of neutral Rydberg excitation of atoms in strong laser fields,” Phys. Rev. A 94, 053403 (2016). [CrossRef]  

7. H. G. Muller, “Numerical simulation of high-order above-threshold-ionization enhancement in argon,” Phys. Rev. A 60, 1341 (1999). [CrossRef]  

8. B. B. Li, S. L. Hu, X. T. He, and J. Chen, “Resonance-like enhancement in high-order above threshold ionization of atoms and molecules in intense laser fields,” Opt. Express 26, 13012 (2018). [CrossRef]   [PubMed]  

9. P. M. Paul, T. O. Clatterbuck, C. Lyngå, P. Colosimo, L. F. DiMauro, P. Agostini, and K. C. Kulander, “Enhanced High Harmonic Generation from an Optically Prepared Excited Medium,” Phys. Rev. Lett . 94, 113906 (2005). [CrossRef]   [PubMed]  

10. B. B. Wang, J. Chen, J. Liu, Z. C. Yan, and P. M. Fu, “Attosecond-pulse-controlled high-order harmonic generation in ultrashort laser fields,” Phys. Rev. A 78, 023413 (2008). [CrossRef]  

11. J. Heslar, D. A. Telnov, and S. I. Chu, “Enhancement of VUV and EUV generation by field-controlled resonance structures of diatomic molecules,” Phys. Rev. A 93, 063401 (2016). [CrossRef]  

12. Y. C. Han and L. B. Madsen, “Minimum in the high-order harmonic generation spectrum from molecules: role of excited states,” J. Phys. B 43, 225601 (2010). [CrossRef]  

13. Y. C. Han and L. B. Madsen, “Internuclear-distance dependence of the role of excited states in high-order-harmonic generation of H2+,” Phys. Rev. A 87, 043404 (2013). [CrossRef]  

14. X. B. Bian and A. D. Bandrauk, “Multichannel Molecular High-Order Harmonic Generation from Asymmetric Diatomic Molecules,” Phys. Rev. Lett . 105, 093903 (2010). [CrossRef]   [PubMed]  

15. X. B. Bian and A. D. Bandrauk, “Nonadiabatic molecular high-order harmonic generation from polar molecules: Spectral redshift,” Phys. Rev. A 83, 041403 (2011). [CrossRef]  

16. S. Beaulieu, S. Camp, D. Descamps, A. Comby, V. Wanie, S. Petit, F. Légaré, K. J. Schafer, M. B. Gaarde, F. Catoire, and Y. Mairesse, “Role of Excited States In High-order Harmonic Generation,” Phys. Rev. Lett . 117, 203001 (2016). [CrossRef]   [PubMed]  

17. H. Stapelfeldt and T. Seideman, “Colloquium: Aligning molecules with strong laser pulses,” Rev. Mod. Phys . 75, 543 (2003). [CrossRef]  

18. I. V. Litvinyuk, K. F. Lee, P. W. Dooley, D. M. Rayner, D. M. Villeneuve, and P. B. Corkum, “Alignment-Dependent Strong Field Ionization of Molecules,” Phys. Rev. Lett . 90, 233003 (2003). [CrossRef]   [PubMed]  

19. D. Pavičić, K. F. Lee, D. M. Rayner, P. B. Corkum, and D. M. Villeneuve, “Direct Measurement of the Angular Dependence of Ionization for N2, O2, and CO2 in Intense Laser Fields,” Phys. Rev. Lett . 98, 243001 (2007). [CrossRef]  

20. J. L. Hansen, L. Holmegaard, J. H. Nielsen, H. Stapelfeldt, D. Dimitrovski, and L. B. Madsen, “Orientation-dependent ionization yields from strong-field ionization of fixed-in-space linear and asymmetric top molecules,” J. Phys. B 45, 015101 (2012). [CrossRef]  

21. S. F. Zhao, C. Jin, A. T. Le, T. F. Jiang, and C. D. Lin, “Analysis of angular dependence of strong-field tunneling ionization for CO2,” Phys. Rev. A 80, 051402 (2009). [CrossRef]  

22. S. Petretti, Y. V. Vanne, A. Saenz, A. Castro, and P. Decleva, “Alignment-Dependent Ionization of N2, O2, and CO2 in Intense Laser Fields,” Phys. Rev. Lett . 104, 223001 (2010). [CrossRef]  

23. R. Johansen, K. G. Bay, L. Christensen, J. Thøgersen, D. Dimitrovski, L. B. Madsen, and H. Stapelfeldt, “Alignment-dependent strong-field ionization yields of carbonyl sulfide molecules induced by mid-infrared laser pulses,” J. Phys. B 49, 205601 (2016). [CrossRef]  

24. V. P. Majety and A. Scrinzi, “Dynamic Exchange in the Strong Field Ionization of Molecules,” Phys. Rev. Lett . 115, 103002 (2015). [CrossRef]   [PubMed]  

25. M. D. Śpiewanowski and L. B. Madsen, “Alignment-and orientation-dependent strong-field ionization of molecules: Field-induced orbital distortion effects,” Phys. Rev. A 91, 043406 (2015). [CrossRef]  

26. G. L. Kamta and A. D. Bandrauk, “Three-dimensional time-profile analysis of high-order harmonic generation in H2+,” Phys. Rev. A 71, 053407 (2005). [CrossRef]  

27. T. K. Kjeldsen, L. B. Madsen, and J. P. Hansen, “Ab initio studies of strong-field ionization of arbitrarily oriented H2+ molecules,” Phys. Rev. A . 74, 035402 (2006). [CrossRef]  

28. S. Petretti, Á. Magaña, A. Saenz, and P. Decleva, “Wavelength-and alignment-dependent photoionization of N2 and O2,” Phys. Rev. A . 94, 053411 (2016). [CrossRef]  

29. D. A. Telnov and S. I. Chu, “Ab initio study of the orientation effects in multiphoton ionization and high-order harmonic generation from the ground and excited electronic states of H2+,” Phys. Rev. A 76, 043412 (2007). [CrossRef]  

30. B. Zhang, J. M. Yuan, and Z. X. Zhao, “Alignment-dependent ionization of Hharmonic generation from the ground and excited electronic states of H2+: From multiphoton ionization to tunneling ionization,” Phys. Rev. A 85, 033421 (2012). [CrossRef]  

31. D. B. Milošević, “Strong-field approximation for ionization of a diatomic molecule by a strong laser field,” Phys. Rev. A 74, 063404 (2006). [CrossRef]  

32. B. Zhang and Z. X. Zhao, “Strong-field approximation for the ionization of N2 and CO2,” Phys. Rev. A 82, 035401 (2010). [CrossRef]  

33. S. Selstø, M. Førre, J. P. Hansen, and L. B. Madsen, “Strong Orientation Effects in Ionization of H2+ by Short, Intense, High-Frequency Light Pulses,” Phys. Rev. Lett . 95, 093002 (2005). [CrossRef]  

34. S. L. Hu, J. Chen, X. L. Hao, and W. D. Li, “Effect of low-energy electron interference on strong-field molecular ionization,” Phys. Rev. A 93, 023424 (2016). [CrossRef]  

35. J. Itatani, J. Levesque, D. Zeidler, H. Niikura, H. Pépin, J. C. Kieffer, P. B. Corkum, and D. M. Villeneuve, “Tomographic imaging of molecular orbitals,” Nature (London) 432, 867 (2004). [CrossRef]  

36. S. L. Hu, Z. X. Zhao, and T. Y. Shi, “B-spline one-center method for molecular Hartree-Fock calculations,” Int. J. Quan. Chem. 114, 441 (2014). [CrossRef]  

37. S. L. Hu, Z. X. Zhao, J. Chen, and T. Y. Shi, “Ionization dynamics of C2H2 in intense laser fields: Time-dependent Hartree-Fock approach,” Phys. Rev. A 92, 053409 (2015). [CrossRef]  

38. H. Bachau, E. Cormier, P. Decleva, J. E. Hansen, and F. Martín, “Applications of B-splines in atomic and molecular physics,” Rep. Pro. Phys . 64, 1815 (2001). [CrossRef]  

39. Y. J. Jin, X. M. Tong, and N. Toshima, “Enhanced ionization of hydrogen molecular ions in an intense laser field via a multiphoton resonance,” Phys. Rev. A 81, 013408 (2010). [CrossRef]  

40. L. V. Keldysh, “Ionization in the field of a strong electromagnetic wave,” Zh. Eksp. Teor. Fiz. 47, 1945 (1964) [Sov. Phys. JETP 20, 1307 (1965)].

41. F. H. M. Faisal, “Multiple absorption of laser photons by atoms,” J. Phys. B 6, L89 (1973). [CrossRef]  

42. H. R. Reiss, “Effect of an intense electromagnetic field on a weakly bound system,” Phys. Rev. A 22, 1786 (1980). [CrossRef]  

43. J. Chen and S. G. Chen, “Gauge problem in molecular high-order harmonic generation,” Phys. Rev. A 75, 041402 (2007). [CrossRef]  

44. W. Becker, J. Chen, S. G. Chen, and D. B. Milošević, “Dressed-state strong-field approximation for laser-induced molecular ionization,” Phys. Rev. A 76, 033403 (2007). [CrossRef]  

45. Y. Li, X. Y. Jia, S. P. Yang, W. D. Li, and J. Chen, “S-matrix study on alignment dependence of single ionization of molecular ions with different active orbitals,” Chin. Phys. B 19, 063302 (2010). [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 Two-dimensional distribution of the wave function for different orbitals. (a): 1σg; (b): 1σu; (c): 2σg; (d): 1πu .
Fig. 2
Fig. 2 Alignment-dependent ionization probabilities of H 2 + for the initial state 1πu as a function of the laser intensity calculated by different methods. (a): TDSE; (b): SFA without considering 2σg state in Eq. 7 (see text). I0=1014W/cm2.
Fig. 3
Fig. 3 Time-varying probabilities of 1πu and 2σg, and other excited states calculated by TDSE with increasing alignment angle β for different laser intensities. (a)–(c): 0.3I0, (d)–(f): 0.5I0. T = 2π/ω and I0=1014W/cm2.
Fig. 4
Fig. 4 Time-varying probabilities of 2σu and 3σg states obtained by TDSE as a function of alignment angles β for different laser intensities. (a)–(c): 0.3I0, (d)–(f): 0.5I0. T = 2π/ω and I0=1014W/cm2.
Fig. 5
Fig. 5 Simulations of H 2 + via TDSE by excluding 2σg state in the basis expansion with the initial state of 1πu. (a): Alignment-dependent ionization probability. (b) and (c): Time-varying probabilities of other excited states with increasing alignment angles for different laser intensities. T = 2π/ω and I0=1014W/cm2.
Fig. 6
Fig. 6 Alignment-dependent ionization probabilities calculated by SFA with increasing laser intensities for the initial state 1πu (2σg state is taken into account in Eq. (7)). (a): The ionization yields of 1πu state. (b): The ionization probabilities of 2σg state. (c): Sum of the ionization probabilities of 1πu and 2σg states. I0=1014W/cm2.
Fig. 7
Fig. 7 Time-dependent probabilities of 1πu and 2σg states calculated by SFA with increasing alignment angles for different laser intensities (2σg state is taken into account in Eq. (7)). (a)–(c): 0.3I0; (d)–(f): 0.5I0. T = 2π/ω and I0=1014W/cm2.

Tables (1)

Tables Icon

Table 1 The eigenvalue (a.u.) of the bound states for magnetic quantum numbers m=0 and 1, and the internuclear separation R=2 a.u.

Equations (11)

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

i t Ψ ( r , t ) = [ H 1 ( r ) r E ( t ) ] Ψ ( r , t ) .
H 1 ( r ) = 1 2 r 2 r r 2 r + 1 2 r 2 [ ξ ( 1 ξ 2 ) ξ m 2 1 ξ 2 ] λ = 0 r < λ r > λ + 1 P λ ( ξ ) λ = 0 ( 1 ) λ r < λ r > λ + 1 P λ ( ξ ) .
H t = r E ( t ) ( ξ cos β + 1 ξ 2 sin β cos φ ) ,
H 1 ( r ) ψ j ( r , ξ , φ ) = ϵ j ψ j ( r , ξ , φ ) .
ψ j ( r , ξ , φ ) = 1 2 π μ v C μ v j B μ k ( r ) r B v k ( ξ ) ( 1 ξ 2 ) | m | 2 e i m φ ,
Ψ ( r , t ) = j C j ( t ) ψ j ( r , ξ , φ ) .
Ψ ( r , t ) = e i I p t [ a 0 ( t ) | 0 > + a 1 ( t ) | 1 > + d P b ( P , t ) | P > ] ,
| 0 > = c 1 [ ϕ 2 p x ( r + R / 2 ) + ϕ 2 p x ( r R / 2 ) + 0.1 ( ϕ 2 p x ( r + R / 2 ) + ϕ 2 p x ( r R / 2 ) ) ] , | 1 > = c 2 [ ϕ 2 s ( r + R / 2 ) + ϕ 2 s ( r R / 2 ) ] .
i a ˙ 0 ( t ) = E ( t ) [ a 0 ( t ) d 00 + a 1 ( t ) d 01 ] , i a ˙ 1 ( t ) = E ( t ) [ a 0 ( t ) d 10 + a 1 ( t ) d 11 ] + ( I p 1 I p 2 ) a 1 ( t ) .
W s = 2 π N e n = n 0 | T p n s | 2 δ ( E p + I p n ω ) d P ,
T p n s = 1 T 0 T d t exp [ i ( E 0 2 8 ω 3 sin 2 ω t + E 0 e P ω 2 sin ω t ) ] × exp [ i ( P 2 2 + I p + U p ) t ] E 0 sin ω t ϕ ~ [ P + A ( t ) ] a s ( t ) .
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.