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

Ultrafast attosecond-magnetic-field generation of the charge migration process based on HeH2+ and H2+ electronically excited by circularly polarized laser pulses

Open Access Open Access

Abstract

The ultrafast process by the electron in molecular ions from one site or region to another that has come to be known as charge migration (CM), which is of fundamental importance to photon induced chemical or physical reactions. In this work, we study the electron current and ultrafast magnetic-field generation based on CM process of oriented asymmetric (HeH2+) and symmetric (H2+) molecular ions. Calculated results show that they are ascribed to quantum interference of electronic states for these molecular ions under intense circularly polarized (CP) laser pulses. The two scenarios of (i) resonance excitation and (ii) direct ionization are considered through appropriately utilizing designed laser pulses. By comparison, the magnetic field induced by the scenario (i) is stronger than that of scenario (ii) for molecular ions. However, the scheme (ii) is very sensitive to the helicity of CP field, which is opposite to the scenario (i). Moreover, the magnetic field generated by H2+ is stronger than that by HeH2+ through scenario (i). Our findings provide a guiding principle for producing ultrafast magnetic fields in molecular systems for future research in ultrafast magneto-optics.

© 2021 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

With the development of laser technology, it has become a reality to image and manipulate the nuclear and electronic dynamics at ultra-short time and ultra-micro spatial scales in photophysics and photochemsitry processes. This has prompted the researches of laser-matter interactions in atomic and molecular systems, among which one of the basic issues is the study of electronic ultrafast dynamics in molecular reactions. Attosecond (10$^{-18}$ s) pulses [1,2] have provided important extreme physical conditions for these ultrafast processes, which has become a new tool for studying and ultimately controlling the electrons dynamics at the natural time scale [36]. Recently, the shortest attosecond pulses with a pulse duration of 43 as are available for new ultrafast photoelectron imaging technology [7]. However, the synthesis of ultrafast pulses on the sub-second time scale for imaging and controlling electronic movement is mostly limited to the linear polarization process [1,8]. Until recently, this new capability to generate circularly polarized coherent short wavelength harmonics using intense femtosecond bichromatic circularly polarized (CP) pulses has proved the potential value in experiments, which regard the X-ray system as a new imaging tool for new application of ultrafast X-ray magnetic circular dichroism [911]. Currently, CP has been widely used in many aspects, such as the generation of high-order harmonic and attosecond pulses in experiments and theories [1216].

Ultrafast charge migration (CM) induced by the coherent population excitation in multiple electronic states has attracted widespread attention in the fields of photophysics and photochemistry, or even the application of attosecond technology [1730], which provides convenience for controlling chemical reactions and detecting molecular orbitals. Recent work has shown that attosecond CM can be detected by using iodoacetylene molecular ions to generate high harmonics in heavy collisions [31]. On this basis, the current-carrying state in a semiconductor quantum ring is completely controlled using the quantum optimal control theory for terahertz laser pulses, which proves the importance of ultrafast pulses for magneto-optics [32]. So far, many efforts have been devoted to the generation of strong magnetic fields. Researchers demonstrated that the generation of electronic ring current in aromatic molecules can produce static magnetic fields through linear [33] and CP $\pi$ UV pulses resonance with the degenerate $\pi$ orbitals [3436]. Especially, this generated magnetic fields can be much larger than those obtained by traditional static fields [37]. Morever, the electron wave packets in H$_{2}^{+}$ generated by means of intense few-cycles attosecond CP UV pulses can also generate strong attosecond magnetic fields of several tens of Teslas [3843]. In addition, when the current-carrying state $2p_{+}$ of Ne atom is irradiated by an intense linearly polarized laser field, an spatial-localized oscillating magnetic field with an intensity up to 47 Tesla at the center can be induced [44]. However, the influence of molecular symmetry and asymmetry on the generation of magnetic fields remains unclear, which requires further exploration.

In this work, we report the generation of electron currents and ultrafast magnetic-field on the selected different orbitals of excited states for oriented symmetric H$_{2}^{+}$ and asymmetric HeH$^{2+}$ molecular ions under the CP attosecond XUV laser pulses by numerically solving the corresponding time-dependent Schrödinger equation (TDSE). The two scenarios of (i) resonant excitation, (ii) direct ionization are considered. For different molecular ions, we compare the electron currents and the corresponding ultra-fast strong magnetic field generated under these two scenarios. The results show that, in comparison to the simplest symmetric molecular ion H$_{2}^{+}$ from ground $1s\sigma _{g}$ electronic state, in the asymmetric molecular ion HeH$^{2+}$ from first excited $2p\sigma$ electronic state, the results obtained have similar features, but overall, the electron current and corresponding ultrafast magnetic field of symmetric molecular ion are stronger [1719].

The paper is organized as follows: We briefly describe computational methods for numerically solving 2D TDSEs of an aligned molecular ion HeH$^{2+}$ and H$_{2}^{+}$ by intense CP attosecond UV laser pulses in Sec. II. The results for aligned symmetric (H$_{2}^{+}$) and asymmetric (HeH$^{2+}$) molecular ions are discussed in Sec. III. Finally, we summarize our findings in Sec. IV. Throughout this paper, atomic units (a. u.) are used unless otherwise noted.

2. Theoretical model and computational methods

In this work, we perform numerical simulations on oriented single-electron molecular ions H$_{2}^{+}$ and HeH$^{2+}$ with a static nuclear frame. The corresponding TDSE is appropriately written with respect to the centre of mass of the two nuclei as

$$i\frac{\partial{}\psi{}(\textbf{r},t)}{\partial{}t}=[\frac{1}{2}{{\mathop\nabla\nolimits _\textbf{r}}^{2}}+V_{c}\left ( \textbf{r}\right )+V_{L}\left ( \textbf{r},t\right )]\ \psi{}\left(\textbf{r},t\right)$$

Here, a 2D (planar) model of the laser polarization in the molecular plane is adopted, which is adequate to describe electron motions in laser fields due to the 2D confinement of electron via CP laser pulses. The molecular axis is strictly oriented with the x-polarization axis. The corresponding soft-core Coulomb potential can be expressed as:

$$V_{c}\left ( \textbf{r}\right )={-}\frac{Z_{1}}{\sqrt{\left ( x+R_{1}\right )^{2}+y^{2}+\alpha }}-\frac{Z_{2}}{\sqrt{\left ( x-R_{2}\right )^{2}+y^{2}+\alpha}}$$
where x, y represent electronic coordinates, Z$_{1}$ and Z$_{2}$ are the charges of the nuclei. For H$_{2}^{+}$, Z$_{1}$ = Z$_{2}$ = 1 and R$_{1}$ = R$_{2}$ = R/2 (R= R$_{1}$ + R$_{2}$ =2 a. u. is the internuclear distance). For HeH$^{2+}$, Z$_{1}$ = 2, Z$_{2}$ = 1. For HeH$^{2+}$, the positions of the two nuclei coordinates can be written in a general form with respect to the centre of mass of two nuclei: R$_{1}$=-M$_{2}$R/(M$_{1}$+M$_{2}$) for He nuclei, and R$_{2}$ = M$_{1}$R/(M$_{1}$+M$_{2}$) for H nuclei, M$_{1}$ and M$_{2}$ are the masses of the He and H nuclei with M$_{1}$ = 4, M$_{2}$ = 1, and R = R$_{2}$ +R$_{1}$ = 3.89 a.u.. $\alpha$ is regularization parameter. Here, $\alpha$ = 0.74 is used to remove the singularity so as to accurately generate the electronic state potential energies of the I$_{p}$ = 1.08 a.u. for $1s\sigma _{g}$ orbitals in H$_{2}^{+}$, and $\alpha$ =0.5 is used to obtain the I$_{p}$ = 1.03 a.u. for $2p\sigma$ orbitals in HeH$^{2+}$. In these simulations, the H and He nuclei are placed at x$_{H}$ = +R$_{2}$ and x$_{He}$ = -R$_{1}$ on the x-axis.

The field-molecule interaction $V_{L}\left ( \textbf {r},t\right )=\textbf {r}\cdot E\left ( t\right )$ is treated under the length gauge and dipole approximation, since by exact solutions of the TDSE, numerical results are gauge invariant. The laser fields $E\left ( t\right )$ always propagates along the z axis, which is perpendicular to the molecular (x, y) plane. The forms for the CP pulses are defined as

$$\textbf{E}(t)=E_0f\left(t\right)\left[\cos{\left(\omega{}t\right)}{\hat{e}}_x+ \xi \sin{\left(\omega{}t\right)}{\hat{e}}_y\right] $$
where $\hat {e_{x}}$ / $\hat {e_{y}}$ are the laser field polarization vectors. $\omega$ is frequencies of the pulses. $\xi$ is the helicity of CP pulses. $\xi$= $\pm 1$ denote left/ right handed circularly polarized pulses in the (x, y) plane of the molecule. A temporal slowly varying envelope $f\left ( t\right )$= $sin^{2}\left ( \pi t/T\right )$ is used with duration T = 10 $\tau$ , where $\tau =2\pi /\omega$, in one optical cycle of pulses. Multiple optical cycle duration ensures that the field area $\int E\left ( t\right )dt=0$ to eliminate effects of static fields and carrier envelope phases of pulses.

The 2D TDSE in Eq. (1) is numerically solved by using the second order split-operator method algorithm combined with fast Fourier transform (FFT) technique [45]:

$$\begin{aligned}\psi(\textbf{r},t+\Delta{t})&=e^{{-}i(\frac{{p}^2}{4})\Delta{t}}e^{({-}iV_{c}(\textbf{r})-iV_{L}(\textbf{r},t))\Delta{t}}e^{{-}i(\frac{{p}^{2}}{4})\Delta{t}}{\psi(\textbf{r},t)}\\ &+O(\Delta{t})^3 \ \end{aligned}$$

In our calculation, the time step is taken to be $\Delta$t = 0.05 a. u. = 1.2 as. The spatial grid has 512 $\times$512 = 262144 grid points with the same step sizes for the x- and y-directions: $\Delta$x = $\Delta$y = 0.25 a.u. The grid ranges from −64 a.u. to 64 a.u. in each direction. The wave function is multiplied by a $cos^{\frac {1}{8}}$ mask function at each time step to prevent unphysical effects originating from the reflection of the wave packet from the boundary. And the domain of absorption ranges from $\left | x,y\right |$ = 50 a.u. to $\left | x,y\right |$ = 64 a.u..

The time-dependent electronic current density is defined by the quantum expression also in the length gauge,

$$\textbf{j}(\textbf{r},t)=\frac{i}{2}\left[\psi{}(\textbf{r},t){\nabla{}}_\textbf{r}{\psi{}}^*(\textbf{r},t)-{\psi{}}^*(\textbf{r},t){\nabla{}}_\textbf{r}\psi{}(\textbf{r},t)\right] $$
where ${\nabla {}}_{\textbf {r}}=\frac {\partial {}}{\partial {}x}\vec {i}+\frac {\partial {}}{\partial {}y}\vec {j}$ and $\psi {}\left (\textbf {r},t\right )$ is the exact Born-Oppenheimer (static nuclei) electron wave packet obtained from Eq. (1). The corresponding time-dependent magnetic field B(r,t) is obtained using the following classical Jefimenko’s equation [46]:
$$\begin{aligned}\textbf{B}(\textbf{r},t)=&\frac{{\mu{}}_0}{4\pi{}}\;\;\; \int[\frac{\textbf{j}({\textbf{r}}',t_r)}{{\vert{}\textbf{r}-{\textbf{r}}' \vert{}}^3}+\frac{1}{{\vert{}\textbf{r}-{\textbf{r}}'\vert{}}^2c}\frac{\partial{}\textbf{j}({\textbf{r}}',t_r)}{\partial{}t}]\\ &\times \;\;\;(\textbf{r}-{\textbf{r}}')d^3{\textbf{r}}' \ \end{aligned}$$
where $t_r=t-r/c$ is the retarded time and ${\mu {}}_0=4\pi {}\times {}{10}^{-7}$ NA$^{-2}$ $(6.692\times {}{10}^{-4}$ a.u.). For the static induced magnetic field after the laser pulse, Eq. (6) reduces to the classical Biot-Savart law $B\left ( \textbf {r},t\right )=\frac {\mu _{0}}{4\pi }\int \left [\frac {j\left ( {\textbf {r}}',t\right )\times \left ( \textbf {r}-{\textbf {r}}'\right )}{\left |\textbf {r}-{\textbf {r}}' \right |^{3}} \right ]d^{3}{\textbf {r}}'$.

3. Results and discussions

We study the electron currents and corresponding time-dependent magnetic fields induced by coherent electronic wave packets for symmetric H$_{2}^{+}$ and asymmetric HeH$^{2+}$ molecular ions at equilibrium aligned along the x-axis. Molecular ions are excited by using a CP XUV laser pulse with its electric field vector polarized in the (x, y) plane, propagating along the z axis, as illustrated in Fig. 1. The pulse intensity is $\ I_0 = 1\times {}{10}^{13} \ \mbox {W}/{\mbox {cm}^2}$ ($E_{0} = 1.688 \times 10^{-2}$ a.u.). The research content will follow two designed scenarios. (i) We use XUV laser pulses with photon energy $\hbar \omega$ = $\Delta E$ = $E_{e}-E_{g}$ ( $E_{e}$ and $E_{g}$ represent the eigenenergies of the excited state and ground state) to create coherent excitation and CM of the $1s\sigma _{g}-2p\sigma _{u}$ and $2p\sigma -2p\pi$ electronic state combinations. Where, $1s\sigma _{g}-2p\sigma _{u}$ resonant excitation occurs from the electronic ground state $1s\sigma _{g}$ to the electronic first excited state $2p\sigma _{u}$ for H$_{2}^{+}$. Here, the $2p\sigma -2p\pi$ resonant excitation is considered for HeH$^{2+}$. Since the ground state $1s\sigma$ of HeH$^{2+}$ is mainly localized on the He$^{2+}$, its potential energy curve is repulsive, while the first excited $2p\sigma$ state has a minimum at R = 3.89 a.u., the mean lifetime of $2p\sigma$ state is about 4 ns [47]. Of course, we also compare the results of the HeH$^{2+}$ when the initial states are $2p\sigma$ and $1s\sigma$ respectively (Supplement 1, part 1). (ii) We use the XUV ultra-short laser pulse with photon energy ($\hbar \omega >I_{p}$) to directly ionize molecular ions and explore the CM process. Figure 2(a) shows the diagram of these two scenarios. The corresponding ionization potentials $I_{p}$ for $(1s\sigma _{g})$ and $(2p\sigma _{u})$ are 1.08 a.u. and 0.65 a.u. in the H$_{2}^{+}$ (R = 2 a.u.) [48,49], and that for $1s\sigma$, $2p\sigma$ , $2p\pi$ are 2.25 a.u., 1.03 a.u., 0.74 a.u. in the HeH$^{2+}$ (R = 3.89 a.u.) [5052]. Meanwhile, the densities of initial wavefunction of the selected electronic states for two molecular ions are presented in Fig. 2(b).

 figure: Fig. 1.

Fig. 1. Illustration of the attosecond magnetic field pulses $B\left ( \textbf {r},t\right )$. It is generated in the molecular ion by a right-handed circular attosecond pulses E(t) in the $\left ( x,y\right )$ plane. The corresponding electron currents $j\left ( \textbf {r},t\right )$ are induced in the (x, y) plane and the magnetic field is perpendicular to $j\left ( \textbf {r},t\right )$, along the z axis.

Download Full Size | PDF

 figure: Fig. 2.

Fig. 2. Illustration of the possible photoexcitation pathways to produce electron currents and magnetic fields by CP laser pulses. (a) The left and right column represent the two scenarios designed for HeH$^{2+}$ and H$_{2}^{+}$. (b) The initial wavefunction densities of selected electronic states. The upper row represents the initial $2p\sigma$ excited state and the second excited state $2p\pi$ for HeH$^{2+}$, and the lower row represents the ground state $1s\sigma _{g}$ and the first excited state $2p\sigma _{u}$ for H$_{2}^{+}$.

Download Full Size | PDF

We first consider the scenario (i). For HeH$^{2+}$, the $2p\sigma - 2p\pi$ resonant excitation from the electronic first excited state $2p\sigma$ ($|\psi _{g0}\left ( \textbf {r}\right )\rangle$ with the eigenenergy $E_{g0}$) to the electronic excited $2p\pi$ state ($|\psi _{e0} \left ( \textbf {r}\right )\rangle$ with the eigenenergy $E_{e0}$) is studied. To induce molecular resonant excitation, the wavelength $\lambda$=150 nm ( $\omega$ = $E_{e0}$ - $E_{g0}$ = 0.3 a.u.) of the CP XUV pulses is required. For H$_{2}^{+}$, the $1s\sigma _{g}-2p\sigma _{u}$ resonant excitation from the electronic ground state $1s\sigma _{g}$ ($|\psi _{g1} \left ( \textbf {r}\right )\rangle$ with the eigenenergy $E_{g1}$) to the electronic excited $2p\sigma _{u}$ state ($|\psi _{e1} \left ( \textbf {r}\right )\rangle$ with the eigenenergy $E_{e1}$) is studied. The wavelength with $\lambda$=100 nm ($\omega$= $E_{e1}$ - $E_{g1}$ = 0.455 a.u.) of the CP pulses is used to produce such resonant excitation. Then, a coherent superposition of the two electronic states is created due to a strong charge-resonance excitation for both molecular ions, respectively.

$${\psi{}}_0\left(\textbf{r},t\right)=c_{gn}\left ( t\right )|\psi _{gn}\left ( r\right )\rangle e^{{-}iE_{gn}t/\hbar}+c_{en}\left ( t\right )|\psi _{en}\left ( r\right )\rangle e^{{-}iE_{en}t/\hbar}$$
where, the $c_{gn}\left ( t\right )$ and $c_{en}\left ( t\right )$ (n = 0, 1) are occupation coefficients. Then, the corresponding time-dependent coherent electron density distributions are described by
$$\begin{aligned}\mathcal{A}\left(\textbf{r},t\right)&={\left\vert{}{\psi{}}_0\left(\textbf{r},t\right)\right\vert{}}^2\\ &=|c_{gn}\left ( t\right )\psi _{gn}\left ( \textbf{r}\right )|^{2}+|c_{en}\left ( t\right )\psi _{en}\left ( \textbf{r}\right )|^{2}\\ & +2\left | c_{gn}\left ( t\right )c_{en}\left ( t\right )\right |\psi _{gn}\left ( \textbf{r}\right )\psi _{en}\left ( \textbf{r}\right )cos\left ( \Delta Et\right ) \end{aligned}$$

The coherent electron dynamic is composed of two electronic state densities, $\mathcal {A}^{(gn)}$ = $|c_{gn}\left ( t\right )\psi _{(gn)}\left ( \textbf {r}\right )|^{2}$ and $\mathcal {A}^{(en)}$ = $|c_{(en)}\left ( t\right )\psi _{(en)}\left ( \textbf {r}\right )|^{2}$, which are insensitive to the time. Their interfering superposition can be expressed: $\mathcal {A}^{(g,e)}$ = $2\left | c_{(gn)}\left ( t\right )c_{(en)}\left ( t\right )\right |\psi _{(gn)}\left ( \textbf {r}\right )\psi _{(en)}\left ( \textbf {r}\right )cos\left ( \Delta Et\right )$. Whereas the coherent superposition term $2\left | c_{(gn)}\left ( t\right )c_{(en)}\left ( t\right )\right |\psi _{(gn)}\left ( \textbf {r}\right )\psi _{(en)}\left ( \textbf {r}\right )cos\left ( \Delta Et\right )$ is determined by the time t.

The time-dependent superposition term $\mathcal {A}^{\left ( g,e\right )}\left (\textbf {r},t \right )$ describes the attosecond electronic coherence with the oscillation period $\Delta \tau = 2 \pi /\Delta E = 506.8$ as, where $\Delta E$ = $E_{e0}$ - $E_{g0}$ = 0.3 a.u. for HeH$^{2+}$. The coherent electron density probability $\mathcal {A}\left ( \textbf {r},t\right )$ oscillates between the two protons with time t (Fig. 3(a)). At t = 5 $\tau$, the coherent electron density is mainly localized at H$^{+}$, while, as time increases, the electron is gradually localized at ion He$^{2+}$. That is, at 5.0 $\tau$ $\leq$ t $\leq$ 5.25 $\tau$, the electron migration from H$^{+}$ to He$^{2+}$. As the time t increases further, the electron density come back to the H$^{+}$, finally, the electron migrates from H$^{+}$ back to He$^{2+}$ at t = 6 $\tau$. Therefore, the CM presents a periodic discipline i.e., the electron oscillates between the H$^{+}$ and He$^{2+}$, and the time is exactly one optical cycle.

 figure: Fig. 3.

Fig. 3. Electronic dynamics in CM for HeH$^{2+}$ at equilibrium R = 3.89 a.u. from the initial $2p\sigma$ excited state. (a) and (c) show the electronic density probabilities $\mathcal {A}\left (\textbf {r},t \right )$ at five different times (top row). (b) and (d) are the electronic current densities $j\left ( \textbf {r},t\right )$ at five different times. The white arrows in (b) and (d) label the directions of the electronic currents. All these are obtained by using right-handed CP pulses ( $\xi$= −1) with $\lambda$ = 150 nm in (a-b) and 30 nm in (c-d). Arbitrary units of distributions are used.

Download Full Size | PDF

The electronic density probabilities intuitively reflect the distribution of charges in space, while the electronic current can show the distribution of electron velocity in space. To visualize the coherent electronic density, we present the electronic current distribution at different times (Fig. 3(b)). At t = 5.0 $\tau$, the electronic current is localized at H$^{+}$, while at time t = 5.25 $\tau$, the current density is completely localized at He$^{2+}$. The electron current directions for two moments above are both to the left, then, the direction starts to point to the right at t = 5.5 $\tau$, and finally at t = 6 $\tau$, the direction migrate to the left again. The currents alternate between He$^{2+}$ and H$^{+}$ along the molecular x axis with the change of time $\tau$. Actually, under the resonance excitation, the coherent electronic density cannot be ignored, and it is accompanied by the periodic oscillation at time $\tau$, since $\mathcal {A}^{\left ( g,e\right )}\left (\textbf {r},t \right )\sim cos(\Delta Et)$. Meanwhile, the superposition state is the non-stationary state. Therefore, the electronic density distributions oscillates back and forth between the two protons. We can see that the process of CM between the two protons is significant by resonance excitation between $2p\sigma$ and $2p\pi$ electronic states, which can better explain the enhancement of the CM efficiency.

Next, we discuss the scenario (ii), which uses ultrashort pulses with $\lambda$ = 30 nm ($\omega$ = 1.5 a.u., $\omega >I_{p}$) to directly ionize initial $2p\sigma$ state of HeH$^{2+}$. The electronic density probability and the electron current density distribution over time are calculated, as shown in Fig. 3(c-d). It can be seen that the electronic density probability does not change with time, and the CM between the He$^{2+}$ and H$^{+}$ is not significant, which result in the low CM efficiency. And from the perspective of the direction of electron currents, the electron no longer alternately migrate between two protons over time, while exhibits a periodic and clockwise rotation with an optical cycle of 506 as. That is, at time t = 5.0 $\tau$, 5.25 $\tau$, 5.5 $\tau$, 5.75 $\tau$ and 6.0 $\tau$, the electron current directions are downward, left, upward, right and downward, respectively. Since at higher laser frequency $\omega >I_{p}$, the molecular ion is ionized after absorbing one photon, the ionized electrons with relatively small kinetic energies move away directly with the velocity perpendicular to the molecular R axis, that is, the y direction, due to a nonzero drift velocity in the y direction [53]. In fact, the induced electron currents (Fig. 3(d)) result from the electron wave packets in the continuum state generated by the driving CP pulses, and the coherence of electronic wave packet is very weak and almost negligible. Therefore, the electron wave packets with near zero initial velocities move in the (x, y) plane as a free electron, following the driving circularly polarized pulses. It explains the electron currents rotating clockwise in an optical cycle, and the corresponding CM efficiency decreases.

In order to examine the purely repulsive of the potential energy in the ground state $1s\sigma$ of HeH$^{2+}$, we also show the electronic density probability and electronic currents distribution at different times, as shown in Supplement 1, part 1. It was found that the generation of electronic current and corresponding magnetic field induced by resonance excitation $1s\sigma -2p\sigma$ are both weaker than those induced by resonance excitation $2p\sigma -2p\pi$. Interestingly, the electron current induced by former is sensitive to the helicity of CP laser pulses, which is similar to the scenarios (ii) from the initial $2p\sigma$ state.

Then, we turn our attention to the symmetric molecular ion H$_{2}^{+}$. Figure 4(a-b) present the electron density probabilities $\mathcal {A}\left (\textbf {r},t \right )$ and the density distribution $j\left ( \textbf {r},t\right )$ of electron currents for a resonant excitation ($1s\sigma _{g}-2p\sigma _{u}$) by using right-handed CP laser pulses ($\xi$ = −1) with $\lambda$ = 100 nm. It can be seen that both of them show the oscillating behavior similar to that in resonant excitation ($2p\sigma -2p\pi$ of HeH$^{2+}$). Meanwhile, the corresponding $\mathcal {A}\left (\textbf {r},t \right )$ and $j\left ( \textbf {r},t\right )$ under the case of direct ionization are also calculated by using same ultrashort pulses with $\lambda$ = 30 nm (Fig. 4(c-d)). We found that the results obtained were consistent with that of direct ionization for HeH$^{2+}$. Thus, it can be known that no matter in the scenario (i) or (ii), the two molecular ions have similar disciplines. That is, the electron current induced by the alternating CM between two parts, the CM efficiency is high, and it is not sensitive to the helicity of circular polarization in the scenario (i). However, in the scenario (ii), the CM efficiency is low, and it is extremely sensitive to the helicities of the laser pulses.

 figure: Fig. 4.

Fig. 4. Electronic dynamics in CM for H$_{2}^{+}$ at equilibrium R = 2 a.u. from the initial $1s\sigma _{g}$ ground state. (a) and (c) show the electronic density probabilities $\mathcal {A}\left (\textbf {r},t \right )$ at five different times (top row). (b) and (d) are the electronic current densities $j\left ( \textbf {r},t\right )$ at five different times. The white arrows in (b) and (d) label the directions of electronic currents. All these are obtained by using right-handed CP pules ($\xi$ = −1) with $\lambda$ = 100 nm in (a-b) and 30 nm in (c-d). Arbitrary units of distributions are used.

Download Full Size | PDF

In order to better observe above phenomena of the H$_{2}^{+}$ and HeH$^{2+}$, we display the axial electronic current (i.e., $j_{x}\left ( x,t\right )$) as a function of position x and time t (in units of period 5 $\tau$ $\sim$ 6 $\tau$ = 506.8 as). Here, we operate on Eq. (5) with $j_{x}\left ( x,t\right )=\int _{-\infty }^{+\infty }dyj_{x}\left ( \textbf {r},t\right )$ (i.e., by integrating over the coordinate y), where $j\left (\textbf {r},t \right )=j_{x}\left ( \textbf {r},t\right )\hat {e}_{x}+j_{y}\left ( \textbf {r},t\right )\hat {e}_{y}$, and $\hat {e}_{x}$, $\hat {e}_{y}$ are the unit vector along x and y directions. Taking scenario (i) as an example, we display the results of axial electronic currents for the $2p\sigma -2p\pi$ resonant excitation of HeH$^{2+}$ in Fig. 5(a) and for the $1s\sigma _{g}-2p\sigma _{u}$ resonant excitation of H$_{2}^{+}$ in Fig. 5(c). It can be seen that the direction of axial electron currents as a function of time remains the same between 5 $\tau$ and 5.25 $\tau$, while the direction changes when the time evolves from 5.25 $\tau$ to 5.5 $\tau$, then at time from 5.5 to 5.75 $\tau$, the direction remains unchanged. Finally, the reverse of direction occurs between t = 5.75 $\tau$ and t = 6 $\tau$. These phenomena can more clearly reflect the enhanced resonance. And it is consistent with the electron current direction shown in Fig. 3(b)-Fig. 4(b) (white arrow). In addition, the result under the scenario (ii) is also calculated (Fig. 5(b-d)). We can see that only at t= 5.5 $\tau$, the direction of axial electron current reverses, which is consitent with Fig. 3(d) and Fig. 4(d).

 figure: Fig. 5.

Fig. 5. Profiles of axial electronic currents $j_{x}\left ( x,t\right )$) as a function of position x and time t (in units of period $\tau$ = 350 as (refers to the optical period of the H$_{2}^{+}$ in the case of resonant excitation)). (a-b) The $j_{x}\left ( x,t\right )$ for resonant excitation and direct ionization of HeH$^{2+}$. (c-d) The $j_{x}\left ( x,t\right )$ for resonant excitation and direct ionization of H$_{2}^{+}$. Dotted white lines indicate positions of protons. Scale of magnitude and direction are shown at right.

Download Full Size | PDF

For further quantifying the CM process in the above two scenarios, such as the level of migration efficiency, we define the axial electronic yield along the molecular axis. That is, a change in the probability of finding an electron from an area of one nucleus position to another nucleus position during a period of 5$\tau$ $\leq$ t $\leq$ 6$\tau$. It can be expressed as $y_{x}=\int _{5\tau }^{6\tau }dtj\left ( x,t\right )$, where $j\left ( x,t\right )=\int _{-\infty }^{+\infty }dyj\left (\textbf {r},t \right )$ . We exhibit the axial electronic yield in Fig. 6.

 figure: Fig. 6.

Fig. 6. Axial electronic yield $y_{x}$ for two molecular ions. (a) The $y_{x}$ for resonant excitation of $2p\sigma -2p\pi$ (blue solid line) and $1s\sigma -2p\sigma$ (green solid line), direct ionization (orange solid line) in HeH$^{2+}$. (b) The $y_{x}$ for resonant excitation of $1s\sigma _{g}-2p\sigma _{u}$ (blue solid line ) and direct ionization (orange solid line) in H$_{2}^{+}$. The black arrows denote positions of protons.

Download Full Size | PDF

For HeH$^{2+}$, the maximum electronic yield (0.266 a.u.) of $2p\sigma -2p\pi$ resonant excitation is much higher than that (0.008 a.u.) of $1s\sigma -2p\sigma$ resonant excitation. They are both higher than that (0.004 a.u.) of direct ionization. These indicate that the maximum electronic yield of resonant excitation is higher that of direct ionization. This feature also applies to H$_{2}^{+}$, for which the yields of $1s\sigma _{g}-2p\sigma _{u}$ resonant excitation and direct ionization are 0.436 a.u. and 0.006 a.u.. Additionally, the electronic yield of HeH$^{2+}$ is different from that of H$_{2}^{+}$. On the one hand, the $y_{x}$ of H$_{2}^{+}$ has an maximum at x = 0, but for HeH$^{2+}$, it does not appear. This is because the spatial reflection symmetry or parity of the different nuclear charges in HeH$^{2+}$ with respect to the midpoint of the two nuclei is no longer a conserved quantity. On the other hand, the yields of H$_{2}^{+}$ is almost twice higher than that of HeH$^{2+}$ in the scenario (i), which also shows that this symmetric molecular ion has a better CM efficiency and is more likely to induce stronger electronic currents and magnetic fields than asymmetric HeH$^{2+}$ molecular ion.

Based on the above results, we can reasonably speculate that electronic currents induced by the excitation resonance ($1s\sigma _{g}-2p\sigma _{u}$ or $2p\sigma -2p\pi$ ) will be stronger than that induced by direct ionization. To further illustrate the physical mechanism about this phenomenon, the electronic currents are calculated, based on the principle of the continuity of the electronic current in quantum mechanics, that is, the probability of the measured electron in any cross-sections generated through the center point is equal [44]. Therefore, we can examine the dependence of the electronic currents on time by integrating $j\left ( \textbf {r},t\right )$ over the line along the y axis through the section -$\infty$ to 0. It is expressed by [44]

$$J\left ( t\right )=\int_{-\infty }^{0}j\left ( 0,y,t\right )\cdot e_{x}dy$$
Where $j\left (\textbf {r},t \right )=j_{x}\left ( \textbf {r},t\right )\hat {e}_{x}+j_{y}\left ( \textbf {r},t\right )\hat {e}_{y}$, $\hat {e}_{x}$, $\hat {e}_{y}$ are the unit vectors along x and y directions. $J\left (t \right )$ is presented in Fig. 7(a) and (c). The duration time of electronic current in scenario (ii) is T= 10 $\tau$. Taking H$_{2}^{+}$ as an example, we can see that the electronic current in scenario (ii) is almost zero, while that in scenario (i) can reach the maximum value of 0.056 a.u., indicating that compared with $1s\sigma _{g}-2p\sigma _{u}$ resonant excitation, the electronic current induced by direct ionization is particularly weak, which is consistent with the low CM efficiency shown in the electron density distribution (Fig. 4(c)). This is due to the fact that corresponding electron in continuum state move away and then the ionization rates are also weak, thus weaker electron current are obtained. This feature also applies to HeH$^{2+}$. Actually, the time-dependent electronic current can be viewed as a time-dependent magnetic dipole, which implies that an internal ultrafast-varying magnetic field can be generated in the molecules. Figure 7(b) and (d) display the magnetic field $B\left ( \textbf {r},t\right )$, at the molecular center in H$_{2}^{+}$ and the nuclei r = $+R_{2}$ ($x_H{^{+}}$) in HeH$^{2+}$, induced by the electronic currents $j\left ( \textbf {r},t\right )$ on the polarization plane (x, y) for both molecular ions. Since HeH$^{2+}$ is an asymmetrical molecular ion, the electron probability of the initial $2p\sigma$ is mainly on the H$^{+}$. For both H$_{2}^{+}$ and HeH$^{2+}$, the induced time-dependent magnetic field excited by ($1s\sigma _{g}-2p\sigma _{u}$ or $2p\sigma -2p\pi$) resonance in scenario (i) is much stronger than the magnetic field in scenario (ii), which is almost zero. At the same time, compared with the HeH$^{2+}$, the electronic current and magnetic field of H$_{2}^{+}$ are still stronger than that of the HeH$^{2+}$. For example, in the case of resonance excitation, the strongest current can reach 0.056 a.u. for H$_{2}^{+}$ and the one for HeH$^{2+}$ can reach 0.038 a.u.. The corresponding magnetic field for H$_{2}^{+}$ is also the strongest, which can reach 1.2 Tesla, while that for HeH$^{2+}$ can only reach 1.05 Tesla. This is consistent with the calculated electronic yield of the CM mentioned above. Additionally, we also calculated the corresponding magnetic fields at other locations of H$_{2}^{+}$ and HeH$^{2+}$ (Supplement 1, part 2) and of $1s\sigma -2p\sigma$ resonant excitation for HeH$^{2+}$ (Supplement 1, part 3). In contrast, we can see that the magnetic field presented in Fig. 7 is stronger and more regular.

 figure: Fig. 7.

Fig. 7. The time-dependent electronic currents $J\left (t \right )$ and magnetic fields for two molecular ions. (a) and (c) show the $J\left (t \right )$ passing through the section -$\infty$ to 0 along y axis in equation (9) of HeH$^{2+}$ and H$_{2}^{+}$ respectively. The required wavelength of resonance excitation, intensity of the laser pulses are marked in the figure and at intensity $\ I_0=1\times {}{10}^{13} \ \mbox {W}/{\mbox {cm}^2}$. (b) and (d) represent magnetic fields $B\left (\textbf {r}=+R_{2},t \right )$ for HeH$^{2+}$, and $B\left (\textbf {r}=0,t \right )$ for H$_{2}^{+}$. Both magnetic fields $B\left (\textbf {r}, t\right )$ are perpendicular to the laser (x, y) polarization plane (shown in Fig. 1) by using laser pulses at $\lambda$ = 150 nm ($\omega$= 0.3 a.u.) (pink solid line) and $\lambda$ = 30 nm ($\omega$ = 1.5 a.u.) (black solid line) for HeH$^{2+}$, and at $\lambda$= 100 nm ($\omega$ = 0.455 a.u.) (pink solid line) and $\lambda$=30 nm ($\omega$=1.5 a.u.) (black solid line) for H$_{2}^{+}$. The durations are t=10 $\tau$, where $\tau$=2$\pi$/$\omega$ (i.e., 1 o.c. is 506.8 as for $\lambda$=150 nm) is fixed.

Download Full Size | PDF

4. Conclusion

In summary, we focused on the electronic coherence dynamics and explored the ultrafast oscillating electron current and corresponding magnetic field generated from CM for aligned symmetric molecular ion (H$_{2}^{+}$) and asymmetric molecular ion (HeH$^{2+}$) by numerically solving TDSE. The two scenarios were considered for HeH$^{2+}$ and H$_{2}^{+}$, including (i) resonance excitation and (ii) direct ionization. In scenario (i), the electronic density distribution and density distribution of electronic current exhibit the oscillating behavior as a function of time, which illustrates the strong electron coherence. Moreover, it can be known that the process of CM is very significant by resonance excitation ($1s\sigma _{g}-2p\sigma _{u}$ for H$_{2}^{+}$ or $2p\sigma -2p\pi$ for HeH$^{2+}$), which can better reflect the enhancement of the CM efficiency. And electronic current is not sensitive to the helicity of circular polarization of the laser pulses. In addition, the resonance excitation of HeH$^{2+}$ from the ground electronic state $1s\sigma$ was also studied. It was found that a weak magnetic field is generated and the electron current is sensitive to the helicity of circular polarization. However, in the scenario (ii), the CM is not significant, and the CM efficiency is low. Weaker magnetic fields are also obtained. By comparing the two scenarios, we found that in scenario (i), the electron current and magnetic field of resonance excitation for HeH$^{2+}$ from the initial first excited $2p\sigma$ electronic state and H$_{2}^{+}$ from the ground electronic state $1s\sigma _{g}$ are much stronger than that of ionization for HeH$^{2+}$ and H$_{2}^{+}$ in scenario (ii). Moreover, the magnetic field generated by H$_{2}^{+}$ is stronger than that by HeH$^{2+}$ through scenario (i). The generation of the ultrafast oscillating magnetic field provides possibility to detect and manipulate the ultrafast magnetic and chiral dynamics, and study electron dynamics in ultrafast magneto-optics. We expect the magnetic source with high oscillation frequency will be applied in other areas, such as the high-speed information processing and storage.

Funding

National Natural Science Foundation of China (12074146).

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available currently but may be obtained from the authors upon reasonable request.

Supplemental document

See Supplement 1 for supporting content.

References

1. F. Krausz and M. Ivanov, “Attosecond physics,” Rev. Mod. Phys. 81(1), 163–234 (2009). [CrossRef]  

2. Z. H. Chang and P. B. Corkum, “Attosecond photon sources: the first decade and beyond,” J. Opt. Soc. Am. B 27(11), B9–B17 (2010). [CrossRef]  

3. P. Hockett, C. Z. Bisgaard, O. J. Clarkin, and A. Stolow, “Time-resolved imaging of purely valence-electron dynamics during a chemical reaction,” Nat. Phys. 7(8), 612–615 (2011). [CrossRef]  

4. T. Okino, Y. Furukawa, Y. Nabekawa, S. Miyabe, A. A. Eilanlou, E. Takahashi, K. Yamanouchi, and K. Midorikawa, “Direct observation of an attosecond electron wave packet in a nitrogen molecule,” Sci. Adv. 1(8), e1500356 (2015). [CrossRef]  

5. F. Lépine, M. Y. Ivanov, and M. J. J. Vrakking, “Attosecond molecular dynamics: fact or fiction?” Nat. Photonics 8(3), 195–204 (2014). [CrossRef]  

6. K. Ramasesha, S. R. Leone, and D. M. Neumark, “Real-time probing of electron dynamics using attosecond time-resolved spectroscopy,” Annu. Rev. Phys. Chem. 67(1), 41–63 (2016). [CrossRef]  

7. T. Gaumnitz, A. Jain, Y. Pertot, M. Huppert, I. Jordan, F. Ardana-Lamas, and H. J. Wörner, “Streaking of 43-attosecond soft-x-ray pulses generated by a passively cep-stable mid-infrared driver,” Opt. Express 25(22), 27506–27518 (2017). [CrossRef]  

8. L. Plaja, R. Torres, and A. Zaïr, Attosecond Physics: Attosecond Measurements and Control of Physical Systems (Springer, Berlin, 2013).

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

10. T. Fan, P. Grychtol, R. Knut, C. H. Garcia, D. D. Hickstein, D. Zusin, C. Gentry, F. J. Dollar, C. A. Mancuso, C. W. Hogle, O. Kfir, D. Legut, K. Carva, J. L. Ellis, K. M. Dorney, C. Chen, O. G. Shpyrko, E. E. Fullerton, O. Cohen, P. M. Oppeneer, D. B. Milosevic, A. Becker, A. A. Jaron-Becker, T. Popmintchev, M. M. Murnane, and H. C. Kapteyn, “Bright circularly polarized soft x-ray high harmonics for x-ray magnetic circular dichroism,” Proc. Natl. Acad. Sci. U. S. A. 112(46), 14206–14211 (2015). [CrossRef]  

11. O. Kfir, P. Grychtol, E. Turgut, R. Knut, D. Zusin, D. Popmintchev, T. Popmintchev, H. Nembach, J. M. Shaw, A. Fleischer, H. Kapteyn, M. Murnane, and O. Cohen, “Generation of bright phase-matched circularly-polarized extreme ultraviolet high harmonics,” Nat. Photonics 9(2), 99–105 (2015). [CrossRef]  

12. O. Kfir, P. Grychtol, E. Turgut, R. Knut, D. Zusin, A. Fleischer, E. Bordo, T. Fan, D. Popmintchev, T. Popmintchev, H. Kapteyn, M. Murnane, and O. Cohen, “Helicity-selective phase-matching and quasi-phase matching of circularly polarized high-order harmonics: towards chiral attosecond pulses,” J. Phys. B: At. Mol. Opt. Phys. 49(12), 123501 (2016). [CrossRef]  

13. X. F. Zhang, L. Li, X. S. Zhu, X. Liu, Q. B. Zhang, P. F. Lan, and P. X. Lu, “Helicity reversion in high-order-harmonic generation driven by bichromatic counter-rotating circularly polarized laser fields,” Phys. Rev. A 94(5), 053408 (2016). [CrossRef]  

14. D. B. Milosevic, “Possibility of introducing spin into attoscience with spin-polarized electrons produced by a bichromatic circularly polarized laser field,” Phys. Rev. A 93(5), 051402 (2016). [CrossRef]  

15. A. D. Bandrauk, F. Mauger, and K. J. Yuan, “Circularly polarized harmonic generation by intense bicircular laser pulses: electron recollision dynamics and frequency dependent helicity,” J. Phys. B: At. Mol. Opt. Phys. 49(23), 23LT01 (2016). [CrossRef]  

16. D. M. Reich and L. B. Madsen, “Rotating-frame perspective on high-order-harmonic generation of circularly polarized light,” Phys. Rev. A 93(4), 043411 (2016). [CrossRef]  

17. R. Weinkauf, P. Schanen, S. D. Yang, S. D. Soukara, and E. W. Schlag, “Elementary processes in peptides: Electron mobility and dissociation in peptide cations in the gas phase,” J. Phys. Chem. 99(28), 11255–11265 (1995). [CrossRef]  

18. L. S. Cederbaum and J. Zobeley, “Ultrafast charge migration by electron correlation,” Chem. Phys. Lett. 307(3-4), 205–210 (1999). [CrossRef]  

19. I. Barth and J. Manz, “Periodic electron circulation induced by circularly polarized laser pulses: Quantum model simulations for mg porphyrin,” Angew. Chem. Int. Ed. 45(18), 2962–2965 (2006). [CrossRef]  

20. H. Mineo, M. Yamaki, Y. Teranishi, M. Hayashi, S. H. Lin, and Y. Fujimura, “Quantum switching of π-electron rotations in a nonplanar chiral molecule by using linearly polarized uv laser pulses,” J. Am. Chem. Soc. 134(35), 14279–14282 (2012). [CrossRef]  

21. B. Mignolet, R. D. Levine, and F. Remacle, “Charge migration in the bifunctional penna cation induced and probed by ultrafast ionization a dynamical study,” J. Phys. B 47(12), 124011 (2014). [CrossRef]  

22. V. Despré, A. Marciniak, V. Loriot, M. C. E. Galbraith, A. Rouzée, M. J. J. Vrakking, F. Lépine, and A. I. Kuleff, “Attosecond hole migration in benzene molecules surviving nuclear motion,” J. Phys. Chem. Lett. 6(3), 426–431 (2015). [CrossRef]  

23. G. Hermann, C. Liu, J. Manz, B. Paulus, J. F. Prez-Torres, V. Pohl, and J. C. Tremblay, “Multidirectional angular electronic flux during adiabatic attosecond charge migration in excited benzene,” J. Phys. Chem. A 120(27), 5360–5369 (2016). [CrossRef]  

24. K. J. Yuan and A. D. Bandrauk, “Monitoring coherent electron wave packet excitation dynamics by two-colorattosecond laser pulses,” J. Chem. Phys. 145(19), 194304 (2016). [CrossRef]  

25. F. Remacle and R. D. Levine, “An electronic time scale in chemistry,” Proc. Natl. Acad. Sci. U. S. A. 103(18), 6793–6798 (2006). [CrossRef]  

26. H. C. Shao and A. F. Starace, “Imaging population transfer in atoms with ultrafast electron pulses,” Phys. Rev. A 94(3), 030702 (2016). [CrossRef]  

27. D. Jia, J. Manz, B. Paulus, V. Pohl, J. C. Tremblay, and Y. Yang, “Quantum control of electronic fluxes during adiabatic attosecond charge migration in degenerate superposition states of benzene,” Chem. Phys. 482, 146–159 (2017). [CrossRef]  

28. G. Hermann, C. Liu, J. Manz, B. Paulus, V. Pohl, and J. C. Tremblay, “Attosecond angular flux of partial charges on the carbon atoms of benzene in non-aromatic excited state,” Chem. Phys. Lett. 683, 553–558 (2017). [CrossRef]  

29. D. J. Diestler, G. Hermann, and J. Manz, “Charge migration in eyring, walter and kimball’s 1944 model of the electronically excited hydrogen-molecule ion,” J. Phys. Chem. A 121(28), 5332–5340 (2017). [CrossRef]  

30. W. Li, A. A. Jaron-Becker, C. W. Hogle, V. Sharma, X. Zhou, A. Becker, H. C. Kapteyn, and M. M. Murnane, “Visualizing electron rearrangement in space and time during the transition from a molecule to atoms,” Proc. Natl. Acad. Sci. U. S. A. 107(47), 20219–20222 (2010). [CrossRef]  

31. P. M. Kraus, B. Mignolet, D. Baykusheva, A. Rupenyan, L. Horný, E. F. Penka, G. Grassi, O. I. Tolstikhin, J. Schneider, F. Jensen, L. B. Madsen, A. D. Bandrauk, F. Remacle, and H. J. Wörner, “Measurement and laser control of attosecond charge migration in ionized iodoacetylene,” Science 350(6262), 790–795 (2015). [CrossRef]  

32. E. Räsänen, A. Castro, J. Werschnik, A. Rubio, and E. K. U. Gross, “Optimal control of quantum rings by terahertz laser pulses,” Phys. Rev. Lett. 98(15), 157404 (2007). [CrossRef]  

33. M. Kanno, H. Kono, Y. Fujimura, and S. H. Lin, “Nonadiabatic response model of laser-induced ultrafast π-electron rotations in chiral aromatic molecules,” Phys. Rev. Lett. 104(10), 108302 (2010). [CrossRef]  

34. I. Barth, J. Manz, Y. Shigeta, and K. Yagi, “Unidirectional electronic ring current driven by a few cycle circularly polarized laser pulse: Quantum model simulations for mg-porphyrin,” J. Am. Chem. Soc. 128(21), 7043–7049 (2006). [CrossRef]  

35. I. Barth and J. Manz, “Electric ring currents in atomic orbitals and magnetic fields induced by short intense circularly polarizedπlaser pulses,” Phys. Rev. A 75(1), 012510 (2007). [CrossRef]  

36. I. Barth, L. Serrano-Andrés, and T. Seideman, “Nonadiabatic orientation, toroidal current, and induced magnetic field in beo molecules,” J. Chem. Phys. 129(16), 164303 (2008). [CrossRef]  

37. K. Nobusada and Y. Kazuhiro, “Photoinduced electric currents in ring-shaped molecules by circularly polarized laser pulses,” Phys. Rev. A 75(3), 032518 (2007). [CrossRef]  

38. K. J. Yuan and A. D. Bandrauk, “Attosecond-magnetic-field-pulse generation by intense few-cycle circularly polarized uv laser pulses,” Phys. Rev. A 88(1), 013417 (2013). [CrossRef]  

39. A. D. Bandrauk and K. J. Yuan, “Circularly polarised attosecond pulses: generation and applications,” Mol. Phys. 114(3-4), 1–12 (2015). [CrossRef]  

40. K. J. Yuan and A. D. Bandrauk, “Attosecond-magnetic-field-pulse generation by electronic currents in bichromatic circularly polarized uv laser fields,” Phys. Rev. A 92(6), 063401 (2015). [CrossRef]  

41. J. Guo, K. J. Yuan, H. Z. Lu, and A. D. Bandrauk, “Spatiotemporal evolution of ultrafast magnetic-field generation in molecules with intense bichromatic circularly polarized uv laser pulses,” Phys. Rev. A 99(5), 053416 (2019). [CrossRef]  

42. K. J. Yuan, J. Guo, and A. D. Bandrauk, “Generation of ultrafast magnetic fields from molecular coherent electron currents,” Phys. Rev. A 98(4), 043410 (2018). [CrossRef]  

43. A. D. Bandrauk, J. Guo, and K. J. Yuan, “Circularly polarized attosecond pulse generation and applications to ultrafast magnetism,” Phys. Rev. A 19(12), 124016 (2017). [CrossRef]  

44. X. F. Zhang, X. S. Zhu, D. Wang, L. Li, X. Liu, Q. Liao, P. F. Lan, and P. X. Lu, “Ultrafast oscillating-magnetic-field generation based on electronic-current dynamics,” Phys. Rev. A 99(1), 013414 (2019). [CrossRef]  

45. M. D. Feit and J. A. F. Jr, “Solution of the schrödinger equation by a spectral method ii: Vibrational energy levels of triatomic molecules,” J. Chem. Phys. 78(1), 301–308 (1983). [CrossRef]  

46. O. D. Jefimenko, Electricity and Magnetism: An Introduction to the Theory of Electric and Magnetic Fields, 2nd ed (Electret Scientific, 1989).

47. I. B. Itzhak, I. Gertner, O. Heber, and B. Rosner, “Experimental evidence for the existence of the 2 bound state of HeH2 + and its decay mechanism,” Phys. Rev. Lett. 71(9), 1347–1350 (1993). [CrossRef]  

48. K. J. Yuan, S. Chelkowski, and A. D. Bandrauk, “Molecular photoelectron angular distributions with intense attosecond circularly polarized UV laser pulses,” Chem. Phys. Lett. 592, 334–340 (2014). [CrossRef]  

49. H. Wind, “Electron energy for H2+ in the ground state,” J. Chem. Phys. 42(7), 2371–2373 (1965). [CrossRef]  

50. X. X. Guan, E. B. Secor, R. C. DuToit, and K. Bartschat, “Diffraction patterns in the ionization of the heteronuclear HeH2 + ion by attosecond x-ray radiation,” Phys. Rev. A 86(5), 053425 (2012). [CrossRef]  

51. X. B. Bian and A. D. Bandrauk, “Phase control of multichannel molecular high-order harmonic generation by the asymmetric diatomic molecule HeH2 + in two-color laser fields,” Phys. Rev. A 83(2), 023414 (2011). [CrossRef]  

52. J. A. Campos, D. L. Nascimento, D. T. Cavalcante, A. L. A. Fonseca, and A. O. C. Nunes, “Determination of electronic energy levels for the heteromolecular ions HeH2 +, LiH3 +, and BeH4 + from the hamilton-jacobi equation,” Int. J. Quantum Chem. 106(13), 2587–2596 (2006). [CrossRef]  

53. P. B. Corkum, N. H. Burnett, and F. Brunel, “Above-threshold ionization in the long-wavelength limit,” Phys. Rev. Lett. 62(11), 1259–1262 (1989). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       Supporting Information

Data availability

Data underlying the results presented in this paper are not publicly available currently but may be obtained from the authors upon reasonable request.

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. Illustration of the attosecond magnetic field pulses $B\left ( \textbf {r},t\right )$. It is generated in the molecular ion by a right-handed circular attosecond pulses E(t) in the $\left ( x,y\right )$ plane. The corresponding electron currents $j\left ( \textbf {r},t\right )$ are induced in the (x, y) plane and the magnetic field is perpendicular to $j\left ( \textbf {r},t\right )$, along the z axis.
Fig. 2.
Fig. 2. Illustration of the possible photoexcitation pathways to produce electron currents and magnetic fields by CP laser pulses. (a) The left and right column represent the two scenarios designed for HeH$^{2+}$ and H$_{2}^{+}$. (b) The initial wavefunction densities of selected electronic states. The upper row represents the initial $2p\sigma$ excited state and the second excited state $2p\pi$ for HeH$^{2+}$, and the lower row represents the ground state $1s\sigma _{g}$ and the first excited state $2p\sigma _{u}$ for H$_{2}^{+}$.
Fig. 3.
Fig. 3. Electronic dynamics in CM for HeH$^{2+}$ at equilibrium R = 3.89 a.u. from the initial $2p\sigma$ excited state. (a) and (c) show the electronic density probabilities $\mathcal {A}\left (\textbf {r},t \right )$ at five different times (top row). (b) and (d) are the electronic current densities $j\left ( \textbf {r},t\right )$ at five different times. The white arrows in (b) and (d) label the directions of the electronic currents. All these are obtained by using right-handed CP pulses ( $\xi$= −1) with $\lambda$ = 150 nm in (a-b) and 30 nm in (c-d). Arbitrary units of distributions are used.
Fig. 4.
Fig. 4. Electronic dynamics in CM for H$_{2}^{+}$ at equilibrium R = 2 a.u. from the initial $1s\sigma _{g}$ ground state. (a) and (c) show the electronic density probabilities $\mathcal {A}\left (\textbf {r},t \right )$ at five different times (top row). (b) and (d) are the electronic current densities $j\left ( \textbf {r},t\right )$ at five different times. The white arrows in (b) and (d) label the directions of electronic currents. All these are obtained by using right-handed CP pules ($\xi$ = −1) with $\lambda$ = 100 nm in (a-b) and 30 nm in (c-d). Arbitrary units of distributions are used.
Fig. 5.
Fig. 5. Profiles of axial electronic currents $j_{x}\left ( x,t\right )$) as a function of position x and time t (in units of period $\tau$ = 350 as (refers to the optical period of the H$_{2}^{+}$ in the case of resonant excitation)). (a-b) The $j_{x}\left ( x,t\right )$ for resonant excitation and direct ionization of HeH$^{2+}$. (c-d) The $j_{x}\left ( x,t\right )$ for resonant excitation and direct ionization of H$_{2}^{+}$. Dotted white lines indicate positions of protons. Scale of magnitude and direction are shown at right.
Fig. 6.
Fig. 6. Axial electronic yield $y_{x}$ for two molecular ions. (a) The $y_{x}$ for resonant excitation of $2p\sigma -2p\pi$ (blue solid line) and $1s\sigma -2p\sigma$ (green solid line), direct ionization (orange solid line) in HeH$^{2+}$. (b) The $y_{x}$ for resonant excitation of $1s\sigma _{g}-2p\sigma _{u}$ (blue solid line ) and direct ionization (orange solid line) in H$_{2}^{+}$. The black arrows denote positions of protons.
Fig. 7.
Fig. 7. The time-dependent electronic currents $J\left (t \right )$ and magnetic fields for two molecular ions. (a) and (c) show the $J\left (t \right )$ passing through the section -$\infty$ to 0 along y axis in equation (9) of HeH$^{2+}$ and H$_{2}^{+}$ respectively. The required wavelength of resonance excitation, intensity of the laser pulses are marked in the figure and at intensity $\ I_0=1\times {}{10}^{13} \ \mbox {W}/{\mbox {cm}^2}$. (b) and (d) represent magnetic fields $B\left (\textbf {r}=+R_{2},t \right )$ for HeH$^{2+}$, and $B\left (\textbf {r}=0,t \right )$ for H$_{2}^{+}$. Both magnetic fields $B\left (\textbf {r}, t\right )$ are perpendicular to the laser (x, y) polarization plane (shown in Fig. 1) by using laser pulses at $\lambda$ = 150 nm ($\omega$= 0.3 a.u.) (pink solid line) and $\lambda$ = 30 nm ($\omega$ = 1.5 a.u.) (black solid line) for HeH$^{2+}$, and at $\lambda$= 100 nm ($\omega$ = 0.455 a.u.) (pink solid line) and $\lambda$=30 nm ($\omega$=1.5 a.u.) (black solid line) for H$_{2}^{+}$. The durations are t=10 $\tau$, where $\tau$=2$\pi$/$\omega$ (i.e., 1 o.c. is 506.8 as for $\lambda$=150 nm) is fixed.

Equations (9)

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

i ψ ( r , t ) t = [ 1 2 r 2 + V c ( r ) + V L ( r , t ) ]   ψ ( r , t )
V c ( r ) = Z 1 ( x + R 1 ) 2 + y 2 + α Z 2 ( x R 2 ) 2 + y 2 + α
E ( t ) = E 0 f ( t ) [ cos ( ω t ) e ^ x + ξ sin ( ω t ) e ^ y ]
ψ ( r , t + Δ t ) = e i ( p 2 4 ) Δ t e ( i V c ( r ) i V L ( r , t ) ) Δ t e i ( p 2 4 ) Δ t ψ ( r , t ) + O ( Δ t ) 3  
j ( r , t ) = i 2 [ ψ ( r , t ) r ψ ( r , t ) ψ ( r , t ) r ψ ( r , t ) ]
B ( r , t ) = μ 0 4 π [ j ( r , t r ) | r r | 3 + 1 | r r | 2 c j ( r , t r ) t ] × ( r r ) d 3 r  
ψ 0 ( r , t ) = c g n ( t ) | ψ g n ( r ) e i E g n t / + c e n ( t ) | ψ e n ( r ) e i E e n t /
A ( r , t ) = | ψ 0 ( r , t ) | 2 = | c g n ( t ) ψ g n ( r ) | 2 + | c e n ( t ) ψ e n ( r ) | 2 + 2 | c g n ( t ) c e n ( t ) | ψ g n ( r ) ψ e n ( r ) c o s ( Δ E t )
J ( t ) = 0 j ( 0 , y , t ) e x d 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.