† Corresponding author. E-mail:
Project supported by the National Natural Science Foundation of China (Grant Nos. 11961131008, 11725416, and 11574010) and the National Key Research and Development Program of China (Grant No. 2018YFA0306302).
In this review, we will focus on recent progress on the investigations of nondipole effects in few-electron atoms and molecules interacting with light fields. We first briefly survey several popular theoretical methods and relevant concepts in strong field and attosecond physics beyond the dipole approximation. Physical phenomena stemming from the breakdown of the dipole approximation are then discussed in various topics, including the radiation pressure and photon-momentum transfer, the atomic stabilization, the dynamic interference, and the high-order harmonic generation. Whenever available, the corresponding experimental observations of these nondipole effects are also introduced respectively in each topics.
In most cases, the dipole approximation is used to simplify the theoretical treatment of the laser–matter interaction. The spatial dependence of the electromagnetic field of light is thus neglected and so one drops all higher orders of multipole interactions than the dipole one. This approximation is valid when the electromagnetic field has a wavelength much larger than the scale of the atomic and molecular system, which means that there is an upper limit on the light frequency to apply the dipole approximation.[1,2] In the short-wavelength regime, free-electron lasers (FELs)[3–5] can now produce x-rays at wavelengths down to 1 Å with an unprecedented intensity around 1020 W/cm2 (see, e.g., Ref. [6] and references therein). There also exists a lower limit to the frequency at which one can neglect the magnetic force of light.[7] In particular, recent advances in intense ultrafast lasers in the long wavelength regime,[8–14] has extended the strong-field sciences into mid-infrared (mid-IR) regime.[15–19] Technical developments in these new light sources have opened the door to experimentally investigate highly non-perturbative and relativistic effects in the laser–matter interaction. For the theoretical studies, the dipole approximation become invalid and nondipole effects can arise when the laser wavelength is comparable with the size of atoms or when the laser intensity increases so drastically that the electric and magnetic components of lights become equally important in the exhibition of the relativistic effects.[20,21]
The breakdown of the dipole approximation is expected to lead to a forward/backward asymmetry in the photoelectron angular distribution (PAD). For decades, a large number of researches have concentrated on experimentally measuring or theoretically calculating the nondipolar asymmetry parameters. Since the year around 1930, deviations from the dipole predictions have been shown in PADs for hard x-ray pulses with photon-energy well above 5 keV,[22–26] see also Ref. [27] and the references therein. Experimental[25,28–38] and theoretical[39–47] investigations on the nondipole effects in the PADs for the soft x-ray pulses with photon-energy below 5 keV can be traced back to the year of 1969. Nondipole effects have been observed at extremely low energies,[48–58] in which situations the dipole transition amplitude is usually depressed around a Cooper minimum[59] while the nondipole effect is resonance-enhanced. Measurements and calculations for the nondipole parameter of molecules have also been carried out, with a few debates on N2.[60–71] We note that, recent measurements have been made on the angular emission distributions of the 1s-photoelectrons of N2 ionized by linearly polarized synchrotron radiation at photon energy of 40 keV, in which an unexpected asymmetry with respect to the polarization direction has been observed.[72] Photoionization cross sections beyond the dipole approximation have been calculated for atoms through the full field operator,[73,74] which has also been used in Refs. [64,75] for molecules.
By integrating over those asymmetric electron distributions, one can obtain a net momentum of photoelectron along the direction of the light propagation, which comes from the linear photon-momentum partition between the photoelectrons and the residual ion. The photon-momentum transfer, which has been investigated respectively in the perturbative regime,[76–81] and in the tunneling regime,[77,79,82–95] can be used to measure the radiation pressure and give insights into the breakdown of the dipole approximation.
For the laser pulse at a high frequency and a super-high intensity, effects such as the atomic stabilization[96–108] and the dynamic interference,[109–120] which has been theoretically observed for decades, are in prospect of being demonstrated experimentally. The stabilization effect is expected to occur in a laser pulse at super-high intensities with its photon energy usually exceeding the binding potential of the system. In the case of atomic hydrogen, the stabilization can happen for photon energies above 13.6 eV and intensities higher than 1016 W/cm2.[102,119,121,122] Calculations[123,124] show that the manifestation of the nondipole effect on the atomic stabilization relies on the pulse duration, in which a window of intermediate pulse durations where the atomic stabilization is enhanced by the nondipole terms was proposed. In particular, the magnetic component of the laser field shows a detrimental effect on the atomic stabilization for short external pulses.[122] For atomic hydrogen, the atomic stabilization is feasible for an external laser pulse at a moderately high frequency (∼ 50 eV) and an extremely high intensity (above 1018 W/cm2), thus making the dynamic interference possible.[119] In such a high-intensity laser field, dynamic interference patterns beyond the dipole approximation have been theoretically studied.[120]
In the long-wavelength limit, the magnetic field-induced drift of the photoelectron in the laser propagation direction may reduce the chance of the ionized electron to revisit the residual ion,[125–128] which is expected to weaken the rescattering and recombination process. This will lead to nondipole effects on the high-order above-threshold ionization (ATI)[129,130] and the high-order harmonic generation (HHG)[131–137] spectra, which in the latter case shows a decrease in the yield of photons emitted along the laser polarization direction.[138–145] Besides, the symmetry of the system under the dipole approximation will be broken due to the magnetic field, with the presence of even harmonics[146–148] and with photons emitted along the laser propagation direction.[140,144,149,150] Both of them are forbidden in the dipole approximation because of the parity and symmetry of the system.
With the consideration of the magnetic component of the light field, attention should be also paid to the electronic spin dynamics. The spin–magnetic coupling is in the order of O(α) and the spin–orbit coupling is in the order of O(α2). Though comparable with the leading order of the nondipole terms, which includes the electric quadrupole and the magnetic dipole, the spin–magnetic coupling is negligible due to its spatial independence in the model with the lowest-order corrections beyond the dipole approximation. If higher-order multipole effects are considered, the spin–magnetic coupling becomes spatially dependent and the spin–orbit coupling is comparable with the second-order nondipole terms. Therefore, both of them need to be included in theoretical models and the electronic spin may play an important role in higher-order nondipole effects.[151–155] In practice, one needs to numerically solve the Dirac equation, instead of solving the time-dependent Schrödinger equation beyond the dipole approximation.
The rest of this review is arranged as follows. In Section
To describe the laser–matter interaction, an appropriate gauge form is crucial to the numerical convergence[156] and to the physical interpretation of the numerical results.[85] In the radiation gauge, the time-dependent Hamiltonian of a single-electron system interacting with a laser field is given by
The Hamiltonian in Eq. (
To solve the time-dependent Schrödinger equation corresponding to the Hamiltonian (
For a light field at a relatively weak intensity, the lowest-order perturbation theory (LOPT) can be applied to the one-photon ionization process. Deriving from the Hamiltonian in Eq. (
Likewise, the second-order perturbation theory can be used in the two-photon ionization process.[168,169] In the formulation of the Green’s function, the two-photon transition amplitude beyond the dipole approximation can be shown to be[81]
With the above formulations, the three-dimensional photoelectron momentum distribution (PMD) can be obtained by P(
Non-perturbative phenomena can occur when the laser intensity is sufficiently strong, which is beyond the description of the perturbation theory. In the case of intense infrared (IR) or mid-IR pulses, the electron dynamics can be treated semi-classically in the combined force of the electromagnetic field and the Coulomb potential. For example, the classical trajectory Monte Carlo (CTMC) simulations[173,174] is widely used to interpret the experimental data and to analyze the Coulomb effects in atomic ionization. In the CTMC model, the probability W(ti, v⊥) of an electron released from the atomic bound state at the moment ti and with an initial transverse velocity v⊥ is weighted by the PPT[175] ionization rate (or the ADK[176] ionization rate in the adiabatic picture). The tunneling electron is assumed to have a Gaussian transverse (perpendicular to the instantaneous electric field) velocity distribution,
After sampling the initial conditions of all electrons, their classical motion is governed by the Newton’s equation until the external field is turned off. With the leading-order (O(α)) nondipole term included, the motion equation of electrons ionized from the hydrogen atom in the combined electromagnetic field and the Coulomb field is
After the end of the laser pulse, the electron with a positive energy is regarded as ionized, and its final momentum is given by the Kepler’s formula. Finally, the electrons appearing at the tunnel exit at the time ti and with a similar final momentum
The above CTMC model beyond the dipole approximation have been applied to investigate the nondipole effects, such as the transfer and the partition of the photon linear momentum in atomic ionization.[84,93,177,178]
Besides the aforementioned semiclassical description, one can also develop a quantum mechanical formulation by generalizing the usual strong field approximation to the nondipole case. Applying an analytical expression of the nondipole Volkov state[79] and keeping the leading-order nondipole term, one can arrive at the transition amplitude from the initial bound state to the final state
In Eq. (
Numerically solving the time-dependent Schrödinger equation (TDSE) based on the Hamiltonian (
After the end of the external fields, we get the final wavefunction Ψf(
For the two-electron atom of helium, the Hamiltonian beyond the dipole approximation in multipole gauge is given by
Similarly, the two-electron wavefunction can be expanded into a set of coupled spherical harmonics which is in the time-dependent close-coupling (TDCC) scheme,[182]
For the two-center problem of a molecule, it is convenient to solve the corresponding TDSE in the spheroidal coordinates. The readers can get more details from Ref. [165] and references therein. Of course, in the perturbative regime, one can also develop a few-photon time-dependent perturbation theory in this coordinate system.
According to Cooper’s derivation,[43,44] the differential cross section of photoelectrons ejected from an isotropic system, or an anisotropic system but randomly orientated such as atoms or molecules in the gas phase can be written as
The first two terms on the right of Eq. (
The significance of higher-order nondipole effects, e.g., the pure-quadrupole and octupole–dipole interference terms, has also been demonstrated in PADs by a soft-x-ray pulse.[45] In this case, the vector potential of Eq. (
In the ionization process, the nondipole path leads to a partial wave of different angular momenta, and it has been understood for many years that nondipole effects can be observed in the PAD for the single electron system.[44] For two-electron systems, there exist additional features: (i) there could be a nondipole resonance resulting in the magnification of nondipole effects; (ii) core relaxation exists and brings a nonzero parameter δ in PADs; (iii) the total angular momentum and energy are shared between two electrons with the consideration of the correlation, so for a nonsequential process the formula of PAD for the single-electron case becomes invalid, and one has to look at the double differential cross section (DDCS). In 2003, the observation of PAD in helium photoionization[51,52] confirmed the nondipole parameter can be enhanced by the quadrupole Fano resonance (which does not exist in single electron systems), and a good agreement between the ab initio calculations and experimental measurements was achieved. After several years, the DDCS (similar to PAD with a varying energy) of the single photon double ionization (SDPI) is theoretically investigated with the lowest-order perturbation theory and the convergent close-coupling theory (CCC),[188] revealing a nonzero parameter δ, which is usually zero for a single electron photoionization from 1s state.[44] Recently, the nondipole PAD parameters were experimentally observed in the two photon double ionization (TPDI) process in a sequential region using FELs at FERMI,[58] compared with theoretical calculations developed by Grum-Grzhimailo et al.[56,189] This study reveals a large nondipole parameter
In the XUV region, the single photon double ionization of helium atom has been well studied with dipole approximation, for reviews, see Refs. [190,191]. However, when the photon energy reaches hundreds of eV, the nondipole effects become pronounced. In 2004, Istomin et al. reported their calculations for nondipole effects in VUV region.[192] At an excess energy as low as 80 eV, a forward–backward asymmetry in the triply differential cross section (TDCS) was confirmed with a linearly polarized light. The case of an elliptically polarized light was discussed in the same year,[193] and they found that for the equal energy sharing, the circular dichroism exists, which is expected to be zero within the dipole approximation. Later, the details of their methods were published,[194] in which the electron’s evolution is separately calculated with the ground state correlation and the final state correlation included. They showed TDCS at different geometries at an excess energy of 20, 100, and 450 eV, but their absolute TDCSs failed to get agreement with the the experimental results at excess energy of 450 eV. However, the ratio of the difference of the TDCSs in the forward and backward half-planes to the TDCS in the forward half-plane (Eq. (69) in Ref. [194]) is in a qualitative agreement. Considering the drawbacks of LOPT in handling the electron correlation, they investigated the TDCS at an excess energy of 450 eV with CCC method in 2006,[195] showing an improved agreement with the experiment measurements.
Besides these work by Istomin and coworkers, there were also several investigations on the nondipole effects in helium. Meharg et al. developed the time-dependent close-coupling method (for a review, see Ref. [196]) in the full-dimensionality and presented the applications for helium and hydrogen in intense IR laser fields. It should be noted that they dropped the term −i
The early studies about nondipole effects in molecular PADs focused on the randomly oriented molecules. The first observation of a significance nondipole effect on molecule was made for N2 in 2001.[60] They measured the ζ parameter for the photoelectron ionized from K-shell with a photon energy about hundreds of eV. A large nondipole signal (ζ close to one) was observed for a near-threshold ionization with photoelectron energy around 50 eV. A corresponding theoretical calculation[200,201] was performed based on the fixed-nuclear and the frozen-core Hartree–Fock (HF) approximation, and the main feature can be repeated from the computation. The resonance peak near the threshold is contributed by the σ* resonance of the dipole transition moment with photoelectron energy around 10 eV,[202,203] which comes from the multiple-electron correlation, though their positions do not coincide with each other. In the next year, a coincidence experiment was carried out where the angular distribution for fixed-in-space N2 (Ref. [61]) was measured at 660-eV photon energy for the K-shell photoelectron. A clear asymmetric structure can be directly seen from the angular distribution, as shown in Fig.
However, an independent experiment performed at the Photon Factory in Tsukuba, Japan[68] showed a negligible nondipole effect for the near-threshold ionization of K-shell electron from N2. Their computation based on the relaxed-core HF with the random phase approximation (RPA) for the electron correlation confirmed their experimental observation. A similar conclusion for C K-shell of CO was also obtained.[67]
The third experiment was carried out at the Elettra synchrotron radiation source, Italy.[204] Together with their theory based on the density functional theory (DFT) with the single center expansion,[66] their results are consistent with the Japanese group. Oscillation of nondipole parameters with photo energy from σg and σu states was observed theoretically, which can be understood with the classical Cohen–Fano model:[205] electrons emitted from the two nuclei interfere with each other, resulting in an oscillation factor of ei
Aside from the debates on N2, there were many theoretical works focused on nondipole effect of other molecules. Grum-Grzhimailo[62] reformulated the theory of angular distributions and angular correlations of photoelectrons and recoil ions in the molecular photoionization in terms of the density matrix and statistical tensor formalism, which incorporates a full multipole expansion of the radiation field. This theory was applied to compute the nondipole correlation for N 1s photoionization of NO (Refs. [70,209] and the circular dichroism in the molecular ion orientation distribution.[69] Seabra et al.[64] computed the nondipole correlation at photo energy of several keV for various molecules with the HF and the random phase approximation, including CH4, NH3, H2O, and HF. Brumboiu et al.[210] calculated the nondipole ionization cross section for molecules with various of sizes by using the Gaussian-type orbitals for initial state and plane waves for final states.} Baltenkov et al.[211] turned back to the single electron case, giving an analytic solution to the photoionization from a diatomic molecule with a zero-range potential, which provided some physical insights to the two-center nondipole effects.
Photoionization of an atom by a photon with energy ω involves absorption of the photon’s linear momentum
Within the perturbative regime, the momentum transfer in the photoelectric effect for the electron ionized from the ground state of the hydrogenlike atom is
Things would be much more complex for other systems, e.g., molecules and two-electron systems. The average momentum of the ionized electron is not necessarily collinear with the photon momentum, since the interaction between the electron and the residual ions can be anisotropic. Lao et al.[212] provided an expression of photo momentum transfer for the diatomic molecule with zero-range potential. Liang et al.[165] carried out ab initio calculations on the nondipole correlation of
About 44 years ago, Amusia et al. took nondipole effects into account in the single photon double ionization, predicting the so called quasi-free mechanism (QFM),[213] by which the electrons can break the well-known selection rules within the dipole approximation that the equal energy sharing back-to-back emission is forbidden.[190] The details of QFM can be found in Ref. [214]. In 2012, Galstyan et al. calculated the nondipole effects by the perturbatiion theory with the final state described by the three-body double-continuum Coulomb function (3C) and the uncorrelated 2C function.[215] Their results showed the breakdown of the selection rules mentioned above, and this effect is strongest in helium than other helium-like atoms. When electrons with an equal energy sharing emit back-to-back, the ion momentum equals to the photon momentum absorbed in the photon ionization process, which is about 0.2 a.u. for a photon energy at 800 eV, and the typical ion momentum from the dipole path is about 6 a.u. in this situation. As a result, the yield of ions which have a momentum close to zero is a sign of the nondipole effects. Schöffler et al.[216] observed such a phenomenon in 2013 at photon energies of 440 eV and 800 eV, but their theoretical calculations by CCC and TDCC method overestimated the nondipole effects (i.e., the yield of ion momentum close to zero is as strong as the dipole yield). Five years later, they managed to realize kinematically complete measurement of SPDI, reported the differential cross section for the coplanar geometry and equal sharing, in which the back-to-back emission can be clearly seen and separated from the dipole yield, and the experiment is in great agreement with the CCC theory.[217]
In Ref. [218], the photon momentum transfer in the single-photon double ionization of helium is investigated experimentally and theoretically. Their main conclusions are shown in Fig.
In the atomic two-photon ionization, the total cross sections and angular distributions of electrons beyond the dipole approximation have been widely discussed for a long time.[219–227] Numerical fitting shows that a good linear relationship like Eq. (
By increasing the laser intensity of the IR or Mid-IR light, Smeenk et al.[82] and Ludwig et al.[88] have successively reported experimental observations of a photoelectron momentum shift along the direction of laser propagation in the tunneling ionization process. Theoretical works have confirmed this asymmetrical electron momentum distribution and a nonzero momentum shift along the laser’s propagation direction by quantum-mechanical calculations,[83,91] semiclassical models with the Lorentz force included,[84,88,92] solution to the time-dependent Dirac equation,[90] and strong field approximation models beyond the dipole approximation.[77,79,128,180] The pioneering work by Chelkowski et al.[77] has predicted different photon-momentum partitioning rules for one-photon ionization and multiphoton processes, the former of which has been discussed above and the latter of which has been supported by a recent experiment.[229] Within the long-wavelength limit of a linearly polarized light, the Coulomb interaction and the rescattering events may result in a momentum shift opposite to the direction of laser propagation,[79,88,92] particularly for the low-energy electrons. Besides, the under-barrier motion caused by the laser–magnetic-field-induced Lorentz force has been reported to be relevant to the electron momentum shift in the laser propagation direction.[86,87,89,95]
In the tunneling regime, figure
However, a constant net momentum shift equal to 0.3αIp has been proposed in theoretical works mainly based on the SFA,[77,79,86,87] which makes the partition rule in the tunneling regime change into
Besides the momentum transfer law in the long-wavelength limit, Ludwig et al.[88] experimentally observed an increasing shift of the peak of f(pz) opposite to the beam propagation direction (z axis) with increasing laser intensities of the linearly-polarized pulse, as shown in Fig.
Nondipole effects for the one-photon ionization in the perturbative regime are discussed in Subsection
According to the Einstein’s photoelectric law, a single peak at the energy of ω – Ip is expected in the photoelectron energy spectrum for one-photon ionization in the perturbative process. However, the single peak can gradually evolve into a multi-peak structure due to the dynamic interference when the laser intensity is increased to reach the nonperturbative regime, see the photoelectron energy distribution D(E) at various laser intensities in Figs.
In Fig.
Detailed angularly distinguished momentum spectra of the photoelectron within/beyond the dipole application are presented, for example the momentum distributions in the polar coordinates of the polarization–propagation (px–pz) plane in Figs.
By including the nondipole corrections for the Volkov phase[79,106] into the semi-analytical model previously developed for the dynamic interference,[119] the momentum shift observed in the TDSE calculations is nicely reproduced, as is shown in Fig.
In Subsection
On one hand, for photons emitted along the laser polarization direction, calculations for He+ ions by a semiclassical theory[138] indicate that the nondipole corrections can lead to a decrease of the high harmonic yield and a shift of the HHG spectra. Similar results have been obtained by a fully relativistic treatment[139,141,142] and theoretical works based on SFA.[140,142,143,145] In 2016, Zhu et al.[144] proposed different rules for the decrease of harmonic yield due to the nondipole effects in the high-intensity region and in the long-wavelength region respectively.
On the other hand, emitted photons have been found along the laser propagation direction, which is forbidden within the dipole approximation. Potvliege et al.[149] claimed that photon emission by He+ ions along the propagation direction of a few-cycle laser pulse with wavelength of 800 nm and peak intensity of 5.6 × 1015 W/cm2 was few orders of magnitude weaker than that along the laser polarization direction. Similar conclusions have been reported for hydrogen-like ions[140] and multiply charged ions.[150]
Besides, the magnetic-field component of the laser will break the symmetry of the system and even harmonics will appear in the spectra.[148] Vázquez et al.[146] have seen intense even harmonics emitted from a 3D hydrogen atom interacting with a high-frequency super-intense linearly polarized laser field. In 2006, Bandrauk et al.[147] presented the TDSE calculations of the nondipole effects in HHG for a Born–Oppenheimer 3D molecule
The dipole approximation has been commonly used in the theoretical treatment of the laser pulse interacting with atoms and molecules. When the laser wavelength is comparable with the atomic size or when the photoelectrons can be accelerated to a considerable velocity at high laser intensities, this approximation fails and physical phenomena related to the nondipole effects become obvious. The inclusion of the nondipole corrections will break the symmetry of the system’s Hamiltonian, which can be seen from theoretical models introduced above like numerically solving the time-dependent Schrödinger equation, the perturbation theory, the classical trajectory Monte Carlo method, and the strong field approximation. In this review, we have discussed the exhibition of the nondipole effects in both the perturbative and the non-perturbative regime, especially including topics such as the asymmetry photoelectron angular distributions, the photon momentum transfer, the dynamic interference, and high-order harmonic generation. Experimental and theoretical works in few-electron and molecular systems have been discussed.
[1] | |
[2] | |
[3] | |
[4] | |
[5] | |
[6] | |
[7] | |
[8] | |
[9] | |
[10] | |
[11] | |
[12] | |
[13] | |
[14] | |
[15] | |
[16] | |
[17] | |
[18] | |
[19] | |
[20] | |
[21] | |
[22] | |
[23] | |
[24] | |
[25] | |
[26] | |
[27] | |
[28] | |
[29] | |
[30] | |
[31] | |
[32] | |
[33] | |
[34] | |
[35] | |
[36] | |
[37] | |
[38] | |
[39] | |
[40] | |
[41] | |
[42] | |
[43] | |
[44] | |
[45] | |
[46] | |
[47] | |
[48] | |
[49] | |
[50] | |
[51] | |
[52] | |
[53] | |
[54] | |
[55] | |
[56] | |
[57] | |
[58] | |
[59] | |
[60] | |
[61] | |
[62] | |
[63] | |
[64] | |
[65] | |
[66] | |
[67] | |
[68] | |
[69] | |
[70] | |
[71] | |
[72] | |
[73] | |
[74] | |
[75] | |
[76] | |
[77] | |
[78] | |
[79] | |
[80] | |
[81] | |
[82] | |
[83] | |
[84] | |
[85] | |
[86] | |
[87] | |
[88] | |
[89] | |
[90] | |
[91] | |
[92] | |
[93] | |
[94] | |
[95] | |
[96] | |
[97] | |
[98] | |
[99] | |
[100] | |
[101] | |
[102] | |
[103] | |
[104] | |
[105] | |
[106] | |
[107] | |
[108] | |
[109] | |
[110] | |
[111] | |
[112] | |
[113] | |
[114] | |
[115] | |
[116] | |
[117] | |
[118] | |
[119] | |
[120] | |
[121] | |
[122] | |
[123] | |
[124] | |
[125] | |
[126] | |
[127] | |
[128] | |
[129] | |
[130] | |
[131] | |
[132] | |
[133] | |
[134] | |
[135] | |
[136] | |
[137] | |
[138] | |
[139] | |
[140] | |
[141] | |
[142] | |
[143] | |
[144] | |
[145] | |
[146] | |
[147] | |
[148] | |
[149] | |
[150] | |
[151] | |
[152] | |
[153] | |
[154] | |
[155] | |
[156] | |
[157] | |
[158] | |
[159] | |
[160] | |
[161] | |
[162] | |
[163] | |
[164] | |
[165] | |
[166] | |
[167] | |
[168] | |
[169] | |
[170] | |
[171] | |
[172] | |
[173] | |
[174] | |
[175] | |
[176] | |
[177] | |
[178] | |
[179] | |
[180] | |
[181] | |
[182] | |
[183] | |
[184] | |
[185] | |
[186] | |
[187] | |
[188] | |
[189] | |
[190] | |
[191] | |
[192] | |
[193] | |
[194] | |
[195] | |
[196] | |
[197] | |
[198] | |
[199] | |
[200] | |
[201] | |
[202] | |
[203] | |
[204] | |
[205] | |
[206] | |
[207] | |
[208] | |
[209] | |
[210] | |
[211] | |
[212] | |
[213] | |
[214] | |
[215] | |
[216] | |
[217] | |
[218] | |
[219] | |
[220] | |
[221] | |
[222] | |
[223] | |
[224] | |
[225] | |
[226] | |
[227] | |
[228] | |
[229] | |
[230] | |
[231] | |
[232] | |
[233] | |
[234] | |
[235] | |
[236] | |
[237] | |
[238] | |
[239] | |
[240] | |
[241] |