• Aucun résultat trouvé

Photodissociation of the HeH+ ion into excited fragments (n=2,3) by time-dependent methods

N/A
N/A
Protected

Academic year: 2021

Partager "Photodissociation of the HeH+ ion into excited fragments (n=2,3) by time-dependent methods"

Copied!
9
0
0

Texte intégral

(1)

Photodissociation of the HeH

+

ion into excited fragments (n = 2 , 3) by time-dependent methods

K. Sodoga,1,2J. Loreau,3 D. Lauvergnat,1Y. Justum,1N. Vaeck,3and M. Desouter-Lecomte1,4,

*

1

Laboratoire de Chimie Physique, Université de Paris-Sud, UMR 8000, Bât 349, Orsay F-91405, France 2

Université de Lomé, Faculté des Sciences, Département de Physique, BP 1515 Lomé, Togo 3

Laboratoire de Chimie Quantique et Photophysique, Université Libre de Bruxelles, CP160/09 50, av. F. Roosevelt, 1050 Bruxelles, Belgium

4Département de Chimie, B6c, Université de Liège, Sart-Tilman, B-4000 Liège 1, Belgium

共Received 12 June 2009; published 28 September 2009兲

The total and partial photodissociation cross sections of the molecular ion HeH+ are computed by

time-dependent methods for fragmentation into the excited shells n = 1 , 2 , 3 up to a photon energy of 40 eV.1⌺+and 1⌸ states are considered for parallel and perpendicular transitions for different initial rotational or vibrational

excitations. Nonadiabatic radial and rotational couplings are taken into account. The results from coupled-channel equations are compared with the Born-Oppenheimer approximation. A time-dependent calculation with a femtosecond laser pulse is carried out to simulate a recent crossed beam photodissociation imaging experiment with vacuum ultraviolet free-electron laser关H. B. Pedersen et al., Phys. Rev. Lett. 98, 223202 共2007兲兴. The dominance of photodissociation perpendicular to the photon polarization is confirmed.

DOI:10.1103/PhysRevA.80.033417 PACS number共s兲: 33.80.⫺b, 34.70.⫹e

I. INTRODUCTION

The HeH+ helium-hydride molecular ion has been the subject of numerous experimental 关1–6兴 and theoretical works关7–16兴 due to its interest in the context of astrophysics 关17–19兴. It has been suggested that HeH+could be abundant enough in different celestial objects composed of primordial material such as metal-poor stars, white dwarfs or the plan-etary nebula, to be observed in the infrared domain 关20兴. However, despite numerous efforts, none of the several at-tempts of extraterrestrial observation of HeH+ have been conclusive. One of the key points in the evaluation of the fractional abundance of HeH+is to determine a correct bal-ance between the formation and destruction rate processes; one of these processes being photodissociation. The rate of photodissociation was determined either by detailed balanc-ing applied to the reverse reaction, i.e. radiative association 关19,21兴 or by absorption from the ground state to the vibra-tional continuum of the first excited state only 关17,22兴. Re-cently, the first experimental data on the photodissociation of HeH+, i.e., a crossed beam photodissociation imaging with vacuum ultra violet free-electron laser in Hamburg FLASH 关23兴 has renewed the interest in the fragmentation into highly excited fragments 共nⱖ2兲. The experiment involves photons of wavelength 32 nm共38.7 eV兲 and resolves the kinetic en-ergy release of the neutral He fragment and the correspond-ing angular distribution with respect to the photon polariza-tion. This experiment on the HeH+ system reveals the dominance of photodissociation perpendicular to the laser polarization and therefore the important role of the⌸ states during dissociation with energetic photons. This shows the necessity of taking into account the perpendicular orientation generally not considered or not completely considered in previous theoretical models. The photodissociation has been recently revisited for both orientations 关16兴 and compared

with previous results for parallel orientation 关12兴. The cross section is computed in Ref. 关16兴 by the Fermi Golden rule within the Born-Oppenheimer approximation with a possible correction to consider the avoided crossing between the sec-ond and third excited states of the⌺ manifold. In this work, we present a time-dependent approach of the photodissocia-tion by including all the radial nonadiabatic couplings and the rotational couplings between the 1⌺ and 1⌸ states. The time-dependent approach is equivalent to the Golden rule treatment关24兴. In a first approach, the total photodissociation cross section is computed using the autocorrelation function of the promoted state by considering different vibrational and rotational initial states and the partial cross section is ob-tained by two different methods using the asymptotic wave packets. Secondly, we simulate the free-electron laser experi-ment by exciting the ground initial state with a Gaussian laser pulse of 30 fs and a carrier frequency corresponding to the experimental wavelength. The kinetic energy of the He fragment is estimated from the asymptotic wave packets in order to compare with the experimental data.

II. COMPUTATIONAL DETAILS A. Molecular data

We consider the singlet 1⌺+ and the 1⌸ adiabatic elec-tronic states which dissociate into He共1snl兲+H+ and He+共1s兲+H共nl兲 with n=1,2,3. There exist 12 e-labeled1+ states, 6 e-labeled1⌸eand 6 f-labeled 1f. The parity under

the Eⴱ inversion operator in the laboratory frame 关25兴 of a ro-electronic state is given by共−1兲J␧, where J is the quantum number of the rotation and␧=+1 or −1 correspond to e or f states 关26兴. For the electronic part, the Eⴱ inversion corre-sponds to the reflection in the xz plane of the molecular frame. Due to symmetry, there is no nonadiabatic coupling between the e and f states 关27兴. The potential energy curves and couplings are computed at the state average complete active space self-consistent field 共CASSCF兲 关28兴 level by *Corresponding author; mdesoute@lcp.u-psud.fr

(2)

using the ab initio quantum chemistry packageMOLPRO关29兴. A large nonstandard basis set is used by adding one con-tracted Gaussian per orbital per atom up to n = 4 to the aug-cc-pV5Z basis set共augmented correlation-consistent valence quintuple zeta兲 关30兴. Details and comparison with other the-oretical works are given in Ref.关31兴. The CASSCF approach allows the computation of the radial nonadiabatic coupling elements of the first derivative⳵R operator. The F共R兲 matrix

F␣␤共R兲=具␹ad兩⳵R兩␹

ad典 is used to build diabatic electronic

states by solving the coupled equations ⳵RD共R兲=

−F共R兲D共R兲 for the elements of the adiabatic-to-diabatic D共R兲 matrix transformation

Vdia共R兲 = DT共R兲Vadia共R兲D共R兲,

where V is the matrix of the electronic Hamiltonian Hˆel

which is diagonal in the adiabatic representation. We use an approximate F共R兲 matrix by retaining only the main cou-plings between neighboring states. This structure corre-sponds to a succession of 2⫻2 coupling cases. This ap-proach, which reduces considerably the computational cost and simplifies the diabatization procedure, has been shown to give similar results to the use of the complete F共R兲 matrix in curve crossing dynamics关32兴. The adiabatic and diabatic po-tential energy curves are represented in Fig.1. The dissocia-tion channels for the 1⌺+and1⌸ states are gathered in Table

I. The Stark effect due to He+ lifts the degeneracy of the excited hydrogen atom as it has been already mentioned in Ref. 关8兴.

We also include in the time-dependent treatment the effect of the rotational couplings but we neglect the spin-dependent interactions. The field free Hamiltonian is Hˆ0= Hˆel+ Tˆrad

+ Tˆrot. We use a basis set of electronic-rotational functions to

get time-dependent coupled equations for the nuclear mo-tion. These parity adapted functions are

兩mJM⌳␧典 = 1

关2共1 +␦⌳0兲兴1/2关兩JM⌳典兩m⌳典 +共− 1兲J␧兩JM,− ⌳典兩m共− ⌳兲典兴

where m numbers the electronic states for a given ⌳. -3 -2.5 -2 -1.5 4 8 12 16 20 24 28 Energy (hartree) R (bohr) Adiabatic -3 -2.8 -2.6 -2.4 -2.2 -2 -1.8 -1.6 4 8 12 16 20 24 28 R (bohr) Energy (hartree) Diabatic

FIG. 1. 共Color online兲 Adiabatic and diabatic potential energy curves of the 1⌺+ 共full lines兲 and 1⌸ 共dashed lines兲 of the HeH+

molecular ion. The adiabatic curves are computed at the state aver-age CASSCF level. The diabatic curves are obtained from the trans-formation matrix D共R兲 solution of the equationRD共R兲= −F共R兲D共R兲 where F共R兲 is the approximate “2⫻2” nonadiabatic radial coupling matrix.

TABLE I. CASSCF asymptotic energies of the1⌺+and1⌸ states included in the calculations. State

m

Energy共h兲

at 50 bohr Dissociative atomic states

1 −2.8980458 ⌺, He共1s2 1S兲+H+ 2 −2.4999359 ⌺, He+共1s兲+H共1s兲 3 −2.1448183 ⌺, He共1s2s1S兲+H+ 4 −2.1261793 ⌺, He+共1s兲+1/

2关H共2s兲−H共2p兲兴 5 −2.1237656 ⌺, He+共1s兲+1/

2关H共2s兲+H共2p兲兴 6 −2.1225613 ⌺, He共1s2p1Po兲+H+ 7 −2.0623295 ⌺, He共1s3s1S兲+H+ 8 −2.0600078 ⌺, He+共1s兲+1/

3 · H共3s兲+1/

2 · H共3p兲+1/

6 · H共3d兲 9 −2.0568303 ⌺, He共1s3d1D兲+H+ 10 −2.0552158 ⌺, He+共1s兲+1/

3 . H共3s兲−

2/3.H共3d兲 11 −2.0528106 ⌺, He共1s3p1Po兲+H+ 12 −2.0523058 ⌺, He+共1s兲+1/

3 · H共3s兲−1/

2 · H共3p兲+1/

6 · H共3d兲 1 −2.1248933 ⌸, He+共1s兲+H共2p兲 2 −2.1236547 ⌸, He共1s2p1Po兲+H+ 3 −2.0571436 ⌸, He+共1s兲+1/

2关H共3p兲−H共3d兲兴 4 −2.0567818 ⌸, He共1s3d1D兲+H+ 5 −2.0536464 ⌸, He+共1s兲+1/

2关H共3p兲+H共3d兲兴 6 −2.0531961 ⌸, He共1s3p1Po兲+H+

(3)

兩JM⌳典=关共2J+1兲/4␲兴1/2D

M

共J兲共␾,, 0兲 are the eigenstates of the total angular momentum representation where D is a Wigner function关27兴. M and ⌳ are the projection of the total electronic angular momentum L onto the laboratory Z axis and onto the internuclear z axis, respectively. ⌳ is also the projection of the total molecular angular momentum J onto the internuclear axis because in this case the total electronic spin S is zero共⍀=⌳+S=⌳兲. The rotational kinetic operator takes the form Tˆrot= Nˆ2/共2␮R2兲 with Nˆ2= Jˆ2+ Lˆ2− 2Jˆ . Lˆ and

Jˆ . Lˆ = JˆzLˆz+共Jˆ++ Jˆ+兲/2. After the action of the ladder op-erator Jˆ兩JM⌳典=关J共J+1兲−⌳共⌳⫿1兲兴1/2兩JM ,⌳⫿1典, the ro-tational couplings between symmetry adapted ⌺共⌳=0兲 and ⌸共⌳= ⫾1兲 states become 具m

J

M

⌺兩Tˆrot兩mJM⌸e典 = −关J共J + 1兲兴 1/2 2␮R2 1 21/2共具m

⌺兩Lˆ兩m⌳ = 1典 +具m

⌺兩Lˆ+兩m⌳ = − 1典兲␦J,J⬘␦MM⬘.

For diatomic molecules, MOLPRO works in the C2v group,

which allows us to calculate the matrix elements of Lxand Ly

between the⌺ states and the ⌸xand⌸ycomponents of the⌸

states. We obtain the matrix elements 具m

⌺兩Lˆy兩m⌸x典=

−iammand具m

⌺兩Lˆx兩m⌸y典=iamm, and transform them in

or-der to get 具m

⌺兩Lˆ兩m⌳= ⫿1典. The final expression of the rotational couplings is

具m

J

M

⌺兩Tˆrot兩mJM⌸典 = 关J共J + 1兲兴1/2amm/共␮R2兲␦J,J⬘␦MM⬘.

We neglect the contribution from the共Lˆx2+ Lˆy2兲/2␮R2term. In the diagonal rotational correction 共m=m

and ⌳=⌳

兲, we include the term

具m

J

M

兩Tˆrot兩mJM⌳典

=关J共J + 1兲 − 2⌳2兴/共2␮R2兲␦

mm⬘␦⌳⌳⬘␦J,J⬘␦MM⬘.

The rotational coupling matrix is computed in the adiabatic basis set and transformed to the diabatic one

Trot

dia共R兲 = DT共R兲T rot

adia共R兲D共R兲.

B. Photodissociation cross section

The photodissociation cross section towards 1⌺ or 1⌸ states from a ro-vibrational state ␹0adiavJ共R兲 of energy E0vJ

where 0 denotes the ground electronic state1⌺+共⌳

= 0兲 is ␴0vJ→⌳共E兲 =

J

SJ,J⬘⌳

2J

+ 1 ␴0vJ→⌳J共E兲, 共1兲 where the Hönl-London factor for symmetry adapted states is given by关33兴 SJ⬙⌳⬙,J⬘⌳⬘=共1 +␦⌳⬘0+␦⌳⬙0− 2␦⌳⬘0␦⌳⬙0兲共2J

+ 1兲 ⫻ 共2J

+ 1兲

J

1 J

−⌳

−⌳

2 . Each contribution ␴0vJ→⌳J共E兲 is obtained by Fourier

transforming the autocorrelation function C0vJJ共t兲

=兺m具⌽m⌳⬘J⬘ 0vJ共t=0兲兩⌽m⌳⬘J⬘ 0vJ共t兲典 of a promoted state ⌽m⌳⬘J⬘ 0vJ共t = 0兲. In the adiabatic representation, each component ⌽m⌳⬘J

0vJad

共t=0兲=␮j,0→m⌳ad

共R兲␹0vJ

ad 共R兲 is built by multiplying

the initial nuclear state ␹0advJ共R兲 of the ground adiabatic electronic state by the adiabatic components of the transition moment ␮adj,0→m⌳, with j = x , y , z. The vector is then trans-formed to the diabatic representation ⌽dia共t=0兲

= D共R兲⌽adia共t=0兲 and propagated with the field free coupled

equations on the excited states with angular momentum J

with J

= J

− 1 and J

+ 1 for the1⌺→1⌺ transitions and J

= J

− 1 and J

+ 1共involving 1⌸e states兲 and J

= J

共involv-ing 1⌸f states兲 for1⌺→1⌸ transitions 共J

= 1 only when J

= 0兲. The cross section ␴0vJ→⌳J共E兲 in SI units is then

given by 关34,35兴

␴0vJ→⌳J共E兲 = 4␲2␣a02EAJ

0vJ共E兲, 共2兲

where␣= e2/共បc4␲␧0兲 is the fine structure constant and a0is the Bohr radius and

A0vJJ共E兲 = 2Re

1

2␲

0 ⬁

C0vJJ共t兲ei共E0vJ+E兲t/បdt

The partial cross sections can be obtained by different meth-ods from an analysis of the wave packets in the asymptotic region. The first one uses the Fourier transformation of each component of the wave packet at a large internuclear dis-tance, R, where the couplings become vanishingly small and the dissociation channels are diabatic or adiabatic 关36兴. Then the partial cross-section is given by

␴0vJ→m⌳J共E兲 = 4␲2␣a02kmE兩Am⌳⬘J⬘ 0vJ共E兲兩2, 共3兲 where km=

2␮共E−Em

as兲/ប is the magnitude of the wave

number in dissociative channel m with an asymptotic energy

Emasand Am0vJJ共E兲 =

1 2␲

0 ⬁ ⌽m⌳⬘J⬘ 0vJ

共R,t兲ei共E0vJ+E兲t/បdt. The second method uses the momentum representation of the wave packet关37,35兴 ⌽¯ m⌳⬘J⬘ 0vJ共k,t⬁兲 =

1 2␲

0 ⬁ ⌽m⌳⬘J⬘ 0vJ共R,t兲e−ikRdR and leads to the expression

␴0vJ→m⌳J共E兲 = 4␲2a 0 2 kn E兩⌽¯0mvJJ共k,t兲兩2. 共4兲 The field free propagation is carried out in diabatic represen-tation by the split operator method关38兴 extended to

(4)

nonadia-batic processes formalism 关39兴. This requires two transfor-mations to change from the diabatic to the rotationally adiabatic representation共by diagonalizing the J

bloc of the matrix of Hˆel+ Tˆrot兲 and two Fourier transformations to

change from the coordinate to the momentum representations in order to apply the kinetic part of the propagator. The method based on Eq. 共4兲 requires a larger spatial grid than the approach based on Eq. 共3兲 since at the final time, the whole wave packet must be in the asymptotic region before the optical potential which avoids the reflection at the end of the grid. We use 213points for a grid of 100 a.u. A quadratic optical potential begins at 80 a.u. The asymptotic point R 关Eq. 共3兲兴 is 35 a.u. and the final time t关Eq. 共4兲兴 is 61 fs. The time step is 2.41⫻10−2 fs.

C. Dynamics with a femtosecond pulse

The propagation is carried out again by the split operator by keeping the field constant during a small time interval. The elementary evolution operator for a time step is given by

U共⌬t兲␺共tk兲 = e−i⌬t/共4ប兲Vei⌬t/共2ប兲␮␧共tke−i⌬t/共4ប兲Ve−i⌬t/បT

⫻e−i⌬t/共4ប兲Vei⌬t/共2ប兲␮␧共tke−i⌬t/共4ប兲V␺共t k兲.

The propagator ei⌬t/共2ប兲␮␧共tkis built by diagonalizing the transition moment matrix ␮one time and by multiplying at each step the eigenvalues by the scalar␧共tk兲.

We use a Gaussian femtosecond laser field with param-eters inspired from the experimental conditions. The carrier frequency is fixed by the experimental wave length共32 nm兲. The half-height width of the envelope is 30 fs. The total duration of the propagation is 100 fs with a time step of

2.41⫻10−2 fs. We use an intensity of 8.775

⫻1011 W cm−2 which is larger than the experimental inten-sity but the probabilities of transition

Pm共t兲 = 具⌽mJ

0vJ共t兲兩⌽ mJ

0vJ共t兲典 共5兲

are very weak and we increased the intensity in order to obtain probabilities of the order of 0.1%.

III. PHOTODISSOCIATION CROSS SECTION

Figure 2 shows the photodissociation cross section ␴0vJ→⌳共E兲 computed by Eqs. 共1兲 and 共2兲 for 1⌺→1⌺, i.e.

for the parallel orientation of the molecular axis with respect to the field for different initial states. The cases J

= 0→J

= 1 共full line兲 and J

= 1→J

= 0 , 2 共dotted line with open circles兲 for v=0 are nearly superimposed showing a negli-gible influence of the rotational excitation共the difference be-tween the maximum of the cross section for J

= 0 and J

= 2 is 0.07%兲. We also show two cases with higher rotational and with vibrational excitation. The example J

= 10→J

= 9 , 11 for v = 0 is drawn in dashed line. The difference

be-tween the maximum of the cross section for J

= 9 and J

= 11 is of 0.6% and thus a little bit larger. The peak around 20 eV is related to the photodissociation via the first excited electronic state leading to He+共1s兲+H共1s兲 and is strongly shifted from the maximum of the J

= 0 case. The mean equi-librium distance for the J

= 10 state is 1.86 a.u. while it is

1.49 a.u. for J

= 0. This leads to a very different Franck Condon gap between the ground and first excited states. However, the cross section is only weakly modified in the range probed by the FEL experiment around 38.7 eV 关23兴. The same tendency is observed for the v = 1, J

= 0 example drawn in dotted line in Fig.2. The effect is discussed again below共see Fig.3兲 for the average cross section over an iso-tropic orientation.

Figure 3 gives the photodissociation cross section for three cases of Fig. 2 averaged over an isotropic orientation by weighting each orientation x, y, z by 1/3. The parallel orientation共␮z兲 concerns the transition1⌺→1⌺ and the

per-pendicular orientations 共␮x and ␮y兲 are related to the 1⌺

1⌸ transitions. One observes that the average cross sec-tions for initial excited rotational and vibrational states are of

0 5 10-18 1 10-17 1.5 10-17 2 10-17 2.5 10-17 15 20 25 30 35 40 45 v=0, J"=0 v=0, J"=1 v=0, J"=10 v=1, J"=0 Cross section (cm 2)

Photon energy (eV)

× ×

× × ×

FIG. 2. 共Color online兲 Total photodissociation cross section ␴0vJ→⌳共E兲 关Eq. 共1兲兴 for the transition1⌺→1⌺ for the parallel

orientation of the molecular axis with respect to the field for the case J⬙= 0→J⬘= 1 共full line兲, J⬙= 1→J⬘= 0 , 2 共dotted line with open circles兲, J⬙= 10→J⬘= 9 , 11共dashes兲 for v=0 and J⬙= 0→J⬘ = 1 forv = 1共dots兲. 0 5 10-18 1 10-17 1.5 10-17 2 10-17 10 15 20 25 30 35 40 45 50 parallel v=0,J"=0 perpendicular v=0,J"=0 average v=0,J"=0 average v=1,J"=0 average v=0,J"=10 Cross section (cm 2)

Photon energy (eV)

×

×

×

×

FIG. 3. 共Color online兲 Photodissociation cross section averaged over an isotropic orientation with respect to the laser field, bold line: J⬙= 0→J⬘= 1 forv = 0 with in dotted line the contribution of

the parallel orientation共1⌺→1⌺ transitions兲 and in dashed line, the contribution of the perpendicular orientations共1⌺→1⌸ transitions兲; open circles: J⬙= 10→J⬘= 9 , 11 for v = 0 ; diamonds: J⬙= 0→J⬘ = 1 forv = 1.

(5)

the same order of magnitude in the range of 38.7 eV but lower than the value for thev = 0, J

= 0 case. The cross sec-tion decreases with v in this range 共4.1⫻10−18 cm2 for v = 0, 2.0⫻10−18 cm2for v = 1 and 2.1⫻10−18 cm2for v = 4兲.

Finally, one observes that the rotational couplings are negli-gible for J

= 0 but begin to give some effect for J

= 10. We have compared the results with and without the rotational couplings for J

= 10. The cross section for parallel transi-tions differs by 0.2% but that for perpendicular transitransi-tions differs by 3%.

A comparison with previous theoretical works of the total cross section from the state v = 0, J

= 1, averaged over an isotropic orientation, is given in Fig. 4. Flower and Roueff 关22兴 consider the ground state and the first excited ⌺ state. They use the potential energy curves and the dipole moment from Green et al.关8兴 and calculate only one point which they estimate valid in an energy range of 22–31 eV. Roberge and Dalgarno关17兴 use the curves from Kolos and Peek 关9兴 and the dipole matrix element from Green et al. 关8兴 Basu and Barua关11兴 include a third ⌺ state as well as the first ⌸ state in their calculation. The agreement with our work is good until about 35 eV, where the contribution from higher excited states becomes dominant. Dumitriu and Saenz关16兴 recently calculated the cross section. At energies lower than 38 eV, our result is shifted to a larger energy by about 0.55 eV. This shift is probably due to the value of the asymptote of the ground state in Ref.关16兴 关−78.3 eV⫾0.1 eV from Fig. 1 of Ref. 关16兴 while it is −78.9 eV 共−2.898 h兲 in this work and the experimental value is −79.0 eV 共−2.903 h兲兴. Above 40 eV, our cross section decreases faster, which is due to the fact that we include only states up to n = 3 in our calculations.

Figure5共a兲shows the partial cross sections computed by Eq.共3兲 for the fragmentations leading to a section larger than 10−18 cm2. Figure 5共b兲 gives a zoom of Fig.5共a兲with sec-tions larger than 5⫻10−19 cm2. For an energy below 32 eV, the dissociation is mainly due to the H共n=1兲 channel and the region from 32 to 37 eV is dominated by the fragmentation into He共1s2s兲 共open circles兲 and H共n=2兲 共dashed line兲. We have verified that the two methods based on an asymptotic analysis 关Eq. 共3兲 共crosses兲 and Eq. 共4兲 共diamonds兲兴 give the

same partial cross sections and that their sum converges to-ward the result obtained by autocorrelation of the promoted state. As shown in the zoom in Fig. 5共b兲, the discrepancy between the total cross section关Eqs. 共1兲 and 共2兲兴 and the sum of the partial cross section is less than 10%.

The partial cross sections obtained in the Born-Oppenheimer approximation by propagating the wave pack-ets on the uncoupled adiabatic states are displayed in Fig.6. The total cross section is the same as expected but the nona-diabatic couplings modify the distribution of the fragments, particularly the yield of He共1s2s兲 and H共n=2兲 in the range around 35 eV.

The effect is more spectacular for the dissociation in the ⌸ states of the n=2 shell by the perpendicular excitation as shown in Figs. 7共a兲and7共b兲 showing the partial cross sec-tions in the BO approximation and in the coupled case. The promoted state mainly populates the lower state in this n = 2 shell which dissociates into He+and H共2p兲 fragment. The nonadiabatic coupling around R = 10.15 bohr leads to a charge exchange and a notable redistribution of the He共1s2p兲 and H共2p兲 fragments 关see Fig.7共b兲兴.

The experimental observations at 32 nm estimate a partial cross section for the fragmentation into H+ and neutral He

0 4 10-18 -18 1.2 10-17 1.6 10-17 15 20 25 30 35 40 45 50

Roberge & Dalgarno 1982 Flower & Roueff 1979 Basu & Barua 1984 Dumitriu & Saenz 2009 This work

Cross

section

(cm

2)

Photon energy (eV)

× ×

×

×

FIG. 4.共Color online兲 Comparison of the total photodissociation cross section for a isotropic orientation of the molecular axis with respect to the field for the casev = 0, J⬙= 1→J⬘= 0 , 2 with previous works. 0 1 10-18 2 10-18 3 10-18 4 10-18 5 10-18 25 30 35 40 45 50 Total Eq.(2) Sum Eq.(3) Sum Eq.(4) He(3p) H+ He+ H(n=3) He+ H(n=2) He(2s) H+ He+ H(n=1) Cross section (cm 2)

Photon energy (eV)

b 0 5 10-18 1 10-17 1.5 10-17 2 10-17 2.5 10-17 10 15 20 25 30 35 40 45 50 Total Eq.(2) Sum Eq.(3) Sum Eq.(4) He+ H(n=2) He(2s) H+ He+ H(n=1) Cross section (cm 2)

Photon energy (eV)

× × × × × (b) (a) × × × × ×

FIG. 5. 共Color online兲 Partial cross sections section computed from Eq.共3兲 for a parallel orientation for transitions J⬙= 0→J⬘= 1 from the vibrational statev = 0. The sum of the partial cross sections

is compared with that obtained by Eq.共4兲 and with the total cross

section calculated by Eq. 共2兲. Panel a: sections larger than

(6)

equal to 共1.4⫾0.7兲⫻10−18 cm2. In Fig. 8, we sum all the partial sections for channels H++ He共1snl兲 with nⱖ2 and give the average over isotropic distribution. Figure8共a兲

pre-sents the BO approach and Fig. 8共b兲 gives the results ob-tained with the coupled channels. The BO profile is similar to the simulation of Ref.关16兴 with a lower maximum value near 40 eV for the contribution of the⌸ states. The coupled-channel profile is very different, particularly due to the peak in between 32 and 33 eV which comes from the fragmenta-tion into He共1s2p兲 from the perpendicular orientafragmenta-tion 关see Fig. 7共b兲兴. The peak at 39 eV has a maximum of 3.14 ⫻10−18 cm2and the value at the experimental wavelength is 2.07⫻10−18 cm2, thus it belongs to the uncertainty bar of the experimental data共1.4⫾0.7⫻10−18 cm2兲. In conclusion, our BO simulation does not agree with the experimental value but the result of the nonadiabatic coupled equations gives good agreement. Furthermore, the simulation confirms that the major part of the fragmentation arise from the ex-cited⌸ states. This is proven again below by the simulation of the photodissociation with a femtosecond laser pulse.

IV. SIMULATION WITH A FEMTOSECOND LASER PULSE

The FEL experiment analyzes the neutral He共1snl兲 frag-ment by resolving the kinetic energy release distribution and

0 1 10-18 2 10-18 3 10-18 4 10-18 5 10-18 25 30 35 40 45 50

Sum of the partial cross sections

He+ H(n=1) He(2s) H+ He+ H(n=2) He(2p) H+ He+ H(n=2) He(3d) H+ Cross section (cm 2)

Photon energy (eV)

× × × ×

×

FIG. 6. 共Color online兲 Partial cross sections larger than 5 ⫻10−19 cm2computed from Eq.3兲 in the Born-Oppenheimer

ap-proximation for a parallel orientation for transitions J⬙= 0→J⬘= 1 from the vibrational statev = 0. The sum of the partial cross sections

BO 共full line兲 is compared with that the results obtained in the nonadiabatic case共triangles兲.

0 5 10-18 1 10-17 1.5 10-17 2 10-17 2.5 10-17 25 30 35 40 45 50

Sum of the partial cross sections

He+ H(n=2) He+ H(n=3) He(3p) H+ Cross section (cm 2)

Photon energy (eV) BO × × × × × 0 5 10-18 1 10-17 1.5 10-17 2 10-17 2.5 10-17 25 30 35 40 45 50

Sum of the partial cross sections He(3d) H+ He+ H(n=3) He(2p) H+ He+ H(n=2) Cross section (cm 2)

Photon energy (eV) non BO × × × × × (b) (a)

FIG. 7. 共Color online兲 Partial cross sections larger than 10−18 cm2computed from Eq. 3兲 for a perpendicular orientation

for transitions J⬙= 0→J⬘= 1 from the vibrational state v = 0. Panel

a: Born-Oppenheimer approximation; panel b: nonadiabatic case.

0 2 10-19 4 10-19 6 10-19 8 10-19 1 10-18 1.2 10-18 1.4 10-18 25 30 35 40 45 50 Parallel Perpendicular Average Cross section (cm 2)

Photon energy (eV) BO × × × × × × × 0 1 10-18 2 10-18 3 10-18 4 10-18 5 10-18 6 10-18 7 10-18 25 30 35 40 45 50 Parallel Perpendicular Average Cross section (cm 2)

Photon energy (eV) non BO × × × × × × × (b) (a)

FIG. 8. 共Color online兲 Average partial cross section for the frag-mentation into H+ and neutral He computed from Eq. 3兲 for an

isotropic orientation for transitions J⬙= 0→J⬘= 1 from the vibra-tional statev = 0. Panel a: Born-Oppenheimer approximation; panel

(7)

the dissociation angle distribution relative to the FEL polar-ization which is directed along the ion beam. The FEL con-sists of a sequence of short pulses of about 30 fs separated by 10 ␮s. The accumulation time to get the photo fragment image is very long of about 13 h 关23兴. We simulate the im-pact of a single laser shot of 30 fs and an intensity which is probably larger than the experimental one but leading to transition probabilities of the order of 10−3. The initial state is thev = 0 J = 0 state of the ground electronic state. Figure9 gives an example of time-dependent occupancies of the elec-tronic states关Eq. 共5兲兴 during the photodissociation for paral-lel共⌺ states兲 and perpendicular 共⌸ states兲 orientations. Only fragments having an asymptotic probability larger than 3 ⫻10−4are shown. One observes a larger probability of popu-lating states dissociating into H fragment and a larger prob-ability to obtain He fragment from the⌸ states than from the ⌺ states in agreement with the experimental results.

We have computed the kinetic energy release distribution of the fragment He共1snl兲 by Fourier transforming the wave packet in the asymptotic region. The distribution averaging over an isotropic orientation is given in Fig.10. We obtain a large intensity in between 13.5 and 14 eV and a second less intense peak at 16.5 eV. The parallel contribution in the range 13.5 and 14 eV is due to He共1s3s兲 and He共1s3p兲 and that at 16.5 eV to He共1s2s兲. The large perpendicular contri-bution at 13.5 eV is mainly due to He共1s3p兲. The simulation

confirms that the major part arises from the excited⌸ states. Finally, we simulate the angular distribution of the He fragment for the peak around 13.5 eV. As discussed in Ref. 关23兴 when the fragmentation is very short 共here less than 100 fs兲 the fragmentation angle is related to the orientation of the molecule at the time of the excitation. The angular distribu-tion is computed by the relation 关40兴, I共␪兲=I0关1 +␤P2共cos␪兲兴/4␲ with P2共cos␪兲=关3 cos2␪− 1兴/2. We take for I0 the maximum value of the peak at 13.8 eV in Fig.10 for the parallel 共␤= 2兲 and the perpendicular 共␤= −1兲 orien-tation, respectively. The distribution is given in Fig.11. This result can be compared with the experimental data shown in Fig. 4c of Ref.关23兴. However, the latter figure corresponds to a summation over the energy released fragments from 13 to 20 eV including therefore the large contribution at 16.5 eV which is mainly due to an⌺ state. This increases the parallel contribution on the experimental graphic when compared with our Fig. 11. Nevertheless, the two figures agree

quali-0 0.002 0.004 0.006 0.008 0.01 0 20 40 60 80 100 Probability Time (fs) perpendicular He+H(n=2) He+H(n=3) He+H(n=3) He(2p) H+ He(3p) H+ He(3d) H+ (b) (a) 0 0.0005 0.001 0.0015 0.002 0.0025 0 20 40 60 80 100 P ro b a bili ty Time (fs) He(2s) H+ He+H(n=2) He(2p) H+ He(3p) H+ He(3s) H+ He(3d) H+ He+H(n=2) He+H(n=3) He+H(n=3) He+H(n=3) parallel

FIG. 9.共Color online兲 Population in the excited electronic states during the photodissociation of the ground v = 0, J = 0 state by a

pulse of 30 fs and a maximum intensity of 8.775⫻1011 W cm−2.

Panel共a兲 parallel orientation 共Only fragments having an asymptotic probability larger than 3⫻10−4are shown兲; panel 共b兲 perpendicular

orientation. 0 0.01 0.02 0.03 0.04 0.05 0.06 12 13 14 15 16 17 18 Parallel Perpendicular Average

Photon energy (eV)

Intensity

(arbitrary

units

)

FIG. 10. 共Color online兲 Kinetic energy release for the fragment He共1snl兲 during the photodissociation of the ground v=0, J=0 state by a pulse of 30 fs. The distribution is averaged over an isotropic orientation. 0 0.001 0.002 0.003 0.004 0.005 0.006 0 30 60 90 120 150 180 Parallel Perpendicular Dissociation angle Intensity (arbitrary units )

FIG. 11. 共Color online兲 Angular distribution of the He共1snl兲 fragment for a kinetic energy release in the range 13.5–14 eV dur-ing the photodissociation of the groundv = 0, J = 0 state by a pulse

of 30 fs. The intensity I0 is taken from the contribution of the

(8)

tatively and present a clear dominance of the⌸ states 共per-pendicular contribution兲 over the ⌺ states 共parallel contribu-tion兲 in both cases.

V. CONCLUSION

We have computed the total and partial photodissociation cross sections of the molecular ion HeH+ for fragmentation into the excited shells n = 1 , 2 , 3 up to a photon energy of 45 eV. We have analyzed the role of the nonadiabatic interac-tions and of the rotational couplings. The ab initio challenge was to get relevant potential energy curves and nonadiabatic couplings for very excited states. The simulations using the Born-Oppenheimer or diabatic potentials give the same total cross section and are in good agreement with the simulation of Dumitriu et al.关16兴. However for the partial cross section in a given channel or the branching ratio between H and He fragmentation, we have illustrated that the simulation ob-tained in the framework of the Born-Oppenheimer model is not sufficient even if it can give accidentally a good agree-ment with experiagree-ment for some energies. The effect of nona-diabatic couplings is more dramatic for perpendicular disso-ciation via the ⌸ states in the range 30–35 eV, i.e., for dissociation toward n = 1 and n = 2 shells. The branching ratio between fragments He or H in the n = 2 shell cannot be pre-dicted by neglecting the⌸ states or by neglecting the nona-diabatic couplings.

By considering a highly excited rotational initial state共J = 10兲, we have observed that the rotational couplings do not play a significant role and could be neglected for lower ro-tational quantum numbers for photodissociation. However,

rotational excitation could play a significant role because the profile of the total cross section strongly depends of the ini-tial value of J, mainly for dissociation into the n = 1 shell. Important information for further simulations is the weight of rotationally or vibrationally excited initial states for non-Boltzmann situations.

Furthermore, time-dependent simulations with very accu-rate methods taking into account couplings among excited states find a renewed interest since the advent of ultrashort laser pulses with energetic photons even for small systems. The time-dependent simulation with a femtosecond Gaussian laser field has provided results which are qualitatively in agreement with the recent FEL laser experiment concerning the order of magnitude of the partial cross section of He neutral fragment at 32 nm and the distribution of the kinetic energy release. We have confirmed the experimental obser-vation that dissociation through⌸ states dominates the pro-cess in the probed energy range.

ACKNOWLEDGMENTS

We thank Dr. B. Lavorel for a helpful discussion. The computing facilities of IDRIS共Projects No. 061247 and No. 2006 0811429兲 as well the financial support of the FNRS in the University of Liège SUN Nic2 project are gratefully ac-knowledged. We thank the support of the COST Action CM0702 CUSPFEL. J.L. and N.V. acknowledge the financial support of the Fond National de le Recherche Scientifique 共IISN projects兲 as well as the “Action de Recherche Con-certée” ATMOS. K.S. would like to thank The University of Paris-Sud 11 for financial support Grant No. 2238. J.L. thanks the FRIA for financial support.

关1兴 T. R. Hogness and E. C. Lunn, Phys. Rev. 26, 44 共1925兲. 关2兴 Z. Liu and P. B. Davis, J. Chem. Phys. 107, 337 共1997兲. 关3兴 W. J. van der Zande, W. Koot, D. P. de Bruijn, and C. Kubach,

Phys. Rev. Lett. 57, 1219共1986兲.

关4兴 D. Strasser, K. G. Bhushan, H. B. Pedersen, R. Wester, O. Heber, A. Lafosse, M. L. Rappaport, N. Altstein, and D. Zajf-man, Phys. Rev. A 61, 060705共2000兲.

关5兴 F. B. Yousif, J. B. A. Mitchell, M. Rogelstad, A. Le Paddelec, A. Canosa, and M. I. Chibisov, Phys. Rev. A 49, 4610共1994兲. 关6兴 Y. Susuki, T. Ito, K. Kimura, and M. Mannami, Phys. Rev. A

51, 528共1995兲.

关7兴 H. H. Michels, J. Chem. Phys. 44, 3834 共1966兲.

关8兴 T. A. Green, H. H. Michels, J. C. Browne, and M. M. Madsen, J. Chem. Phys. 61, 5186共1974兲; A. Green, H. H. Michels, and J. C. Browne, ibid. 64, 3951共1976兲; 69, 101 共1978兲. 关9兴 W. Kolos and J. M. Peek, Chem. Phys. 12, 381 共1976兲. 关10兴 S. Saha, K. K. Datta, and A. K. Barua, J. Phys. B 11, 3349

共1978兲.

关11兴 D. Basu and A. K. Barua, J. Phys. B 17, 1537 共1984兲. 关12兴 A. Saenz, Phys. Rev. A 67, 033409 共2003兲.

关13兴 M. Pavanello, S. Bubin, M. Molski, and L. Adamowicz, J. Chem. Phys. 123, 104306共2005兲.

关14兴 B.-L. Zhou, J.-M. Zhu, and Z.-C. Yan, Phys. Rev. A 73, 064503共2006兲.

关15兴 J. Fernández and F. Martin, J. Phys. B 40, 2471 共2007兲. 关16兴 I. Dumitriu and A. Saenz, J. Phys. B 42, 165101 共2009兲. 关17兴 W. Roberge and A. Dalgarno, Astrophys. J. 255, 489 共1982兲. 关18兴 S. Lepp and J. M. Shull, Astrophys. J. 280, 465 共1984兲. 关19兴 D. Galli and F. Palla, Astron. Astrophys. 335, 403 共1998兲. 关20兴 E. A. Engel, N. Doss, G. J. Harris, and J. Tennyson, Mon. Not.

R. Astron. Soc. 357, 471共2005兲.

关21兴 P. C. Stancil, S. Lepp, and A. Dalgarno, Astrophys. J. 509, 1 共1998兲.

关22兴 D. R. Flower and E. Roueff, Astron. Astrophys. 72, 361 共1979兲.

关23兴 H. B. Pedersen, S. Altevogt, B. Jordon-Thaden, O. Heber, M. L. Rappaport, D. Schwalm, J. Ullrich, D. Zajfman, R. Treusch, N. Guerassimova, M. Martins, J.-T. Hoeft, M. Wellhöfer, and A. Wolf, Phys. Rev. Lett. 98, 223202共2007兲.

关24兴 M. Shapiro and R. Bersohn, Annu. Rev. Phys. Chem. 33, 409 共1982兲.

关25兴 P. S. Shternin and O. S. Vasyutinskii, J. Chem. Phys. 128, 194314共2008兲 and references herein.

关26兴 J. M. Brown, J. T. Hougen, K.-P. Huber, J. W. C. Johns, I. Kopp, H. Lefebvre-Brion, A. J. Mere, D. A. Ramsey, J. Rostas, and R. N. Zare, J. Mol. Spectrosc. 55, 500共1975兲.

关27兴 R. N. Zare, Angular Momentum 共World Scientific, New York, 1988兲.

(9)

关28兴 P. J. Knowles and H.-J. Werner, Chem. Phys. Lett. 115, 259 共1985兲; H.-J. Werner and P. J. Knowles, J. Chem. Phys. 82, 5053共1985兲.

关29兴 H.-J. Werner, P. J. Knowles, R. D. Amos, A. Bernhardsson, A. Berning, P. Celani, D. L. Cooper, M. J. O. Deegan, A. J. Dob-byn, F. Eckert, C. Hampel, G. Hetzer, T. Korona, R. Lindh, A. W. Lloyd, S. J. McNicholas, F. R. Manby, W. Meyer, M. E. Mura, A. Nicklass, P. Palmieri, R. Pitzer, G. Rauhut, M. Schütz, H. Stoll, A. J. Stone, R. Tarroni, and T. Thorsteinsson,

MOLPRO is a package of ab initio programs, Department of Chemistry, University of Birmingham, Birmingham B15 2TT, UK.

关30兴 R. A. Kendall, T. H. Dunning, Jr., and R. J. Harrison, J. Chem. Phys. 96, 6796 共1992兲; D. E. Woon and T. H. Dunning Jr.,

ibid. 98, 1358共1993兲.

关31兴 J. Loreau, P. Palmeri, P. Quinet, J. Liévin, and N. Vaeck 共un-published兲.

关32兴 C. Zhu and H. Nakamura, J. Chem. Phys. 106, 2599 共1997兲. 关33兴 A. Hansson and J. K. G. Watson, J. Mol. Spectrosc. 233, 169

共2005兲.

关34兴 E. J. Heller, J. Chem. Phys. 68, 3891 共1978兲; 68, 2066 共1978兲. 关35兴 D. J. Tannor, Introduction to Quantum Mechanics A

Time-dependent Perspective 共University Science Books, Sausalito,

California, 2007兲.

关36兴 G. G. Balint-Kurti, R. N. Dixon, and C. C. Marston, J. Chem. Soc., Faraday Trans. 86, 1741共1990兲.

关37兴 K. C. Kulander and E. J. Heller, J. Chem. Phys. 69, 2439 共1978兲.

关38兴 M. D. Feit, J. A. Fleck, and A. Steiger, J. Comput. Phys. 47, 412共1982兲; M. D. Feit and J. A. Fleck, J. Chem. Phys. 78, 301 共1983兲.

关39兴 J. Alvarellos and H. Metiu, J. Chem. Phys. 88, 4957 共1988兲. 关40兴 R. N. Zare, Mol. Photochem. 4, 1 共1972兲.

Figure

TABLE I. CASSCF asymptotic energies of the 1 ⌺ + and 1 ⌸ states included in the calculations.
Figure 3 gives the photodissociation cross section for three cases of Fig. 2 averaged over an isotropic orientation by weighting each orientation x, y, z by 1/3
Figure 5共a兲 shows the partial cross sections computed by Eq. 共3兲 for the fragmentations leading to a section larger than 10 −18 cm 2
FIG. 6. 共 Color online 兲 Partial cross sections larger than 5
+2

Références

Documents relatifs

Supporting the Time Resolved data with Molecular Dynamics simulations, static diraction data and the experimental study of the pure solvent response to ultrafast heating,

(52%) des répondants sont toujours clients dans les banques publiques malgré qu’ils soient clients dans les banques privées, les réponses de la prochaine

ABSTRACT - In this paper we consider evolution inclusions driven by a time de- pendent subdifferential operator and a set-valued perturbation term.. For the nonconvex

In order to solve the finite dimensional variational problems, in [14] the authors implement the primal-dual algorithm introduced by Chambolle and Pock in [20], whose core

The main purpose of this paper is to give stability analysis and error estimates of the local discontinuous Galerkin (LDG) methods coupled with three specific implicit-explicit

tions).. We consider the time dependent potential without the bounded part v2, which is a multiplication operator with the function. where denotes a permitted family of

While it is satisfying to know that the bang-bang version of the control law is also the optimal control for an infinite horizon problem in the case of one mode, it is not at all

— Well posedness in the Gevrey Classes of the Cauchy Problem for a Non Strictly Hyperbolic Equation with Coefficients Depending On Time, Ann.. — Two by two strongly hyperbolic systems