Electronic and Vibrational Spectroscopy of CsS - ACS Publications

(MRCI+Q) with and without considering spin-orbit effects. The shapes of the PECs are governed by the interactions between the two ionic states, 2Σ+ a...
0 downloads 0 Views 298KB Size
Subscriber access provided by UNIV OF NEW ENGLAND ARMIDALE

Article

Electronic and Vibrational Spectroscopy of CsS Karim Elhadj Merabti, Sihem Azizi, Roberto Linguerri, Gilberte Chambaud, Muneerah Mogren Al-Mogren, and Majdi Hochlaf J. Phys. Chem. A, Just Accepted Manuscript • DOI: 10.1021/acs.jpca.8b02543 • Publication Date (Web): 29 May 2018 Downloaded from http://pubs.acs.org on May 29, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 19 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

The Journal of Physical Chemistry

Electronic and Vibrational Spectroscopy of CsS Karim Elhadj Merabti,1 Sihem Azizi,1 Roberto Linguerri,2 Gilberte Chambaud,2 Muneerah Mogren Al-Mogren,3 and Majdi Hochlaf 2,* 1

Laboratoire de Physique Théorique, Université Abou Bekr Belkaid Tlemcen, Algeria.

2

Université Paris-Est, Laboratoire Modélisation et Simulation Multi Echelle, MSME UMR 8208

CNRS, 5 bd Descartes, 77454 Marne-la-Vallée, France. 3

Chemistry Department, Faculty of Science, King Saud University, PO Box 2455, Riyadh 11451,

Kingdom of Saudi Arabia.

1 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

Abstract Using multi configurational ab initio methodologies, we compute the potential energy curves (PECs) of the lowest electronic states of the diatomic CsS. These computations are performed using internally contracted multi reference interaction configuration including Davidson correction (MRCI+Q) with and without considering spin-orbit effects. The shapes of the PECs are governed by the interactions between the two ionic states, 2Σ+ and 2Π, correlating at large internuclear separations (RCsS) to the first ionic dissociation limit and the other electronic states correlating to the three lowest neutral dissociation limits. Computations show the importance of considering large amount of electron correlation for the accurate description of the PECs and spectroscopy of this molecular system. As expected, these PECs are also strongly affected by the spin-orbit interaction. For the bound states, we report a set of spectroscopic parameters including equilibrium distances, dissociation energies, vibrational and rotational constants. The effects of spin-orbit-induced changes on these parameters are also discussed. Moreover, we show that the 22Π state presents a “bowl” potential with a rather flat region extending to large RCsS distances. After being promoted to this state, wavepackets should undergo strong oscillations, similar to those observed by Zewail and co-workers for the NaI molecule. These should provide information on the shape of the PEC for the 22Π state and also on the couplings between this and the neighbouring states.

2 ACS Paragon Plus Environment

Page 2 of 19

Page 3 of 19 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

The Journal of Physical Chemistry

I.

Introduction Cesium, Cs, being the alkali element with the lowest (albeit non-negligible) ionization energy

(3.89 eV), can be regarded as a good source material for electrons in plasma heating modules. Such media have complex chemical compositions inducing a variety of physical and chemical processes, where cesium chalcogenides may act as impurities. For instance, CsO is listed as possible impurity in Tokamaks using Cs grids, due to the presence of oxygen traces in the heating chamber

1,2

. Traces of

sulfur and sulfur bearing species can also exist there. After reaction with cesium compounds, they probably result in the formation, for example, of CsS or Cs2S. Moreover, it is known that cesium nitride, Cs3N, can be readily attacked by sulfur 3. Upon heating CsO (for T > 150-200° C) in the presence of dry sulfur dioxide, incandescence is observed 3. Again, molecular species such as CsS or Cs2S, may be produced during these reactions and contribute to the observed effects. To understand the phenomena taking place during these reactions and in plasma, one needs a full characterization of the involved atomic and molecular species. Since alkali-metal monoxides are commonly found in many high temperatures systems, they have been the subject of many studies, both experimental and theoretical

4-16

. However their sulfur isovalent analogues, i.e. alkali-metal

monosulfides (MS), were investigated 17,18, but not to the same extent, consequently there is a lack of information on these molecules. For instance, spectroscopic data on CsS, of interest in the present study, are limited to the determination of the nature of its lowest lying electronic states by Partridge and al. 17 and by Lee and Wright 18. They deduced spectroscopic constants for the ground state, of 2Π symmetry, and for the first close-lying excited state of 2Σ+ symmetry. Both states are coupled by spinorbit interaction. Nevertheless, no information is available on the upper excited electronic states. To date, no experimental studies have been performed on the spectroscopy of gas phase CsS. However, accurate data can be provided by pure theoretical treatments. The goal of this work is to characterize the electronic ground and low-lying excited states of CsS using high-level ab initio calculations. Here we investigate the CsS molecule using multi reference post Hartree-Fock methods and large basis sets. We first compute the CsS potential energy curves (PECs) with and without taking into account the spin-orbit interaction. In addition to the X2Π and 12Σ+ states, we found another bound electronic excited state, the 22Π state. The shapes of all these potentials are strongly affected by considering the spin-orbit interaction and electron correlation. Then, we derived a set of spectroscopic constants that may be used for the identification of CsS in plasma or in laboratory.

II.

Computational methods In this work we study the low-lying doublet and quartet electronic states of CsS. The Cs atom

is described using the ECP46MDF relativistic pseudo-potential, where the 9 external electrons are described by [13s,11p,9d,10f,5g] functions 8. For S, after carefully checking the performance of the

3 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

aug-cc-pVXZ (X=Q, 5, 6) basis sets

19,20

Page 4 of 19

, it was decided to use the aug-cc-pV5Z one. The electronic

computations are performed using the MOLPRO program suite 21. Since MOLPRO does not consider infinite point groups, these calculations are done in the C2v point group where the B1 and B2 representations were equivalently described. Through benchmark calculations, we found out that the shapes and relative position of the CsS PECs are very sensitive to the level of accuracy of the electron correlation description (see below). Thus one needs to use strongly-correlated post-Hartree-Fock methods for a correct characterisation of the CsS electronic states. Also, we found that several of these electronic states exhibit a pronounced multi-configurational character. Hence, this invalidates the use of single reference electronic structure methods for their investigation. Here, we used the state-averaged complete active space self-consistent field (CASSCF)

22,23

followed by multireference configuration interaction (MRCI) techniques

The Davidson correction (MRCI+Q) was also added to the MRCI energies

27

24-26

.

. The CASSCF active

space employed in this study is composed of the 3s, 3p atomic orbitals of S and 6s, 6p orbitals of Cs. This ansatz leads to 7 valence electrons distributed in 8 valence orbitals. In the CASSCF computations we averaged the doublets and quartets together using the default averaging procedure of MOLPRO. In the MRCI step, the complete CASSCF wave functions were used as a reference. For the treatment of spin-orbit interactions, the effective Breit-Pauli SO operator, HSO, as implemented in MOLPRO was used 28. The PECs were later incorporated into variational treatments of the nuclear motions. The spectroscopic parameters were obtained by variational computations using the Numerov package

29

where the nuclear motion problem is solved with the method of Cooley 30.

III.

Results and discussion a. On the nature of the electronic states of CsS

Table 1: Dissociation channels for CsS and molecular states that correlate to these asymptotes. We give also the energy positions (in eV) of the asymptotes. Dissociation fragments

Experimental relative energies (eV)

a)

Calculated relative energies (eV) b)

Cs(2S)+S(3P2)

0.00 c)

0.00

Cs(2S)+S(1D)

1.14

1.16

Cs(2P01/2)+S(3P2)

1.38

1.37

Cs+(1S)+S-(2P3/2)

1.82

2.00

a) Ref. 41. b) MRCI+Q-SO/aug-cc-pV5Z level of theory. c) Used as reference. 4 ACS Paragon Plus Environment

Molecular states 2,4 2 2,4

(Σ−,Π)

(Σ+,Π,∆)

(Σ+,Σ−(2),Π(2),∆) 2

(Σ+,Π)

Page 5 of 19 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

The Journal of Physical Chemistry

The electronic states investigated in the present contribution are correlating at large internuclear distances to [Cs(2S) + S(3P)], [Cs(2S) + S(1D)], [Cs(2P) + S(3P)] and [Cs+(1S) + S-(2P)] dissociation limits (see Table 1).

Figure 1: Potential energy curves of the X2Π, 22Π (solid lines) and 12Σ+ (dashed lines) states of CsS as computed at the CASSCF/aug-cc-pV5Z (red lines), MRCI/aug-cc-pV5Z (black lines) and MRCI+Q/aug-cc-pV5Z (blue lines) levels. The reference energy is chosen to be the energy of CsS (X2Π) at equilibrium.

Figure 1 presents the CASSCF, MRCI and MRCI+Q PECs of the three lowest electronic states of CsS, X2Π, 22Π and 12Σ+, computed with the aug-cc-pV5Z basis set. These electronic states correlate adiabatically to the [Cs(2S)+S(3P)] and [Cs(2S)+S(1D)] dissociation limits at large internuclear separations (Table 1). While the three sets of PECs have smooth shapes and gently reach their limits, we observe drastic differences between the PECs computed using CASSCF and those derived at the MRCI and MRCI+Q levels. This shows the importance of considering dynamical electron correlation (within MRCI) for the accurate description of the electronic states of CsS. Since there is an interplay between covalent (Cs … S) and ionic (Cs+ … S-) bonding character for the formation of these electronic states, we assume that the Cs+ … S- interaction is poorly described at the CASSCF level and only MRCI(+Q) computations recover such effects.

5 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

Page 6 of 19

Figure 2: MRCI+Q/aug-cc-pV5Z potential energy curves of CsS along the internuclear separation (RCsS). The reference energy is chosen to be the energy of CsS (X2Π) at equilibrium.

The nature of the bonding and of the ground electronic state of alkali chalcogenides (MX) was subject to several studies. Indeed, the lowest 2Σ+ and 2Π states are close-lying and their relative ordering, and hence the nature of the electronic ground state of MX, depends on the balance between long-range attractive electrostatic terms that favour the 2Π state, and Pauli repulsion terms that tend to favour the 2Σ+ state at shorter ranges. For instance, alkali-metal monoxides exhibit a change in the electronic ground state when descending the alkali column: LiO

31,32

, NaO

33

and KO

34

have a 2Π

ground state, while RbO 35 and CsO 12 possess a 2Σ+ ground state. For alkali-metal monosulfides, Lee and Wright established a 2Π ground state whatever the nature of the alkali 18. Our results on the two lowest 2Σ+ and 2Π states of CsS confirm these findings, as shown in Figure 2 by the MRCI+Q/aug-ccpV5Z potential energy curves along the Cs-S bond length (RCsS).

6 ACS Paragon Plus Environment

Page 7 of 19 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

The Journal of Physical Chemistry

In addition to those of the X2Π and 12Σ+ states, Figure 2 displays the PECs of the other electronic states correlating at large internuclear distances to [Cs(2S) + S(3P)], [Cs(2S) + S(1D)] and [Cs(2P) + S(3P)] asymptotes (Table 1). An interesting pattern in this figure is the behaviour of the 2Π and 2Σ+ states which are intercepted at long distance by the 2Π and 2Σ+ ionic states correlating to the [Cs+(1S) + S-(2P)] limit, which is the fourth dissociation limit (given in Table 1 but not shown in Figure 2). Figure 2 shows that among the electronic states correlating to the [Cs(2S) + S(3P)] lowest limit, only the ground state presents a deep potential well whereas we compute a repulsive potential (for 14Π) or small potential wells for the other states (i.e. 12Σ- and 14Σ-). For RCsS ~ 5.75 Å, the 12Σstate is crossed by the 12Σ+, where further interaction may take place. Since 12Σ- leads directly to the first dissociation limit, the ro-vibrational levels of 12Σ+ located above this limit are expected to be predissociated by this state, resulting in a shortening of their lifetimes. Additional mutual interactions can be found after inspection of Figure 2. For energies > 3 eV, the situation is even more complicated because of the high density of states correlating to the [Cs(2S) + S(1D)] and [Cs(2P) + S(3P)] dissociation limits. In these series, only 12Σ+ and 22Π states present potential wells greater than 1 eV. The other states present either shallow potentials or repulsive ones. It is worth noting the existence of several avoided crossings between the states of the same space and spin multiplicities. For states with the same spin multiplicity, we also observe crossings between Σ - Π and Π - ∆ doublets where vibronic and also spin-orbit couplings may take place as discussed above for 12Σ+ and 12Σ-. It is worth noting that we have previously computed similar patterns and shapes for the isovalent NaS diatomic 36.

Table 2: CASSCF and MRCI+Q/aug-cc-pV5Z vertical excitation energies (Tv, in eV) and dissociation energies (De, in eV) of the lowest electronic states of CsS. Between parentheses, we give the weight of the corresponding dominant configurations as quoted at the equilibrium distance of CsS (X2Π). Tv

Electronic state

De

Dominant electronic configuration

CASSCF MRCI+Q CASSCF MRCI+Q

X2Π

0.00

0.00

0.72

2.72

(0.88)7σ2 3π3

12 Σ +

0.19

0.15

1.69

3.79

(0.92)7σ1 3π4

14 Σ -

1.22

3.05

(0.99)7σ2 3π2 8σ1

14Π

1.26

3.09

(0.99)7σ1 3π3 8σ1

12 Σ -

1.27

3.10

(0.99)7σ2 3π2 8σ1

22Π

1.44

3.17

12 ∆

2.33

4.22

(0.99)7σ1 3π3 4π1

24Π

2.23

4.27

(0.99)7σ2 3π2 4π1

32Π

2.32

4.30

(0.85)7σ2 3π2 4π1

24 Σ -

2.31

4.31

(0.94)7σ2 3π2 9σ1

0.65

0.99

(0.92)7σ1 3π3 8σ1

7 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

22 ∆

2.65

4.33

(0.99)7σ2 3π2 8σ1

22 Σ +

2.47

4.36

(0.86)7σ1 3π3 4π1

22 Σ -

2.36

4.36

(0.65)7σ1 3π3 4π1 & (0.35)7σ2 3π2 9σ1

34Π

2.39

4.38

(0.99)7σ1 3π3 9σ1

32 Σ -

2.39

4.37

(0.35)7σ1 3π3 4π1 & (0.65)7σ2 3π2 9σ1

34 Σ -

2.36

4.40

(0.94)7σ1 3π3 4π1

42Π

2.50

4.42

(0.85)7σ1 3π39σ1

Page 8 of 19

Table 2 shows the calculated vertical transition energies for the lowest electronic states of CsS together with the dissociation energies for the bound states, obtained at the CASSCF/aug-cc-pV5Z and MRCI+Q/aug-cc-pV5Z levels. We again refer to the in-depth analysis done by Lee and Wright

18

for

the 2Σ+ - 2Π energy ordering. Table 2 lists also the dominant electron configurations of CsS states as quoted at the equilibrium geometry of CsS (X2Π). In the molecular region, the dominant electron configurations of the two lowest electronic states (X2Π and 12Σ+) correspond to occupation of the outermost 7σ or 3π molecular orbitals (MOs) which are mainly located on the sulfur atom, giving a Cs+S- character to these states. For the other states, the dominant configurations correspond mostly to the promotion of one electron from the 7σ or 3π MOs into the vacant 8σ or 9σ or 4π MOs which have more Cs character, resulting in quasi covalent bonds.

8 ACS Paragon Plus Environment

Page 9 of 19 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

The Journal of Physical Chemistry

Figure 3: Potential energy curves of the lowest 2Π states of CsS along the internuclear separation (RCsS) and their dipole moments. The reference energy is chosen to be the energy of CsS (X2Π) at equilibrium.

At medium and large internuclear distances, the wave functions of these electronic states are strongly mixed because of their mutual interactions. For instance, one can identify several avoided crossings between the states presenting the same spin and space symmetry, such as the one between the X2Π and 22Π for RCsS ~ 6 Å and between 22Π and 32Π for RCsS ~ 8 Å (Figure 3). These avoided crossings result in an unconventional shape for the 22Π state. Instead of the usual Morse-like shape, it has a “bowl” profile with a flat region extending from RCsS ~ 3.5 up to ~ 6 Å. This potential shape is interesting from an applicative point of view. For instance, Zewail and co-workers showed that there are two limiting possibilities when a wavepacket is promoted into a potential of this kind: (i) either it is trapped on the adiabatic potential without crossing (resonance) as for NaI 37, or (ii) it jumps on the diabatic potential as for ICN

38-40

. These two limiting cases give rise to different temporal evolutions.

The observation of oscillations gives access to the shape of the potentials of the states and to their mutual couplings. CsS (22Π) falls in case (i) so we expect wavepacket oscillations (resonance) during the dissociation reactions. This can be confirmed experimentally using femtosecond spectroscopy. Since the X2Π-22Π excitation energy is about 3 eV, this transition occurs in the visible for which fslasers are available in laboratories. Figure 4 shows that an avoided crossing takes place between the 12Σ+ and 22Σ+ state at large internuclear distance (RCsS ~ 7.2 Å). Again, this avoided crossing is responsible for an unusual shape of the upper state. 9 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

Figure 4: Potential energy curves of the lowest 2Σ+ states of CsS along the internuclear separation (RCsS) and their dipole moments. The reference energy is chosen to be the energy of CsS (X2Π) at equilibrium.

10 ACS Paragon Plus Environment

Page 10 of 19

Page 11 of 19 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

The Journal of Physical Chemistry

Close to equilibrium geometry of the X2Π state, examination of the dipole moment of CsS (X2Π) (Figure 3) shows that the Cs - S bond is polarized, corresponding to Csq+ - Sq- polarization with a value of q close to unity. Indeed, a Mulliken population analysis shows that q = 0.75 at the MRCI level. On the contrary, for the 22Π state the bond is mostly covalent between Cs and S for this geometry. As in the X2Π state, a large dipole moment is computed for 12Σ+ (Figure 4), signature of the ionic character of the C - S bond in this electronic state.

Figure 5: MRCI+Q/aug-cc-pV5Z potential energy curves of the Ω states of CsS calculated after considering the spin-orbit interactions along the internuclear distance (RCsS). The calculation of the energies of the spin-orbit states has been performed at the MRCI+Q level, where all the vectors correlating to the three lowest dissociation asymptotes from a previous state-averaged CASSCF computation have been used for the multi-reference configuration interaction treatment. The MRCI+Q energies have been taken as diagonal terms in the HSO matrices instead of the

11 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

Page 12 of 19

MRCI energies. This ansatz accounts for 19 Λ-Σ electronic states resulting in 82 Kramers doublets corresponding to 41 fine structure states (with Ω =1/2, 3/2, 5/2 and 7/2). Figure 5 presents the Ω PECs of CsS, in the 0 – 4 eV energy domain, after considering the spin-orbit interaction. Mostly, this figure displays the Ω (= 1/2(5), 3/2(3) and 5/2) components arising from the X2Π, 14Σ-, 12Σ- and 14Π states which correlate to the [Cs(2S1/2) + S(3P2,1,0)] dissociation limit and the Ω (= 1/2(2) and 3/2) components arising from the 12Σ+ and 22Π states which correlate to the [Cs(2S1/2) + S(1D2)] one. For clarity we have not reported in this figure the Ω (= 3/2 and 5/2) components arising from the 2∆ state, also correlated to the second asymptote, since their shapes are very regular. The most interesting feature is the avoided crossing, due to spin-orbit coupling, between the Ω= 1/2 states at ~6 Å. This interaction results in the correlation of the Ω= 1/2 component arising from the 12Σ+ state to the lowest dissociation limit. To have a further insight into the effect of the spin-orbit interaction on the investigated states we give as supplementary material (Tables S1 and S2) the composition of the spinorbit states in terms of the 82 Kramer doublets originating from the manifold of 19 Λ-Σ states, at RCS=3 Å, i.e. close to the ground-state equilibrium geometry, and RCS=6 Å. Using MRCI+Q/aug-ccpV5Z calculations, we compute the following splittings for S(3Pj) states: S(3P2) = 0 eV; S(3P1) = 0.0495 eV and S(3P0) = 0.0720 eV, which compare very well with the experimental splittings (i.e. S(3P2) = 0 eV; S(3P1) = 0.0491 eV; S(3P0) = 0.0711 eV, respectively 41). For the S-(2P) state, the 2P3/22

P1/2 splitting is calculated 0.0609 eV (491 cm-1) with a similar approach.

Table 3: Spectroscopic parameters of

133

Cs32S (X2Π and 12Σ+) states calculated at the MRCI+Q/aug-

cc-pV5Z level, before and after considering the spin-orbit interactions. State

X2Π

12 Σ +

XΩ3/2

1Ω1/2

2Ω1/2

Τe (cm-1)

0

599

0

307

892

T0 (cm-1)

0

608

0

303

968

202.4

213.2

202.3

200.4

b)

ωe (cm ) -1

ωeχe (cm ) -1

Re (Å)

195.8 a) 201.9 a) 195.7 a) 197.7 a) 281.3 a) 1.95

-12.4

0.639 a) 0.698 a) 3.082

2.906

1.84

5.89

b)

0.595 a)

6.12 a)

6.60 a)

3.081

3.077

2.941

3.092 a) 2.905 a) 3.091 a) 3.088 a) 2.955 a)

a)

Ref. 18.

b)

Strongly anharmonic. See Table 4 for the pattern of the vibrational levels.

12 ACS Paragon Plus Environment

Page 13 of 19 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

The Journal of Physical Chemistry

Figure 6: Enlargement of the PECs displayed in Figure 5 in the 0-1 eV energy range. At low energies, Figure 2 shows that the X2Π and 12Σ+ states cross mutually in the repulsive parts of their potentials and close to the minimum of the 12Σ+ state. The spin-orbit interaction between them results in a significant modification of the Ω = 1/2 potential components, whereas the Ω = 3/2 is just pushed down with respect to the unperturbed X2Π potential. In Table 3, we can see that the splitting of the two components of the X2Π state is calculated to be 307 cm-1 at equilibrium geometry which is coherent with the ionic character [Cs+S-] of the X2Π state and the interaction between the two Ω = 1/2 states which pushes down the 1Ω1/2 state. Figure 6 shows that the 1Ω1/2 PEC presents an inflexion point at RCsS ~ 2.9 Å and that the 2Ω1/2 PEC exhibits a steep repulsive wall for RCsS < 2.8 Å. These spin-orbit-induced changes have an influence on the spectroscopic terms and on the vibrational energy pattern of these states. For instance, the MRCI+Q/aug-cc-pV5Z X2Π data are: Re = 3.082 Å, ωe = 202.4 cm-1 and ωexe = 1.95 cm-1 (Table 3). These values are close to those derived for XΩ3/2: Re = 3.081 Å, ωe = 202.3 cm-1 and ωexe = 1.84 cm-1, whereas, large deviations between X2Π and 12Σ+ spectroscopic parameters and those of 1Ω1/2 and 2Ω1/2 are observed. This is more remarkable for the 12Σ+ / 2Ω1/2 sets of data. For instance, the harmonic frequency of 12Σ+ jumps from ωe ~ 213 cm-1 up to unreasonable value for 2Ω1/2 (of ωe ~ 406 cm-1, not shown in Table 3), which is associated with an unphysical very large value of anharmonic term (ωexe ~57 cm-1, not shown in Table 3). These findings are in line with the analysis of Lee and Wright 18. Since there are no experimental measurements for these states, we list in Table 4 the pattern of vibrational levels for v ≤ 12 deduced from variational calculations. For 1Ω1/2 and 2Ω1/2, the spacing between adjacent vibrational levels is not regular where one can see that ∆G(1Ω1/2) decreases by ~15 cm-1 for v ≤ 3 and then it remains unusually constant at ~160 cm-1. For 2Ω1/2, perturbations in the pattern can be seen for the lowest levels as low as v equals 2-3. 13 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

Page 14 of 19

Table 4: Pattern of the vibrational levels (E, in cm-1) of XΩ3/2, 1Ω1/2 and 2Ω1/2 states of

133

Cs32S. ∆G

(in cm-1) is the energy difference between two successive levels. These energies are given with respect to the energy of the corresponding minimum of each state. The calculated (MRCI+Q/aug-cc-pV5Z) absolute energies at the equilibrium geometry of each state are: E(X2Π) = -417.774729 a.u.; E(12Σ+) = -417.771998 a.u.; E(XΩ3/2) = -417.775498 a.u.; E(1Ω1/2) = -417.774100 a.u.; E(2Ω1/2) = -417.771435 a.u. X2Π

v

12 Σ +

XΩ3/2

1Ω1/2

2Ω1/2

E

∆G

E

∆G

E

∆G

E

∆G

E

∆G

0

101.3

101.3

110.3

110.3

101.3

101.3

98.0

98.0

178.0

178.0

1

300.9

199.6

338.9

228.5

300.9

199.7

284.9

186.9

494.1

316.1

2

499.8

198.9

566.2

227.3

499.9

198.9

455.2

170.2

762.0

267.9

3

700.0

200.2

774.9

208.7

700.2

200.2

605.7

150.5 1026.3 264.3

4

899.1

199.1

984.9

209.9

899.2

199.1

755.5

149.8 1277.9 251.6

5

1096.7 197.7 1192.1 207.3 1096.9 197.7

915.0

159.5 1525.7 247.8

6

1293.8 197.0 1396.1 204.0 1294.0 197.1 1075.0 160.0 1768.4 242.7

7

1490.1 196.3 1599.3 203.1 1490.4 196.4 1237.4 162.4 2006.2 237.8

8

1685.6 195.5 1800.6 201.4 1685.9 195.6 1401.0 163.6 2239.7 233.5

9

1880.5 194.9 2000.4 199.8 1880.9 194.9 1564.9 163.9 2469.0 229.3

10 2074.9 194.4 2199.2 198.7 2075.4 194.5 1729.8 164.9 2694.2 225.1 11 2268.7 193.7 2396.4 197.2 2269.2 193.8 1895.2 165.3 2915.7 221.5 12 2461.4 192.8 2592.6 196.2 2461.9 192.8 2060.8 165.6 3134.0 218.3 For the Ω states correlating to the [Cs(2S1/2) + S(3P2,1,0)] asymptotes, the situation is extremely complicated. Indeed, several avoided crossings exist between the Ω = 1/2 components for 5.5 ≤ RCsS ≤ 6.5 Å. This results in local shallow minima. It is unlikely that these minima support rovibrational levels but they may play a crucial role during the photodissociation of CsS. One can see also that the Ω = 3/2 PECs exhibit an avoided crossing resulting in a flat potential well for the 2Ω3/2 state. The 1Ω5/2 PEC is repulsive. The dynamics of wavepackets promoted into these potentials should be rather complicated.

IV.

Conclusions We used post-Hartree-Fock approaches to compute the PECs of the lowest electronic states of

CsS with and without considering spin-orbit interaction. In addition to the already known X2Π and 12Σ+ states, we give for the first time reliable information on the excited states of this cesium chalcogenide including their spectroscopic parameters, excitation energies and vibrational level patterns. We showed that these quantities are affected by spin-orbit interactions, especially we showed 14 ACS Paragon Plus Environment

Page 15 of 19 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

The Journal of Physical Chemistry

that additional avoided crossings and couplings between Ω states are in action. We also identified odd shapes for the upper electronic states of CsS. Wavepackets promoted into these potentials should show complex dynamics as observed earlier by Zewail and co-workers

37

. Finally, the present data and

findings should help in the experimental identification of the CsS molecule in its ground and electronically excited states. The comparison of highly resolved spectra to our theoretical data should represent a critical test for the ab initio methodologies we used for studying this cesium chalcogenide.

Acknowledgments The authors extend their appreciation to the International Scientific Partnership Program (ISPP) at King Saud University for funding this research work through ISPP# 0045. We gratefully acknowledge the support of the COST Action CM1405 entitled MOLIM: Molecules in Motion.

ASSOCIATED CONTENT Supporting Information The Supporting Information is available free of charge on the ACS Publications website at XX. It gives the composition of the 12 lowest spin-orbit eigenstates in terms of the 82 Kramer doublets arising from the 19 Λ-Σ electronic states and an input and output file of MOLPRO.

AUTHOR INFORMATION Corresponding Authors *E-mail: [email protected] (M.H.). ORCID M. Hochlaf: 0000-0002-4737-7978

15 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

References

(1) Mochalskyy, S.; Modeling of the negative ion extraction from a hydrogen plasma source: application to ITER neutral beam injector. PhD thesis, Université Paris-Sud, 2011. (2) ITER (14. March 2015) http://www.iter.org/sci#fusionfuels. (3) Bretherick, L. Handbook of reactive chemical hazards. 4th ed. Boston, MA: ButterworthHeinemann Ltd., 1990, p. 1082. (4) Herm, R. R.; Herschbach, D. R. Molecular beam kinetics: Reactions of alkali atoms with NO2 and CH3NO2. J. Chem. Phys. 1970, 52, 5783-5792. (5) So, S. P.; Richards, W. G.; The electronic ground states of alkali monoxides. Chem. Phys. Lett. 1975, 32, 227-230. (6) Woodward, J. R.; Hayden, J. S.; Gole, J. L. Formation and characterization of a low-lying electronic state of the alkali monoxides LiO…CsO in the red and near infrared. Chem. Phys. 1989, 134, 395-419. (7) Laskowski, B. C.; Langhoff, S. R.; Siegbahn, P. E. M. Theoretical determination of the X2Σ+ and A2Π potentials of CsO using relativistic effective core potentials. J. Quant. Chem. 1983, XXIII, 483-490. (8) Lee, E. P. F.; Lozeille, J.; Soldan, P.; Daire, S. E.; Dyke, J. M.; Wright, T. G. An ab initio study of RbO, CsO and FrO (X2Σ+; A2Π) and their cations (X3Σ−; A3Π). Phys. Chem. Chem. Phys. 2001, 3, 4863-4869. (9) Allison, J. N.; Goddard, W. A. Alkali oxide diatomics: Explanation of the change in ground state symmetry from LiO(2Π) CsO(2Σ+). J. Chem. Phys. 1982, 77, 4259-4261. (10) Langhoff, S. R.; Bauschlicher, Ch. W.; Patridge, H. Theoretical study of the diatomic alkali and alkaline-earth oxides. J. Chem. Phys. 1986, 84, 4474-4480. (11) Sarkas, H. W.; Hendricks, J. H.; Arnold, S. T.; Slager, V. L.; Bowen, K. H. Measurement of the X2Σ+-A2Π splitting in CsO via photoelectron spectroscopy of CsO-. J. Chem. Phys. 1994, 100, 3358-3360. (12) Yamada, C.; Hirota, E. The microwave spectrum of the cesium monoxide CsO radical. J. Chem. Phys. 1999, 111, 9587-9592. (13) Lindsay, D. M.; Herschbach, D. R.; Kwiran, A. L. ESR spectra of matrix isolated CsO and RbO molecules: 2Σ ground states and inner shell bonding. J. Chem. Phys. 1974, 60, 315-317. (14) Allison, J. N.; Cave, R. J.; Goddard, W. A. Alkali oxides. Analysis of bonding and explanation of the reversal in ordering of the 2Π and 2Σ states. J. Phys. Chem. 1984, 88, 1262-1268. (15) Kaeen, D.; Korek, M.; Abdulal, S. N. Electronic structure of the cesium oxide molecule CsO. J. Mod. Phys. 2015, 6, 1889-1894.

16 ACS Paragon Plus Environment

Page 16 of 19

Page 17 of 19 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

The Journal of Physical Chemistry

(16) Zialenina, M.; Kellö, V.; Černušák, I. Ab initio investigation of cesium monoxide of CsO and CsO+. Proceedings of Student Conference PriF, UK 2015. (17) Partridge, H.; Langhoff, S. R.; Bauschlicher Jr, Ch. W. Theoretical study of the alkali and alkaline-earth monosulfides. J. Chem. Phys. 1988, 88, 6431-6437. (18) Lee, E. P. F.; Wright, T. G. Heavier alkali-metal monosulfides (KS, RbS, CsS, and FrS) and their cations. J. Chem. Phys. 2005, 123, 144309. (19) Dunning Jr. T.H. Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen. J. Chem. Phys. 1989, 90, 1007-1023. (20) Kendall, R. A.; Dunning, T. H.; Harrison, R.J. Electron affinities of the first‐row atoms revisited: Systematic basis sets and wave functions. J. Chem. Phys. 1992, 96, 6796. (21) MOLPRO, version 2015.1, a package of ab initio programs, H.-J. Werner, P.J. Knowles, G. Knizia, F.R. Manby, M. Schütz and others, see http://www.molpro.net. (22) Knowles, P. J.; Werner, H.-J. An efficient second-order MC SCF method for long configuration expansions. Chem. Phys. Lett. 1985, 115, 259-267. (23) Werner, H.-J.; Knowles, P. J. A second order multiconfiguration SCF procedure with optimum convergence. J. Chem. Phys. 1985, 82, 5053-5063. (24) Werner, H.-J.; Knowles, P. J. An efficient internally contracted multiconfiguration–reference configuration interaction method. J. Chem. Phys. 1988, 89, 5803-5814. (25) Knowles, P. J.; Werner, H.-J. An efficient method for the evaluation of coupling coefficients in configuration interaction calculations. Chem. Phys. Lett. 1988, 145, 514-522. (26) Shamasundar, K. R.; Knizia, G.; Werner, H.-J. A new internally contracted multi-reference configuration interaction method. J. Chem. Phys. 2011, 135, 054101. (27) Langhoff, S. R.; Davidson, E. R. Configuration interaction calculations on the nitrogen molecule. Int. J. Quant. Chem. 1974, 8, 61-72. (28) Berning, A.; Schweizer, M.; Werner, H.-J.; Knowles, P. J.; Palmieri, P. Spin-orbit matrix elements for internally contracted multireference configuration interaction wavefunctions. Mol. Phys. 2000, 98, 1823-1833. (29) Numerov, B. Publs. Observatoire Central Astrophys. Russ. 1933, 2, 188. (30) Cooley, J. W. An improved eigenvalue corrector formula for solving the Schrödinger equation for central fields. Math. Comput. 1961, 15, 363-374. (31) Yamada, C.; Fujitake, M.; Hirota, E. The microwave spectrum of the LiO radical. J. Chem. Phys. 1989, 91, 137-141. (32) Yamada, C.; Hirota, M. Infrared diode laser spectroscopy of the LiO radical. J. Chem. Phys. 1993, 99, 8489-8497. (33) Yamada, C.; Fujitake, M.; Hirota, E. Detection of the NaO radical by microwave spectroscopy. J. Chem. Phys. 1989, 90, 3033-3037.

17 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

(34) Hirota, E. Recent progress in high-resolution spectroscopic studies on transient molecules. Bull. Chem. Soc. Jpn. 1995, 68, 1-16. (35) Yamada, C.; Hirota, E. The microwave spectrum of the rubidium monoxide RbO radical. J. Chem. Phys. 1999, 110, 2853-2857. (36) Khadri, F.; Hochlaf, M. Theoretical calculations on the NaS and NaS+ radicals: electronic structure, spectroscopy and spin-orbit couplings. European Phys. J. D 2012, 66, 145. (37) Rose, T. S.; Rosker, M. J.; Zewail, A. H. Femtosecond real‐time observation of wave packet oscillations (resonance) in dissociation reactions. J. Chem. Phys. 1988, 88, 6672-6673. (38) Dantus, M.; Rosker, M. J.; Zewail, A. H. Real‐time femtosecond probing of “transition states” in chemical reactions. J. Chem. Phys. 1987, 87, 2395-2397. (39) Rosker, M. J.; Dantus, M.; Zewail, A. H. Femtosecond clocking of the chemical bond. Science 1988, 241, 1200-2. (40) Bersohn, R.; Zewail, A. H. Time dependent absorption of fragments during dissociation. Ber. Bunsenges. Phys. Chem. 1988, 92, 373-378. (41) N. I. of Standards and Technology, 2017, https://www.nist.gov/pml/basic-atomicspectroscopic-data-handbook.

18 ACS Paragon Plus Environment

Page 18 of 19

Page 19 of 19 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

The Journal of Physical Chemistry

TOC Graphic

19 ACS Paragon Plus Environment