Ab Initio Study of Inelastic Collision Processes for Astronomical Applications

Lei Song

Preliminary version – January 29, 2016 The work described in this thesis was performed as part of the Dutch Astro- chemistry Network, a research program of the Netherlands Foundation for Scientific Research (NWO) and its subdivisions Chemistry (CW) and Exact Sciences (EW).

Printed by Ipskamp Drukkers B.V., Enschede

ISBN:

Preliminary version – January 29, 2016 Ab Initio Study of Inelastic Collision Processes for Astronomical Applications

PROEFSCHRIFT

ter verkrijging van de graad van doctor aan de Radboud Universiteit Nijmegen op gezag van de rector magnificus, volgens besluit van het college van decanen in het openbaar te verdedigen op donderdag 3 maart 2016 om 12.30 uur precies

door

Lei Song

geboren op 5 januari 1986 te Qingdao, China

Preliminary version – January 29, 2016 promotoren: Prof. dr. ir. Gerrit C. Groenenboom Prof. dr. ir. Ad van der Avoird manuscriptcommissie: Prof. dr. David H. Parker Prof. dr. Ewine F. van Dishoeck Universiteit Leiden Prof. dr. Marc C. van Hemert Universiteit Leiden

Preliminary version – January 29, 2016 CONTENTS

1introduction 1 1.1 Astrochemistry background 1 1.2 Theoretical methods 4 1.2.1 Solving the electronic Schrödinger equation 6 1.2.2 Solving the nuclear Schrödinger equation 7 1.3 CO collisions with its partners H, He and H2 11 2h-co potential energy surface 15 2.1 Introduction 16 2.2 Ab initio calculations 17 2.3 Fit procedure 19 2.4 Results and discussion 22 2.4.1 Potential surface 23 2.4.2 Vibrational frequencies and rotational constants of HCO 27 2.5 Conclusions 31 3h-co rate coefficients for pure rotational transi- tions 33 3.1 Introduction 34 3.2 Computational approach 37 3.3 Results and discussion 38 3.4 Astrophysical applications 42 3.5 Conclusion 46 4h-co scattering calculation for ro-vibrational tran- sitions 47 4.1 Introduction 48 4.2 Calculations 50 4.2.1 Theoretical methods 50 4.2.2 Computational details 52 4.3 Results and discussion 54 4.3.1 Convergence test 54 4.3.2 Tests of the CS and IOS approximations 54 4.3.3 Cross sections 58 4.3.4 Vibrational quenching rate coefficients 62 4.4 Conclusions 63 5h-co rate coefficients for ro-vibrational transitions 65 5.1 Introduction 66 5.2 Theory 68 5.2.1 Equations for rate coefficients 68 5.2.2 Extrapolation method 69 5.3 Results 70 5.3.1 Rate coefficients from quantum scattering calculations 70 5.3.2 Extrapolated rate coefficients 73 5.4 Astrophysical models 76

v

Preliminary version – January 29, 2016 5.4.1 ProDiMo slab and disk models 76 5.4.2 Slab model results 77 5.4.3 Disk model results 79 5.5 Conclusions 80 6ar-co scattering for pure rotational transitions 81 6.1 Introduction 82 6.2 Experimental Methods 83 6.3 Theoretical Methods 84 6.4 Results and Discussion 85 6.4.1 Overview 85 6.4.2 Mathematical foundation 89 6.4.3 Sensitivity 92 6.5 Conclusion 94 7he-co scattering for pure rotational transitions 95 7.1 Introduction 96 7.2 Experimental Method 98 7.3 Theoretical method 103 7.4 Results and discussion 104 7.4.1 Theoretical results 104 7.4.2 Vector Correlations 106 7.4.3 Experimental results 108 7.4.4 Discussion 114 7.5 Conclusion 118 8h2 -co scattering calculations 121 8.1 Introduction 122 8.2 Computational details 122 8.3 Theoretical results 123 8.4 Conclusion 126 bibliography 127 summary 145 samenvatting 149 XÅ 153 list of publications 155 curriculum vitae 157 acknowledgements 159

Preliminary version – January 29, 2016 ACRONYMS

2D two-dimensional

3D three-dimensional

AGN active galactic nucleus

BF body-fixed

BO Born-Oppenheimer

CASSCF complete active space self-consistent field

CBS complete basis set

CC close-coupling

CCSD(T) single and double excitation coupled cluster method with pertur- bative triples

CI configuration interaction

CS coupled state

DCS differential cross section

DFT density functional theory

DVR discrete variable representation

FIR far-infrared icMRCI internally contracted multireference configuration interaction

ICS integral cross section

IOS infinite order sudden

IR infrared

ISM interstellar medium

LAMDA Leiden Atomic and Molecular Database

LTE local thermodynamic equilibrium

LWS Long Wavelength Spectrometer

HF Hartree Fock

MP Møller-Plesset

vii

Preliminary version – January 29, 2016 MRCI multi-reference configuration interaction

NIR near-infrared

PDDCS polarization-dependent differential cross section

PDR photon-dominated region

PES potential energy surface

PPD protoplanetary disk

REMPI resonance enhanced multiphoton ionization

RHF restricted Hartree-Fock

RKHS reproducing kernel Hilbert space

RKR Rydberg-Klein-Rees

RMSE root-mean-square error

SAPT symmetry-adapted perturbation theory

SCF self-consistent field

SF space-fixed sinc-DVR sinc-function discrete variable representation

SST Spitzer Space Telescope

UV

VMI velocity map imaging

VUV vacuum ultraviolet

WKS Werner-Keller-Schinke

ZEST zero-energy scaling technique

Preliminary version – January 29, 2016 INTRODUCTION 1

This thesis mainly focuses on ab initio calculations for inelastic collision processes of monoxide, CO, colliding with hydrogen and helium atoms, H and He, and with hydrogen molecules, H2. This introductory chapter presents the astrochemistry background, the theoretical methods applied, and the current status of the research on these scattering processes.

1.1 astrochemistry background

The Universe is a combination of time, space, matter, and energy. [1] Dark matter is estimated to compose 85% of the total matter in the observable Universe, while only a small fraction of the Universe is made of molecules. [2] Still, some scientists refer in particular to the “Molecular Universe” [3, 4, 5] to emphasize the importance of molecules in the formation and evolution of planets, and . By now, nearly 180 different molecular species (not counting isotopologues) have been detected, from simple diatomic molecules such as H2, CO, O2, and CS to complex polyatomic molecules like hydrocarbon chains, and the C60 and C70 fullerenes. Figure 1.1 shows that a diversity of molecules is present in the life-cycle of the interstellar medium. The cycle of molecular material in the interstellar medium starts from the diffuse interstellar medium, to dense molecular clouds where stars and planets form, and finally back to the interstellar medium with the ex- plosion of stars. After that, a new cycle starts. In this process, heavier

Figure 1.1: The life cycle of the interstellar medium. Adapted from Ref. [4]

Preliminary version – January 29, 2016 2introduction

Figure 1.2: An overview of the molecular lines and the range of physical conditions where they can probe effectively. Figure from Ref. [4]

elements are created and become more and more abundant. At the same time, more complex molecular material will be present in the next cycle. To understand the origin and evolution of molecules in space and their role in the Universe, astronomical and chemical knowledge should be combined. Astrochemistry is at the interface of astronomy, chemistry, and molecular physics. [6] This discipline was born in the early 1970s, and its essential sub- ject matter was defined by Dalgarno [7] as “the formation, destruction, and excitation of molecules in astronomical environments and their influence on the structure, dynamics, and evolution of astronomical objects”. The formation, destruction and excitation of molecules are strongly related to the collisional, radiative and chemical processes occurring in astronomical environments. A tremendous amount of microscopic molecular processes results in the macroscopic evolution of stars, planets and galaxies. Not only molecular processes are important, but also the molecules themselves play a signifi- cant role in astrochemistry research. This stems from the fact that molecules are outstanding tracers of the physical conditions in their resident regions. The molecular electronic, vibrational and rotational spectra can reveal the abundance and excitation of molecules, which are determined by the ra- diation from surrounding stars [8] and by collisions. Molecular collisions are sensitive to the physical conditions such as the gas kinetic temperature and the density. Therefore, molecules can be used as “thermometers” and “barometers” of emitting gas by means of spectral surveys. [4] Figure. 1.2 presents an overview of often-used molecular lines and the range where these lines can be used as effective “thermometers” and “barometers”. As shown in Fig. 1.2, CO is an excellent molecular tracer, which is effective for a large range of kinetic temperatures and molecular hydrogen density. Ro- tational lines of CO probe the lower density and lower temperature range,

Preliminary version – January 29, 2016 1.1astrochemistrybackground 3

Figure 1.3: A typical CO astronomical spectrum of the vibration-rotation lines at 4.7µm obtained with VLT-CRIRES. Infrared image with the Spitzer Space Telescope of the central part of Ophiuchus. The SR 21 is one of the young stars. (Image credit: NASA/JPL-Caltech/ESO VLT-CRIRES; K.M. Pontoppidan)

while CO ro-vibrational lines are used in the dense, hot gas near a or in a planet-forming disk. [4, 9]. Molecular hydrogen, as the most abundant molecule in the Universe, is another important tracer for studying the phys- ical properties of molecular clouds. It works in the low density and high temperature range, because of facile low-J rotational transitions in the low- density region and its large energy spacing of rotational and vibrational energy levels. In addition, other tracer species such as OH, CS, H2O, and NH3 are present in the regions indicated in Fig 1.2. The electromagnetic spectra of these molecular tracers provide clues in the investigation of our Universe. Astrochemists utilize astronomical ob- servations, laboratory experiments, theoretical computations, and models. Spectroscopy is a key method of investigation for observational astronomy. Progress in telescopes equipped with highly sensitive detectors advances the observations. The observed spectral range becomes larger and the reso- lution of astronomical spectra becomes higher. Figure 1.3 illustrates a typi- cal astronomical CO spectrum of the central part of Ophiuchus molecular cloud, recorded by the ESO VLT-CRIRES, which is created for infrared ob- servations [10]. An image at infrared wavelengths of this molecular cloud, which is a stellar nursery, is depicted in the background of Fig. 1.3 and has been obtained with the Spitzer Space Telescope (SST). The spectral frequen- cies extracted from the positions of the peaks correspond to differences of the ro-vibrational energy levels, which can be assigned to certain ro- vibrational transitions. The intensities of the peaks reveal the abundance of the molecules in different ro-vibrational levels. Two additional features are relevant in the analysis of spectra from telescopes, when compared with spectra measured in the ordinary laboratory. One is the Doppler shift, the

Preliminary version – January 29, 2016 4introduction

other is the lineshape. Both can provide a great amount of information on the motion of the observed body. The physical parameters and chemical composition in the observed re- gion, such as the kinetic temperature, column density, volume density and molecular abundances can be extracted from this type of spectra through molecular modelling. Often, local thermodynamic equilibrium (LTE) is as- sumed when analysing CO ro-vibrational lines. This only holds when the density of the gas is high, and collisions dominate the excitation of the molecules. However, a non-LTE analysis is necessary for dilute molecular gases, where infrared (IR) and ultraviolet (UV) fluorescent excitation and radiative de-excitation compete with collisions. In this case, an enormous amount of collision rate coefficients of CO with its main partners (H, He and H2) are required as input parameters for modelling. In principle, both laboratory experiments and theoretical calculations can provide collision rate coefficients. The experimentally-inimitable harsh environment in space results in a lack of experimental rate coefficients. In addition, only theory can provide the huge amount of state-to-state rate coefficients as functions of the temperature. Thus, theoretical calculations are the most powerful tools for providing collision rate coefficients. Both classical trajectory meth- ods and quantum mechanical approaches are available for such calcula- tions. The classical trajectory methods are more efficient, while quantum- mechanical calculations are more accurate. In this thesis, we used quantum- mechanical calculations to provide state-to-state cross sections and rate co- efficients for astronomical applications.

1.2 theoretical methods

This section presents the quantum-mechanical method that we used to cal- culate state-to-state rate coefficients for inelastic collision processes for as- tronomic application. In 1926, the Austrian physicist Erwin Schrödinger published a partial differential equation that describes how the quantum state of a physical system changes with time, which has come to be known as the Schrödinger equation:

@ ih = Hˆ .(1.1) @t

The Hamiltonian operator Hˆ characterizes the total energy of the system and takes different forms depending on various physical situations. If we can solve this (non-relativistic) Schrödinger equation and obtain the time- dependent wave function , we can find any property of the system under study. The essence of our ab initio study of inelastic collisions of CO with its partners is to solve the Schrödinger equation for these scattering systems. Because the Hamiltonian itself is not dependent on time in our case, we solve the time-independent Schrödinger equation

Hˆ = E,(1.2)

Preliminary version – January 29, 2016 1.2theoreticalmethods 5 where E is the total energy of the system and is the total time-independent wave function. The Schrödinger equation for atomic and molecular systems includes nuclear and electronic degrees of freedom. A common approach to solve this equation applies the Born-Oppenheimer (BO) approximation, which is based on the fact that the motion of nuclei in a molecule is much slower than that of the electrons. Therefore, the electrons can be considered to be moving in the field of fixed nuclei. Then, solving the Schrödinger equation can be separated into two parts. First, the electronic part of the Schrödinger equation is solved for fixed nuclear coordinates to yield a set of electronic energies and wavefunctions, depending parametrically on the nuclear co- ordinates. The electronic energy, as a function of the nuclear coordinates, defines the so-called potential energy surface (PES). The ab initio calculation of such a PES is described in detail in Subsec. 1.2.1. An example of the PES for H-CO is shown in Fig. 1.4. Secondly, using the constructed PES as input, we can solve the nuclear part of the Schrödinger equation to get state-to- state cross sections for scattering systems. The details are presented in the next Subsec. 1.2.2. Then, we illustrate how to get rate coefficients from calcu- lated cross sections. Because full quantum mechanical calculations are pro- hibitively expensive, astronomers use extrapolation to obtain the required rate coefficients from a limited set of known data. An extrapolation method that is currently in use has inherent defects, so in the last part of this sub- section we introduce a new extrapolation method to enlarge the number of rate coefficients.

Figure 1.4: The H-CO potential energy surface with CO fixed at the equilibrium distance r = 2.1322a0.

Preliminary version – January 29, 2016 6introduction

1.2.1 Solving the electronic Schrödinger equation

In this subsection, we briefly introduce the ab initio electronic structure methods for single energy calculations that we used in the construction of PES. The Hartree Fock (HF) method is the simplest approach to solve the multi- electron Schrödinger equation in a stationary state for the wave function and the energy. The essence of HF theory is that the multi-electron system is treated as a one-electron problem by averaging the electron-electron inter- action, so that each electron moves independently in the field of the other electrons. This is known as the mean field approximation. Then, we can define a one-electron effective Hamiltonian operator known as the Fock op- erator Fˆ (the same for all electrons), and solve the one-electron Schrödinger equation Fˆi = ✏ii,(1.3)

where i is a set of one-electron wave functions, called the HF molecular or- bitals, and ✏i are the energies of these molecular orbitals. When considering the spin, the i should be replaced by spinorbitals i. The HF approxima- tion assumes that the wave function of the multi-electron system is given by a single Slater determinant containing all occupied HF molecular spinor- bitals i. Such a determinant wave function is antisymmetric with respect to the interchange of the coordinates of two electrons and satisfies the Pauli exclusion principle. Application of the variational method, with the restric- tion that the (spin)orbitals remain orthogonal, yields the HF equation (1.3). Since the Coulomb and exchange terms in the Fock operator Fˆ depend on all occupied orbitals, the HF equation has to be solved iteratively, in a so-called self-consistent field (SCF) procedure. Because of the antisymmetry of the de- terminant wave function the HF method takes into account some correlation between electrons with parallel spin, but completely ignores the correlation between electrons with opposite spin. The HF method is the starting point for most methods that treat electron correlation problem. In the following, we briefly introduce the coupled cluster method, a post- HF method, which is one of the most prevalent ab initio quantum chemistry methods, and used in the construction of our PES. The basis ansatz of cou- pled cluster theory is CC = exp(Tˆ ) HF,(1.4)

where CC and HF are the coupled cluster wave function and the HF (deter- minant) wave function, respectively. The exponential operator exp(Tˆ ) can be expanded as

Tˆ 2 Tˆ 3 Tˆ K exp(Tˆ )=1 + Tˆ + + + = .(1.5) 2! 3! ··· 1 K! XK=0 The cluster operator Tˆ can be written as

Tˆ = Tˆ + Tˆ + Tˆ + + Tˆ ,(1.6) 1 2 3 ··· N

Preliminary version – January 29, 2016 1.2theoreticalmethods 7

where Tˆ1 is the operator of all single excitations, Tˆ2 stands for the operator of all double excitations, etc. The total energy and the coupled cluster wave function up to a certain excitation level n can be obtained by substitution of this wave function into the Schrödinger equation, and projection of the equation onto the manifold of all n-tuply excited determinants. This yields a set of coupled non-linear equations that can be solved self-consistently. In principle, coupled cluster theory can provide the exact solution to the time-independent Schrödinger equation. In practice, the operator Tˆ should be truncated at a certain level of excitation, balancing computational expen- diture and accuracy. It is most common to truncate at the double excitation level Tˆ1 + Tˆ2 and to take into account the effect of triple excitations Tˆ3 by perturbation theory, which gives the so-called single and double excitation coupled cluster method with perturbative triples (CCSD(T)) method. This is the method that we used in our work. The coupled cluster method belongs to the single-reference methods, be- cause the exponential ansatz includes only one reference function HF. The coupled cluster method is not variational, but in contrast with variational methods such as configuration interaction (CI) it is size-consistent. Size- consistency implies that the energy of a system AB becomes equal to the sum of the energies of the subsystems A and B when they are moved in- finitely far away from each other. This makes the CCSD(T) method suitable for the calculation of interaction energies between atoms and/or molecules. Beside the coupled cluster method, there are many other post-HF meth- ods, such as CI, Møller-Plesset (MP) perturbation theory, quantum chemistry composite methods and so on. Also density functional theory (DFT) is an often applied option for ab initio electronic structure calculations. All these methods have their advantages and shortcomings and one should choose one or more methods suitable for specific applications. To calculate the in- teraction of CO with a hydrogen atom H CCSD(T) is a good option, because it is known to yield accurate interaction energies and it is size-consistent. When the H-CO interaction energies are obtained from electronic structure calculations for many (fixed) nuclear geometries of the H-CO complex, we can fit them to an analytic function of the nuclear coordinates to construct the PES. This process is illustrated in Chap. 2.

1.2.2 Solving the nuclear Schrödinger equation

Here we briefly describe how to solve the nuclear part of the Schrödinger equation for inelastic collisions. First, to understand this physical process, we classically illustrate the key vectors of the collision between a diatomic molecule and a structureless atom in Fig. 1.5. The vector j is the rotational momentum of the diatomic molecule and l, associated with relative motion of the partners, is called the orbital angular momentum. With and without prime we distinguish the final and initial states. The conservation of total angular momentum J is required in the whole process, so that J = j + l = j0 + l0. Because we consider inelastic collisions, there is interchange between

Preliminary version – January 29, 2016 8introduction

the kinetic energy of the relative motion, i.e., the collision energy, and the

Figure 1.5: Key vectors in inelastic collision of a diatomic molecule with an atom. Left: before collision. Right: after collision. Figure from Ref. [11].

vibrational and/or rotational energy of the molecule. Our aim is to solve the nuclear Schrödinger equation quantum-mechanically to yield the state- to-state cross sections and finally to provide rate coefficients contributing to the astronomic database. The nuclear Hamiltonian in a body-fixed (BF) frame for the atom-diatom complex can be written (in atomic units) as

1 @2 Jˆ2 + ˆj2 - 2jˆ Jˆ Hˆ =- R + · + V (r, R, ✓)+Hˆ ,(1.7) 2µR @R2 2µR2 int diatom

where µ is the reduced mass of the complex, r is the diatom bond length, and R is the length of the vector R that connects the atom with the center of mass of the diatom. The angle ✓ is the angle between the vector R and the diatom bond axis, Vint(r, R, ✓) is the PES of the atom-diatom interaction obtained by solving the electronic Schrödinger equation. The operator

2 ˆ2 ˆ 1 @ j Hdiatom =- 2 r + 2 + Vdiatom(r) (1.8) 2µdiatomr @r 2µdiatomr

is the Hamiltonian of the diatom with the reduced mass µdiatom and Vdiatom(r) is the PES of the free diatom.

Preliminary version – January 29, 2016 1.2theoreticalmethods 9

To solve the time-independent Schrödinger equation, Eq. 1.2, with the Hamiltonian of Eq. 1.7, the wave functions are expanded in the BF angular basis |n as i 1 = |n (R),(1.9) R i n n X where n(R) are functions of the radial coordinate. Projecting the resulting equation with n | gives the coupled channel equations h 0 @2 (R)= n0 Wˆ |n (R),(1.10) @R2 n0 i n n X ⌦ where the operator Wˆ is defined as Wˆ = 2µ(Hˆ - Tˆrad - E) and Tˆrad is the radial kinetic energy term, the first term in Eq. 1.7. The detailed derivation of the coupled channel equations is described in Sec. 4.2 of Chap. 4. Here, we will present the numerical algorithm to solve these coupled second-order ordinary differential equations in practice. When we collect the functions n(R) in the column vector (R), we can write Eq. 1.10 in matrix form as

00(R)=W(R) (R).(1.11)

We use the renormalized Numerov algorithm [12, 13]. First, the radial co- ordinate R is discretized into an equidistant grid of points Ri(i = 0, ..., m) with step size = Ri - Ri-1. Secondly, we expand the wave functions i (an abbreviation of (Ri)) and i-1 in a Taylor series 1 1 1 1 = - + 2 - (3)3 + (4)4 - (5)5 + O(6) i-1 i i0 2 i00 6 i 24 i 120 i (1.12) 1 1 3 1 4 1 5 = + + 2 + ( )3 + ( )4 + ( )5 + O(6). i+1 i i0 2 i00 6 i 24 i 120 i (1.13)

In the third step, we add Eqs. 1.12 and 1.13

(4) + = 2 + 2 + i 4 + O(6) (1.14) i-1 i+1 i i00 12 and rearrange the terms

(4) i-1 - 2 i + i+1 i 2 4 00 = - + O( ),(1.15) i 2 12

(4) where i can be written as

(4) Wi-1 i-1 - 2Wi i + Wi+1 i+1 2 =( 00)00 =(W )00 = + O( ).(1.16) i i i i 2

Preliminary version – January 29, 2016 10 introduction

Then, we substitute Eqs. 1.15 and 1.16 into Eq. 1.14 and after some rewriting obtain the following recursion formula

2 5 2 (I - W ) = 2(I + 2W ) -(I - W ) + O(6),(1.17) 12 i+1 i+1 12 i i 12 i-1 i-1

2 where I is the identity matrix. In the next step, we define Fi = I - 12 Wi and 6 substitute Fi into Eq. 1.17, while neglecting the O( ) terms. This yields a recursion formula for i. Finally, we define the Q matrix as i = Qi+1 i+1, substitute it into Eq. 1.17 and rearrange the equation to get the recursion relation, -1 Qi+1 =(12I - 10Fi - Fi-1Qi) Fi+1.(1.18) The above equation is the basic relation of the renormalized Numerov method, by which the matrix Qi can be propagated. Equation 1.18 can be solved to obtain Qm once the initial value Q1 is specified. In scattering prob- lems, the initial 0 is commonly set to zero while 1 is not zero, so that the initial Q1 is zero. After the propagation of Q from R1 to Rm one assumes that Rm-1 and Rm are sufficiently large to make the interaction energy, i.e., the third term in Eq. 1.7, negligibly small. Then, the exact solutions of the nuclear Schrödinger equation in the space-fixed (SF) frame are linear combinations of the proper spheric Bessel functions [14]. K matrix boundary condition are applied by requiring that

(SF) m-1 = Bm-1 - Gm-1K (1.19) (SF) m = Bm - GmK,(1.20)

where the matrices Bi and Gi are related to spherical Bessel functions jl(knRi) and yl(knRi) of the first and second kind as follows:

Bi = Bn,l,n0,l0 (Ri)=nn0 ll0 Rijl(knRi) (1.21)

Gi = Gn,l,n0,l0 (Ri)=nn0 ll0 Riyl(knRi).(1.22)

The spherical Bessel functions are labeled by the SF end-over-end rotational quantum number l, which has become a good quantum number at such 2 2 large separation. The wave number kn is calculated by kn = 2µ(E - ✏n)/h , where the channel energies ✏n are the eigenvalues of Hˆ diatom in Eq. 1.8 and E is the total energy. When dividing Eq. 1.19 by Eq. 1.20, we obtain the relation between the K matrix and the SF Q matrix

(SF) (SF) (SF) -1 -1 Qm = m-1( m ) =(Bm-1 - Gm-1K)(Bm - GmK) .(1.23)

Preliminary version – January 29, 2016 1.3cocollisionswithitspartnersh, he and h2 11

Because the spherical Bessel functions are well known, the K matrix can be (SF) calculated if the matrix Qm is available. We can transform the propagated (BF) (SF) BF Q to a SF Q with the unitary matrix U = BF|SF , m m h i (SF) (BF) Qm = U†Qm U.(1.24)

The BF basis is labeled by (j, K), where K is the projection of the total angular momentum J and, simultaneously, of the rotational quantum number j of the diatomic molecule on the BF z-axis. The SF basis is labeled by (j, l) and the elements of the unitary transformation matrix are given by

jl J (-1)j-l+Kp2l + 1 if K = 0 8 K0-K ! UjK,jl = BF|SF jK,jl = > h i > jl J <> ((-1)j-l+K + p(-1)K) l + 1/2 if K>0. K0-K ! > p > Here, p is the parity.:> From Eqs. 1.23 and 1.24 we can then obtain the K matrix. The S matrix, defined as

S =(I - iK)-1(I + iK),(1.25) includes all the information to deduce the state-to-state collision cross sec- tions. When we consider ro-vibrationally inelastic atom-diatom collisions, with v being the vibrational quantum number of the diatom, the cross sec- tions collisions are given by

J+j J+j ⇡ 0 E 2J 1 S(J) v,j v0,j0 ( v,j)= ( + ) vv0 jj0 ll0 - vjl v j l , ! 2 , 0 0 0 (2j + 1)kv,j J l=|J-j| l0=|J-j0| X X X (1.26) where Ev,j = E - ✏v,j is the collision energy and kv,j is the wave number of the incoming channel with collision energy Ev,j. The rate coefficient is the thermal average of the product of the cross section and the relative velocity of the colliding particles. State-to-state rate coefficients can be calculated from the following formula

1 8kBT 2 1 Ev,j rv,j v ,j (T)= v,j v ,j (Ev,j)exp - Ev,jdEv,j, ! 0 0 ⇡µ 2 ! 0 0 k T ✓ ◆ (kBT) Z01 ✓ B ◆ (1.27) where kB is the Boltzmann constant and T the gas temperature. These cal- culated rate coefficients can be used as inputs for astrophysical modelling.

1.3 co collisions with its partners h, he and h2

Carbon monoxide, as the second most abundant molecule in the Universe, is often used as a tracer for the many regions where it resides as illustrated

Preliminary version – January 29, 2016 12 introduction

Figure 1.6: Pure rotational inelastic cross sections for the transition v = 0, j = 1 ! v0 = 0, j0 = 0 based on four different potential energy surfaces. Figure from Ref. [15]

in Sec. 1.1. Its main collision partners H, He, and H2 are widespread and abundant species in the Universe as well. The complexes formed by CO and its main partners do not have the same chemical properties. The He-CO and H2-CO complexes are bound only by weak van der Waals forces with binding energies of about 23 cm-1 and 94 cm-1, respectively. By contrast, the H-CO complex is a chemically bound species with the much larger binding energy of 6738 cm-1 (0.835 eV). In addition, the H-CO complex has a barrier to dissociation into H + CO with a height of 1138 cm-1 (0.141 eV), whereas there is no such barrier in He- CO and H2-CO. These properties make the H-CO complex very different, and make it much more difficult to calculate collision cross sections than in the other cases. No accurate and complete collision cross sections and rate coefficients have been available for the H-CO system. Recent astrophysical modelling [9] utilized vibrational and rotational rate coefficients calculated by Balakrishnan et al. [16], which are based on the three-dimensional (3D) Werner-Keller-Schinke (WKS) H-CO potential [17, 18]. However, Shepler et al. demonstrated that the inaccurate long range part of the WKS potential seriously affects the precision of pure rotational collision cross section and rate coefficients. [19] As illustrated in Fig. 1.6, Yang et al. [15] recently pre- sented rotational cross sections based on the WKS potential that differ sig- nificantly from values based on two-dimensional (2D) multi-reference con- figuration interaction (MRCI) and CCSD potentials [19]. This discrepancy confirmed the inaccuracy of the long range part in the WKS potential. Thus, a new 3D H-CO potential with an accurate long range part was needed and new cross sections and rate coefficients were required for astrophysical modelling.

Preliminary version – January 29, 2016 1.3cocollisionswithitspartnersh, he and h2 13

For the He-CO system, pure rotational quenching rate coefficients with initial rotation quantum numbers j up to 14 at temperatures of 5 K 6 T 6 500 K and vibrational de-excitation rate coefficients with initial vi- brational quantum numbers v up to 6 for a temperature range of T = 500 - 1300 K were reported by Cecchi-Pestellini et al. [20]. Moreover, var- ious ro-vibrational cross sections and rate coefficients were published for different transitions and temperature ranges [21, 22, 23, 24, 25, 26]. For the H2-CO system, Yang et al. [27] reported pure rotational quenching rate co- efficients from initial states with j = 1 - 40 for a gas temperature range of T = 1 - 3000 K. Reid et al. [28] presented cross sections for the ro-vibrational deactivation of CO (v = 1) by inelastic collisions with H2 for temperatures 30 K 6 T 6 300 K. Ideally, we would like to compare calculated cross sections for H+CO with experimentally determined values. However, a crossed molecular beam experiment on H-CO is extremely hard, because it is difficult to produce a beam of pure H atoms. Experiments of this type are planned by the Molec- ular and Laser Physics group in the Radboud University, which already provided data on the easier CO-He, CO-Ar, and CO-H2 systems. They mea- sured differential cross sections (DCSs) and polarization-dependent differ- ential cross sections (PDDCSs) using the velocity map imaging (VMI) tech- nique combined with (1+10) and/or (2+1) resonance enhanced multiphoton ionization (REMPI) detection of CO in its different final states. We also calcu- lated DCSs and/or PDDCSs for these systems from available ab initio PESs to compare with their experimental data. The differential cross section (DCS) reveals the pair correlation between the initial and final velocity vectors, k and k0, while the polarization-dependent differential cross section (PDDCS) reveals the triple correlation between the initial and final velocities and the orientation of the final rotational angular momentum j0.[29] This informa- tion is valuable to understand the collision dynamics in detail, and to check the accuracy of the PES used to describe the intermolecular and intramolec- ular interactions.

Preliminary version – January 29, 2016 Preliminary version – January 29, 2016 H-CO POTENTIAL ENERGY SURFACE 2

2 In this chapter, we present an ab initio potential for the H-CO(X A0) com- plex in which the CO bond length is varied and the long range interac- tions between H and CO are accurately represented. It was computede using the spin-unrestricted open-shell single and double excitation coupled clus- ter method with perturbative triples [RHF-UCCSD(T)]. Three doubly aug- mented correlation-consistent basis sets were utilized to extrapolate the cor- relation energy to the complete basis set limit. More than 4400 data points were calculated and used for an analytic fit of the potential: long range terms with inverse power dependence on the H-CO distance R were fit to the data points for large R, the reproducing kernel Hilbert space method was applied to the data at smaller distances. Our potential was compared with previous calculations and with some data extracted from spectroscopy. Furthermore, it was used in three-dimensional discrete variable representa- tion calculations of the vibrational frequencies and rotational constants of HCO, which agree very well with the most recently measured data. Also the dissociation energy D0 = 0.623 eV of HCO into H + CO obtained from these calculations agrees well with experimental values. Finally, we made preliminary two-dimensional calculations of the cross sections for rotation- ally inelastic H-CO collisions with the CO bond length fixed and obtained good agreement with recently published two-dimensional results.

Lei Song, Ad van der Avoird, and Gerrit C. Groenenboom, Three-dimensional 2 ab initio potential energy surface for H-CO(X A0), J. Phys. Chem. A, 117, 7571 (2013) e

Preliminary version – January 29, 2016 16 h-co potential energy surface

2.1 introduction

As the second most abundant molecule in the universe, carbon monoxide is typically observed in star-forming regions and protoplanetary disks, and in external galaxies. It is commonly used as a tracer of physical conditions in the regions where CO infrared emission is observed. To extract astro- physical parameters from the observed spectral lines, it is often not appro- priate to assume LTE in the dilute gas of astronomical sources. Non-LTE methods used to analyze the available observations require collisional rate coefficients. [30, 31] The interpretation of the observed CO spectral lines critically relies on the collisional rate coefficients of CO with the dominant collision partners: H, H2, and He. [32, 33] These rate coefficients can be obtained from scattering calculations based on potentials of CO with its collision partners, thus accurate PESs are urgently needed. Compared with CO-H2 or CO-He, the interaction between CO and H is relatively strong and the equilibrium CO bond length in HCO is sub- stantially larger than in free CO. Hence, a 3D potential with CO bond length dependence is needed. Pioneering work on the dynamics of HCO based on the 3D BBH potential constructed by Bowman, Bitman, and Hard- ing was carried out by Bowman and coworkers [34, 35, 36]. Werner et al. discovered later that the BBH potential and its empirically modified ver- sions [35, 36] did not reproduce well the spectroscopic data of Tobiason et al., [37] so they developed the 3D WKS potential [17, 18]. This potential was based on elaborate high-quality complete active space self-consistent field (CASSCF) and MRCI calculations, with special attention for the conical inter- sections between the potentials of different electronic states. Much theoreti- cal [17, 18, 34, 35, 36, 38, 39, 40] and experimental [37, 41, 42, 43, 44, 45, 46] work has been devoted to the study of resonances that play an important role in the of HCO into H + CO. Classical and (approx- imate) quantum mechanical scattering calculations [47, 48, 49] studying the rotational and vibrational (de)excitation of CO in collisions with hot H atoms have been performed with both the BBH and WKS potentials. It was found that the well region of the potential is important especially for vibrationally inelastic collisions, and that the rotationally inelastic scatter- ing cross sections deviated from experimental data probably because the van der Waals region of the potential was insufficiently accurate. In this chapter we focus on the lower energy region of the HCO potential surface. The cross sections for rotational and vibrational (de)excita- tion of CO in collisions with H atoms are important for astrophysical modeling. Shepler et al. [19] have shown recently that the rotationally inelastic H-CO collision cross sections critically depend on the long-range behavior of the PES, especially for collisions at low energies. Calculations on the WKS poten- tial produced much larger values of these cross sections than on the BBH potential and also disagree with the cross sections computed by Shepler et al. on their CCSD(T) and MRCI potentials. The analysis in Ref. [19] shows that the reason for this discrepancy is that the WKS potential does not reflect

Preliminary version – January 29, 2016 2.2 ab initio calculations 17 well the anisotropic long-range van der Waals interaction. Shepler et al. paid much attention to make their potentials describe the correct long-range be- havior, but they only produced 2D potentials [19] with the CO bond length fixed. The goal of this work is to provide an accurate 3D ab initio potential en- 2 ergy surface of the H-CO complex in the ground state (X A0) with explicit consideration of the long-range behavior. Details of the ab initio electronic structure calculations are given in the next section. Sectione 2.3 describes the fit procedure and shows the accuracy of the fit. In Sec. 2.4.1, we present our potential, compare it with previous studies, and discuss some properties. Furthermore, we compare in this section the rotationally inelastic H-CO collision cross sections computed on our potential with results obtained on the 2D potentials of Shepler et al. In Sec. 2.4.2 we use our potential surface to compute the vibrational frequencies and rotational constants of HCO by means of a 3D discrete variable representation (DVR) method, and we show that they are in good agreement with the best available experimental data. The conclusion and outlook are given in Sec. 2.5.

2.2 ab initio calculations

In order to explore the structure of the global potential energy surface, we first performed state-averaged CASSCF [50, 51] calculations. The goal of these calculations was to identify conical intersections of the ground state potential of A0 symmetry with the potentials that correlate with the lowest excited ⇧ state of CO. Starting with orbitals from restricted Hartree- Fock (RHF) calculations in an augmented correlation-consistent polarized valence quadruple-⇣ (aug-cc-pVQZ) basis [52] we performed three-state av- eraged full valence CASSCF calculations with 11 electrons in 9 active or- bitals. They revealed the existence of a conical intersection at nonlinearity, in addition to those that were already found at the linear H-CO and CO-H structures [17, 18]. However, all of these intersections occur at considerably higher energies than the lowest conical intersection identified by Werner et al., which occurs at approximately 1.4 eV 11000 cm-1 above the H + ⇡ CO asymptotic limit [17, 18]. Since we are interested in H-CO collisions for rotational and v = 0 1 vibrational (de)excitation of CO at lower colli- ! sion energies we may assume that we only need the ground state potential of A0 symmetry in the energy region well below all conical intersections. Full CASSCF and MRCI calculations as performed by Werner et al. [17, 18] are therefore not needed. On the other hand, we wish our potential to be especially accurate in the long-range, so we had to use a method that is size-extensive (see Ref. [53], Sec. 13.3). We adopted the spin-unrestricted open-shell single and double excitation coupled cluster method [54, 55] with perturbative triples [56] [UCCSD(T)], with the molecular orbitals from RHF calculations. Some further discussion on the effects of the conical in- tersections is given below, after the presentation of our calculated potential and its analytic fit.

Preliminary version – January 29, 2016 18 h-co potential energy surface

Figure 2.1: Jacobi coordinates for H-CO; Q denotes the CO center of mass.

The ab initio electronic structure calculations were performed with the MOLPRO 2000 package.[57] Doubly augmented correlation-consistent d- aug-cc-pVnZ basis sets [52] were used with n = 3, 4, and 5, and the corre- lation energy was extrapolated to the complete basis set (CBS) limit [58, 59]. The H-CO interaction energy was calculated with the Boys and Bernardi counterpoise correction to avoid the basis-set superposition error [60] (BSSE), i.e., the interaction energies were determined as EHCO - ECO - EH, with the H and CO monomers computed in the same one-electron basis as the H- CO dimer. In total, we calculated 4428 interaction energies: 3744 grid points used in the fit and 684 random points to test the fit. The grid is based on three Jacobi coordinates, shown in Fig. 2.1: the CO bond length r, the distance R, i.e., the length of the vector R pointing from the center of mass of CO to the H nucleus, and the angle ✓ between R and the CO axis r. For ✓ we employed 13 Gauss-Legendre quadrature points in the range from 0 to 180. We chose 32 points for the R coordinate from 2.0a0 to 20.0a0. In the range of 2.0a0 6 R<6.0a0, the step size R is 0.2a0, while R equals 0.4a0 in the range of 6.0a0 6 R<8.0a0.A logarithmically spaced grid with seven points is applied for the range of 8.0a0 6 R 6 20.0a0. The vibrational coordinate r is varied from 1.65 a0 to 2.85 a0, including 1.65, 1.85, 2.05, 2.1322, 2.20, 2.25, 2.45, 2.65, 2.85 a0. We note that the RHF-UCCSD(T) method is a single-reference electron correlation method, which will fail near the conical intersections and in the CO bond dissociation region. The so-called T1 diagnostic [61, 62, 63] is a measure for the amount of multireference behavior; the larger the value of this T1 diagnostic, the less reliable the results from the single-reference electron correlation method are. In this work, we removed or reduced the weight of five ab initio points, because of their bad T1 diagnostic (ranging from 0.085 to 0.142).

Preliminary version – January 29, 2016 2.3 fit procedure 19

2.3 fit procedure

The 3D H-CO potential is written as the sum of the CO diatomic potential and the interaction energy between H and CO,

VHCO(r, R, ✓)=VCO(r)+Vint(r, R, ✓).(2.1)

The CO diatomic molecule potential VCO(r) was obtained from the vibra- tional energies Gv and rotational constants Bv measured by Le Floch [64]by means of the first-order semiclassical Rydberg-Klein-Rees (RKR) procedure, as implemented by Le Roy [65]. The analytical H-CO interaction potential Vint(r, R, ✓) was obtained from a fit to our ab initio RHF-UCCSD(T) H-CO interaction energies. We first fitted its ✓-dependence, then its R-dependence, and finally its r-dependence, with the following procedure. The ✓-dependence of the interaction energy for each of the (r, R) points was expanded as

12 Vint(r, R, ✓)= Vl(r, R)Pl(cos ✓),(2.2) Xl=0 where the Pl(cos ✓) are Legendre polynomials with l = 0 - 12. The ex- pansion coefficients Vl(r, R) are calculated by means of a 13-point Gauss- Legendre quadrature when the vibrational coordinate r is 6 2.20 a0. For larger r, five points with T1 diagnostic larger than 0.085 were removed or weighted by a factor less than 0.01 in the fit process. In this case, the Gauss- Legendre quadrature grid is incomplete, so another method is needed. We employed the linear least-squares fitting procedure with Tikhonov regular- 2 2 ization [66] in which the term ↵ l |l Vl(r, R)| is added to the residual. The factor ↵ was set to 10-10 to ensure that strong oscillations (associated with P large l values) are damped out in the fit. Then, we represented the coefficients Vl(r, R) by analytical functions of R, for each value of r. For the terms with l 6 5, the long-range and short-range contributions were fit separately, and summed

(lr) (sr) Vl(r, R)=Vl (r, R)+Vl (r, R).(2.3)

The dominant long-range interactions between H and CO are dispersion forces. Smaller contributions are the interactions between the (small) dipole and higher multipoles of CO and the multipole moments induced on the H atom. From perturbation theory with the interaction operator in the multi- pole expansion it follows that all of these long-range interactions depend on the H-CO distance as R-n, with the leading term having n = 6. The theory shows, moreover, [67] that this leading power n depends on the anisotropy

Preliminary version – January 29, 2016 20 h-co potential energy surface

(lr) n -n Table 2.1: Coefficients Vnl (r) [in Eh a0 ] of the long-range R terms in Eq. (2.4) obtained from our fits, see Sec. 2.3, with l denoting the indices of the Legendre polynomials Pl(cos ✓) that represent the angular dependence of the potential.

rV6,0 V6,2 V7,1 V7,3 V8,4 V9,5

1.65 18.62154 0.62830 -81.20027 -1.27329 4.35816 -61.11080

1.85 19.58438 0.75523 -75.33132 0.69718 0.31787 2.50351

2.05 20.90374 1.06060 -70.39766 2.72797 3.89982 290.03122

2.1322 21.22015 1.42621 -70.30271 4.75132 -22.07285 230.87453

2.20 21.83178 1.69050 -60.18380 3.94278 -45.09345 -1267.90953

2.25 22.09369 1.63455 -65.00574 4.60370 -20.28939 188.78523

2.45 23.80770 2.58206 -63.52498 4.47302 3.20559 247.74622

2.65 25.32512 3.50240 -64.28439 2.06188 -0.55607 -44.36051

2.85 25.56625 4.61577 -68.00236 0.90311 -9.14573 -85.05427

of the interaction, as defined by the index l in the angular expansion of Eq. 2.2. We took this relation into account in our fit of the analytic form

(lr) (lr) -n Vl (r, R)= -Vnl (r)fn(R) R ,(2.4) n X to the long-range data for R>10a0. Only terms with even n + l were in- -n cluded. For l = 0, 2 the inverse powers n of R start with ni = 6, while for l = 1, 3 the first term has ni = 7. When l > 4, n begins with ni = l + 4. -ni (lr) Only the leading term with R was included in Vl (R, r); the coefficient V(lr)(r) of this term was obtained from a least-squares fit of the first two nil terms with R-ni and R-ni-2 in Eq. 2.4 to the energies at the five long-range R values with R>10a0. The coefficient of the second term changed consid- erably when this fit was restricted to a subset of these long-range R values, while that of the leading term remained practically the same. Therefore, we (lr) included only the latter. The long-range coefficients Vnl (r) obtained from

Preliminary version – January 29, 2016 2.3 fit procedure 21

our fits for all values of r are listed in Table 2.1. The functions fn(R) are Tang-Toennies damping functions [68] defined as

n k -x x fn(x)=1 - e (2.5) k! Xk=0 -1 with x = R and the parameter set to 2.0a0 . They suppress the singular behavior of the long-range terms in the short range. No explicit long-range contributions were calculated for l>5. After subtraction of the long-range contributions with l 6 5 from Vl(r, R) (sr) the short-range contributions Vl (r, R) were obtained for each value of r by interpolation over the full range of R with the reproducing kernel Hilbert space (RKHS) method [69, 70]. The smoothness parameter n of the RKHS method, (see Eq. 9 in Ref. [69]), was fixed to 2, while the RKHS parameter m (sr) that determines the asymptotic behavior of Vl (r, R) was chosen to depend (sr) on the value of l. For l 6 5, m was set to ni + 1, so Vl (r, R) decays asymp- totically as R-ni-2. For l>5, where no separate long-range contributions -n were calculated, m was set to ni - 1, so that Vl(r, R) decays as R i . Finally, the values of the potential Vint(rp, Ri, ✓j) on the grid points rp, calculated at arbitrary values of Ri, ✓j with the functions obtained from the previous fits in ✓ and R, were fitted to a polynomial expansion

5 Vint(r, Ri, ✓j)= Ck(Ri, ✓j)Pk(Ar + B).(2.6) Xk=0

The orthogonal polynomials Pk(Ar + B) are Legendre polynomials in the coordinate Ar + B with scaling parameters A and B chosen such that the range 1.65 a0 6 r 6 2.85 a0 is linearly mapped onto the required interval from -1 to +1. The expansion coefficients Ck(Ri, ✓j) were obtained from a weighted linear least-squares fit. Points with higher energies (Vint >0.1Eh) are less important in bound state and lower energy scattering calculations, so we chose an exponentially decreasing weight

w = e-(Vint-0.1) (2.7) for those points, with = 4 and all quantities in the exponent given in Eh. The fit accuracy and the correct short-range repulsive behavior are con- served by this procedure. The accuracy of the fit was tested in two ways. First, by comparison of potential energies from the fit with the ab initio values used, for all data points with Vint(rp, Ri, ✓j) 6 0.1Eh, but with the exclusion of the five points that had bad T1 diagnostics. The root-mean-square error (RMSE) defined by

1 fit ab initio 2 RMSE = V (rp, Ri, ✓j)-V (rp, Ri, ✓j) (2.8) vn int int u p,i,j u X t

Preliminary version – January 29, 2016 22 h-co potential energy surface

300

250 Ψ (r) v=15

200 ) h

150 (r) (mE CO V

100

50

0 1.6 1.8 2.0 2.2 2.4 2.6 2.8 3.0 r (a ) 0

Figure 2.2: Empirical RKR potential VCO(r) (solid line) and calculated RHF- UCCSD(T) potential (dashed line), with the v = 15 vibrational wave function calculated on the RKR potential.

is 1.31 10-4 E for 2944 points in the short and middle range (2.0a ⇥ h 0 6 R 6 10.0a0). The RMSE relative to the mean absolute values of Vint for each value of R in this range is 0.29%. For the long range (R>10.0a0), the absolute RMSE is 2.66 10-9 E for 585 points and the relative RMSE ⇥ h is 0.20%. Secondly, we calculated 684 additional random points to test the fit. With the exclusion of the points with energy Vint >0.1Eh or bad T1 diagnostics (>0.085), 662 random points were used in this analysis. The RMSE is 3.25 10-4 E over 294 points in the short and middle range, and ⇥ h 2.61 10-8 E over 368 points in the long-range; the overall relative RMSE ⇥ h is 0.56%.

2.4 results and discussion

In order to check first for how large an r value we can still trust the RHF- UCCSD(T)/CBS method used to compute our 3D H-CO potential and how accurate the RHF-UCCSD(T) potential is, we computed the vibrational lev- els and wave functions of the free CO molecule on the empirical RKR poten- tial and on a potential computed for free CO by the same RHF-UCCSD(T)/ CBS method as used in the calculations on H-CO. Figure 2.2 shows the two potentials; for all r up to 3.05 a0 they agree very well with each other. The calculations of the vibrational states were performed with the sinc-function discrete variable representation (sinc-DVR) method [71, 72]. The vibrational

Preliminary version – January 29, 2016 2.4 results and discussion 23 wave function of CO with v = 15 is shown in Fig. 2.2. With the vibra- tional wave functions obtained, we also calculated the rotational constants Bv of the different vibrational states v, and we compared the vibrational transition frequencies Gv - Gv-1 and rotational constants Bv with the ex- perimental data [73, 74]. For v 6 15, the relative errors in Gv - Gv-1 from the empirical RKR potential derived from these frequencies are less than 0.001%; the errors in the ab initio results based on the RHF-UCCSD(T) po- tential are less than 0.3%. The relative errors in Bv with respect to the ex- perimental values [64, 75] are less than 0.02% for the RKR potential and less than 0.5% for the RHF-UCCSD(T) potential. This assessment shows that the RHF-UCCSD(T)/CBS method yields accurate potentials and that it is still reliable up to r = 3.05 a0. Furthermore, it confirms the accuracy of the semi- classical RKR potential which is used as a building block in our 3D H-CO potential.

2.4.1 Potential surface

Figure 2.3 shows a 2D contour map of the H-CO potential in milli-hartree (mEh) with r fixed at the experimental equilibrium bond length [64] re = 2.1322 a0 of the free CO molecule. The minimum (with depth -26.58 mEh) marked with the black cross occurs at R = 3.007 a0 and ✓ = 144.8. The cor- responding properties of the full 3D potential VHCO are given in Table 2.2. The global minimum in the 3D potential has a depth of -30.70 mEh and corresponds to a stable HCO structure with RCH = 2.115 a0, r = 2.221 a0 and \HCO = 124.6 in valence coordinates. The corresponding Jacobi co- ordinates are R = 3.021 a0 and ✓ = 144.8. This minimum is deeper by 4.12 mEh than the minimum in the 2D potential with r fixed at re, and the CO bond is substantially stretched. Table 2.2 shows that there is also shallow local minimum corresponding to a metastable HOC structure with energy +37.0 mEh. Table 2.3 lists the geometry and height of the barrier to linearity (with ✓ = 180). We find a barrier of 7.961 mEh located at r = 2.232 a0 and RCH = 2.019 a0 in valence coordinates, which corresponds to R = 3.294 a0 in Jacobi coordinates. Our results are in good agreement with previous theoretical values of Werner et al. and Keller et al. [17, 18] and with the values of Austin et al. and Johns et al. [76, 77] extracted from experimental data (see Tables 2.2 and 2.3). It should be kept in mind that the structure extracted from microwave spectra corresponds to a vibrationally averaged geometry, while the theoretical data in these tables refer to the equilibrium geometry. A more direct comparison with experimental data is made when we discuss the rotational constants obtained from 3D DVR calculations in Sec. 2.4.2. Our global HCO minimum is slightly deeper than the minimum in the WKS potential and about equally deep as that in the modified WKS po- tential obtained by adjusting the WKS potential to some experimental data. Also the local minimum at the metastable HOC structure agrees well with that in the modified WKS potential. Our barrier to linearity is somewhat lower than the barriers in both these potentials. Table 2.4 gives the location

Preliminary version – January 29, 2016 24 h-co potential energy surface

180 × ×

50 20 600 20 0 150 400 −20 −10 200 × 0 100 70 −10 × 0 120 50 20 70 90

20 (degrees) × 0 50 θ 100 60 70

100

30 200 70

400 20

50 0 × 2 2.5 3 3.5 4 4.5 5 5.5 6 R (a ) 0

Figure 2.3: Contour plot of the VHCO potential (in mEh) with r fixed at the (ex- perimental) equilibrium bond length re = 2.1322 a0 of free CO. Energy zero is the energy of CO with bond length re and the H atom at infinite distance R. The black cross marks the minimum of depth -26.58 mEh at R = 3.007 a0 and ✓ = 144.8, the green cross the barrier to linearity, and the blue cross the barrier to dissociation of H-CO. The red crosses show the locations of the local maxima related to the conical intersections between the ground electronic state and the first excited state.

Table 2.4: Comparison of the barrier for HCO H + CO dissociation in our poten- ! tial, in the WKS potential [17], and in the modified WKS potential [18], in valence coordinates, as well as the experimental activation energy [78] is given.

WKS modified WKS Our work Exp.

RCH [a0] 3.503 3.534 3.525

r [a0] 2.148 2.148 2.141

\HCO [deg] 117.1117.0116.0

V [mE ]a 6.214 4.608 5.185 3.2 0.6 HCO h ± a With respect to H and CO at infinite distance and CO at its equilibrium bond length.

and height of the dissociation barrier of HCO into H + CO or, equivalently, of the activation barrier of the association reaction. This barrier occurs in

Preliminary version – January 29, 2016 2.4 results and discussion 25

Table 2.2: Comparison of the global HCO and local HOC minima in our poten- tial, in the WKS potential [17], and in the modified WKS potential [18], and experimental values from microwave spectroscopy [76], in valence coordinates.

WKS modified WKS Our work Exp.

Global HCO minimum

RCH [a0] 2.112 2.110 2.115 2.098

r [a0] 2.234 2.233 2.221 2.213

\HCO [deg] 124.5124.5124.6127.4

V [mE ] -28.9 -30.65 -30.70 -30.92 0.3a HCO h ±

Local HOC minimum

ROH [a0] 1.845

r [a0] 2.427

\HOC [deg] 107.5

VHCO [mEh] 38.137.0 a Obtained from the experimental value of D0 [17, 42] and the zero-point vibrational energies of HCO and CO from our calculations. our RHF-UCCSD(T) potential at nearly the same geometry as in the WKS and modified WKS potentials based on multiconfigurational CASSCF and MRCI calculations. Our dissociation barrier is lower than in the WKS poten- tial and slightly higher than in the semiempirically modified WKS potential. The experimental activation energy [78] is lower than all of these theoretical values, but it was already pointed out in Ref. [18] that the uncertainty in the measured value is quite large. To illustrate some of the characteristics of our potential more clearly, we display in Fig. 2.4 two other cuts of the 3D potential, with ✓ fixed at 144.8 (upper panel) and 180 (lower panel). These cuts show the global minimum and the saddle point at linearity. At this point let us return to our discussion of the conical intersections in the first paragraph of Sec. 2.2. Our RHF-UCCSD(T) single-reference method cannot reproduce conical intersections, where the potentials of multiple electronic states become degenerate, but we find a local maximum in our A0 potential close to the location of the lowest conical intersection in the WKS potential of Werner et al. [17]. The height of this maximum is not much lower than the energy of the conical intersection in the WKS potential. Also

Preliminary version – January 29, 2016 26 h-co potential energy surface

Table 2.3: Comparison of the barrier to linearity (\HCO = 180) in our potential, in the WKS potential [17], and in the modified WKS potential [18], and experimental values from electronic absorption spectroscopy [77], in va- lence coordinates.

WKS modified WKS Our work Exp.a

RCH [a0] 2.003 2.007 2.019 2.013

r [a0] 2.250 2.248 2.232 2.234

VHCO [mEh] 12.09.853 7.961

a 2 The linear saddle point of the X A0 state is equivalent to the minimum of 2 the A A00 state at linearity e e at the locations of the other conical intersections found in our CASSCF calcu- lations, we find local maxima in our potential fitted to the RHF-UCCSD(T) calculations (see Fig. 2.3). Furthermore, it was found in Ref. [17] that for lin- ear geometries with shorter C-H bond lengths (and C-O bond lengths larger than the free CO equilibrium distance) the energy of the 2⇧ state is lower 2 than that of the ⌃ state, so that the ground state A0 potential becomes de- generate with the lowest A00 potential. The A00 state goes up in energy when the system becomes nonlinear, and the A0 state goes down, so the lowest energy where these states are degenerate defines the barrier to linearity on the A0 potential. Table 2.3 shows that the location of this barrier in our po- tential based on RHF-UCCSD(T) calculations, is similar to the location of the barrier in the WKS potential based on CASSCF and MRCI calculations. The barrier (0.217 eV = 1747 cm-1 above the H + CO limit) in our potential is 34% lower than the barrier in the WKS potential, but only 19% lower than the barrier in the semiempirically adjusted WKS potential. When the energy of the system gets close to the height of this barrier one must, in principle, include the effects of Renner-Teller nonadiabatic coupling between the two degenerate electronic states. In order to test our potential in the long range, we calculated integral cross sections for rotationally inelastic collisions of ground state CO(j = 0) with H atoms into the final states of CO with j0 = 1, 2, ..., 11. We applied the quantum close-coupling method for collision energies of 400 and 800 cm-1, and, in order to compare the results on our potential with cross sections calculated on the 2D H-CO potentials of Shepler et al. [19], CO was consid- ered as a rigid rotor with r = 2.1322 a0 (or 2.20 a0 for the MRCI potential). The close-coupling equations were solved with the renormalized Numerov propagator with R ranging from 2.4 to 20 a0 in steps of 0.02 a0. The calcula- tions were repeated with the R grid extended to 30 a0 and steps of 0.05 a0, but the results were practically the same. The channel basis included rotor functions of CO up to j = 20, and the calculations were made for overall an-

Preliminary version – January 29, 2016 2.4 results and discussion 27

° θ = 144.8 2.8 50 100

2.6 100 10 50 200 −10 2.4 −20 )

0 0 −30 10

r (a 2.2

2 0 10 10 50 50 1.8 100 100 300 200 200 ° θ = 180 2.8 100 100 2.6 200 3000 500 2.4 2000 ) 20 50 0 300 10 20

r (a 2.2 10 2 1000 50 20 100 50 100 1.8 200 200 300 2 2.5 3 3.5 4 4.5 5 5.5 6 R (a ) 0

Figure 2.4: Contour plots of the VHCO potential (in mEh) with ✓ fixed at the equilib- rium value of 144.8 (upper panel) and at 180 (linear geometry, lower panel). gular momenta J = 0 - 60. Figure 2.5 shows for H-CO(j = 0 j ) scattering ! 0 that the integral cross sections on our potential indeed show the predicted [79] propensity for even j transitions and are in close agreement with the cross sections calculated on the 2D CCSD(T) and MRCI potentials of Shepler et al. [19]. Since the latter potentials were also calculated with special atten- tion for the accuracy in the long range, this confirms that also our potential is accurate in that region.

2.4.2 Vibrational frequencies and rotational constants of HCO

A series of vibrational frequencies of HCO was obtained from laser photo- electron spectroscopy of the formyl anion by Murray et al. [41] and with the use of dispersed fluorescense and stimulated emission pumping spec-

Preliminary version – January 29, 2016 28 h-co potential energy surface

12 )

2 our potential 10 CCSD(T) potential MRCI potential 8 − 6 E = 400 cm 1

4

2 Integral Cross Section (Å 0 12 )

2 our potential 10 CCSD(T) potential MRCI potential 8 − 6 E = 800 cm 1

4

2 Integral Cross Section (Å 0 0 1 2 3 4 5 6 7 8 9 10 11 final j’

Figure 2.5: Integral cross sections for inelastic H-CO(v = 0, j = 0 v = 0, j ) ! 0 scattering at collision energies of 400 cm-1 (upper panel) and 800 cm-1 (lower panel), obtained from close-coupling calculations in the rigid rotor approximation. For comparison with cross sections calculated on the two-dimensional potentials of Shepler et al. [19] the CO bond length r was fixed at 2.1322 a0 in our and in the CCSD(T) potential, while in the MRCI potential r was fixed at 2.20 a0.

troscopy of HCO by Tobiason et al. [37]. The ground-state rotational and distortion constants were measured by microwave spectroscopy [76] and, later, by magnetic resonance far-infrared laser spectroscopy [80, 81]. The best way to check our 3D H-CO potential is to use it in accurate calculations of the rovibrational states of HCO and directly compare the results with these experimental data. Actually, the modified BBH potential [35, 36] is a scaled version of the ab initio BBH potential [34], modified such that the vibrational frequencies agree better with the experimental data of [41, 43]. Similarly, the modified WKS potential [18] was obtained from the ab initio WKS potential [17] by scaling to improve agreement with the more recent vi- brational frequencies of Ref. [37]. The method that we used to compute the vibrational states of HCO is the 3D DVR method in Jacobi coordinates with the program of Tennyson et al. [82]. The radial Gauss-Laguerre quadrature grids contained 20 points in the r coordinate and 25 points in R. For the

Preliminary version – January 29, 2016 2.4 results and discussion 29 angle ✓ we used 45 point Gauss-Legendre quadrature. In tests with smaller and larger numbers of DVR grid points we confirmed that the vibrational levels are converged to about 0.001 cm-1 for the lower levels and to about 0.005 cm-1 for the levels at 5000 cm-1 above the zero-point level. An important result from these calculations is the vibrational zero-point energy of HCO. We mentioned above that the binding energy De relative to the free H atom and the free CO molecule at its equilibrium geometry

Table 2.5: Calculated and experimental vibrational term energies (in cm-1) of the 2 a bound levels of X A0 HCO (J = 0) .

e Assignment Experimental Calculated Deviation

(0, 0, 1) 1080.76 1079.6 -0.11%

(0, 1, 0) 1868.17 1871.60.18%

(0, 0, 2) 2142 2145.20.15%

(1, 0, 0) 2434.48 2428.5 -0.25%

(0, 1, 1) 2942 2950.00.27%

(0, 0, 3) 3171 3186.20.48%

(1, 0, 1) 3476 3464.1 -0.34%

(0, 2, 0) 3709 3719.00.27%

(0, 1, 2) 3997 4013.10.40%

(0, 0, 4) 4209* 4195.7 -0.32%

(1, 1, 0) 4302 4301.5 -0.01%

(1, 0, 2) 4501* 4478.7 -0.50%

(2, 0, 0) 4570* 4546.0 -0.53%

(0, 2, 1) 4783.24797.20.29% aMost of the experimental data are from Tobiason et al. [37], the data marked with * are from Rumbles et al. [44]. The assignment (v1, v2, v3) is in terms of the CH stretch level v1, the CO stretch level v2, and the \HCO bend quantum number v3. is 30.6973 mEh = 0.8353 eV. The lowest level obtained from the 3D DVR calculations lies 2794.95 cm-1 = 0.3465 eV above the global minimum in the potential, in other words the vibrational zero-point energy of HCO is 0.3465 eV. Since we also calculated the zero-point vibrational energy of free

Preliminary version – January 29, 2016 30 h-co potential energy surface

CO, 0.1341 eV, we obtain the dissociation energy D0 of H-CO: 0.623 eV. Two values of the H-CO bond enthalpy at 298 K have been measured by Chuang et al. [42] and by Walsch, quoted in Ref. [17]. If these enthalpies are converted [17] to 0 K bond energies, the D values are 0.629 0.008 0 ± eV and 0.603 0.009 eV, respectively. Our ab initio value agrees well with ± these data. The WKS potential gave D0 = 0.569 eV [17] and the empirically modified WKS potential gave D0 = 0.620 eV [18]. The vibrational frequencies from the 3D DVR calculations for total angular momentum J = 0 are listed and compared with the most recent experimen- tal data [37] in Table 2.5. The agreement with the measured frequencies is good, nearly as good as for the modified WKS potential and better than for the modified BBH potential—both these potentials were scaled to get agree- ment with the experimental vibrational frequencies—and substantially bet- ter than with the original ab initio WKS and BBH potentials. Furthermore, we computed the levels for higher rotational states of HCO, up to J = 6. Then, we extracted the rotational constants A, B, C and the centrifugal distortion constants J, JK, K of HCO in its vibrational ground state from a fit of the rotational levels to the eigenvalues of the usual semirigid rotor Hamiltonian for a prolate near-symmetric top including centrifugal distortion

1 2 1 1 H = (B + C) J + A - (B + C) J2 + (B - C) J2 + J2 rot 2 2 z 4 + -  ⇣ ⌘ 4 2 2 4 b + J J + JKb J Jz + K Jz, b b b (2.9)

where J is the totalb angularb b momentumb operator, Jz its projection on the long (a) axis, and J are step-up and step-down operators. The approxi- ± mate quantumb number K is the eigenvalue of Jz. Web did not include off- diagonal centrifugalb distortion terms, because their effect on the energy levels is smaller than the remaining convergenceb error of the rotational lev- els calculated. It turned out that also J and JK are very small (as they are experimentally [80]) and not well converged; the distortion constant K is significant, however. Table 2.6 gives the results and their comparison with the experimental data [80, 81]. Small variations occur in the values of A, B, and C fit to our rotational levels, depending on the highest J value for which these levels were included in the fit. Also spin-orbit coupling constants were fitted to 2 the measured spectra [80, 81] of ground state HCO(X A0), but we did not include spin-orbit coupling in our calculations of the rovibrational levels, so we cannot compare these parameters. The rotationale constants A, B, C and the relatively large centrifugal distortion constant K extracted from the rotational levels agree very well with the experimental data. It is easily checked, for example, that the small discrepancy of 0.5% in the value of A corresponds to a change of \HCO of only 0.1. Also the rotational constants calculated with the assumption that HCO is fixed at its equilibrium geome- try, with RCH = 2.115 a0, r = 2.221 a0, and \HCO = 124.6, are included in Table 2.6. The deviations between these equilibrium geometry values and

Preliminary version – January 29, 2016 2.5 conclusions 31

Table 2.6: Calculated and experimental [80, 81] rotational and distortion constants -1 2 (in cm ) of X A0 HCO in its vibrational ground (v1, v2, v3)=(0, 0, 0) state. e

3 Jmax Kmax ABC10 K Fit error

44 24.4804 1.4945 1.4030 29.4453 0.012

55 24.4640 1.4955 1.4045 27.9362 0.041

66 24.4392 1.4971 1.4066 26.3809 0.100

Experiment 24.3296 1.4940 1.3987 31.4042(57)

Equilibrium structure 24.2667 1.4968 1.4099

the values from the full 3D calculations are larger than the differences be- tween the latter and the experimental values. This illustrates that the direct comparison of the rotational constants calculated for the vibrational ground state with the corresponding experimental data is to be preferred over the comparison of geometric parameters extracted from spectroscopic data, as shown in Table 2.2. The agreement with the experimental vibrational fre- quencies and rotational constants, obtained directly with our 3D ab initio potential surface without any adaptation, demonstrates that this potential is indeed accurate.

2.5 conclusions

We constructed an accurate analytical 3D potential energy surface for the 2 H-CO(X A0) complex based on high-level RHF-UCCSD(T)/CBS ab initio calculations. Both by extending the calculated data points to large values of the H-COe distance R and by taking special care in the fit procedure, we ensured that this potential is also accurate in the long-range region. This will be important for the application of the potential in calculating H-CO inelastic collision cross sections, especially at lower energies. Integral cross sections for rotationally inelastic H-CO(j = 0 j ) collisions were obtained ! 0 with the close-coupling method. The CO molecule was considered as a rigid rotor, in order to be able to compare the results with cross sections calcu- lated on two different 2D potentials for H-CO. The agreement is very good, which confirms that our potential is accurate in the long-range. Further- more, we applied our potential in calculations of the rovibrational states of HCO with the use of a 3D DVR method. The ab initio dissociation energy D0 = 0.623 eV obtained from this calculation agrees well with experimen- tal data. Also, the vibrational levels resulting from the DVR calculations are

Preliminary version – January 29, 2016 32 h-co potential energy surface

in good agreement with experimental data, better than those from previ- ous ab initio potentials and nearly as good as those from the semiempirical “modified WKS potential” that was scaled for better agreement with these vi- brational levels. From the rotational levels of HCO in its vibrational ground state for total angular momentum J = 0 - 6 we extracted the rotational con- stants A, B, C, and the dominant centrifugal distortion constant K. Also, they agree well with measured data. This confirms that our potential is also accurate in the region of the potential well. Further studies are planned to investigate vibration-rotation inelastic col- lision processes. Integral and differential cross sections will be obtained from full three-dimensional close-coupling calculations. The differential cross sections will be compared with results from crossed beam experiments com- bined with velocity map imaging that will soon be performed. The integral cross sections and rate constants will be used in modeling astrophysical processes.

Preliminary version – January 29, 2016 H-CO RATE COEFFICIENTS FOR PURE ROTATIONAL TRAN-3 SITIONS

Carbon monoxide is a simple molecule present in many astrophysical en- vironments, and collisional excitation rate coefficients due to the dominant collision partners are necessary to accurately predict spectral line intensities and extract astrophysical parameters. In this chapter, we report new quan- tum scattering calculations for rotational deexcitation transitions of CO in- duced by H using the three-dimensional potential energy surface described in Chap. 2. State-to-state cross sections for collision energies from 10-5 to 15, 000 cm-1 and rate coefficients for temperatures ranging from 1 to 3000 K are obtained for CO(v = 0, j) deexcitation from j = 1 - 45 to all lower j0 lev- els, where j is the rotational quantum number. Close-coupling and coupled- states calculations were performed in full-dimension for j = 1 - 5, 10, 15, 20, 25, 30, 35, 40, and 45 while scaling approaches were used to estimate rate coefficients for all other intermediate rotational states. The current rate co- efficients are compared with previous scattering results using earlier PESs. Astrophysical applications of the current results are briefly discussed.

Kyle M. Walker, L. Song, B.H. Yang, G. C. Groenenboom, A. van der Avoird, N. Balakrishnan, R. C. Forrey, P. C. Stancil, Quantum calculations of inelastic CO collisions with H. II. Pure Rotational Quenching of High Rotational Levels, Astrophys. J. 811, 27 (2015)

Preliminary version – January 29, 2016 34 h-co rate coefficients for pure rotational transitions

3.1 introduction

Carbon monoxide is the second most abundant molecule in the universe after molecular hydrogen and is found in a variety of astrophysical envi- ronments. The formation of CO in low-density interstellar clouds proceeds mainly through gas-phase chemical reactions involving H2, and therefore detecting and measuring CO spectral lines is one way to trace the molec- ular interstellar medium (ISM) in the region. While H2 rotational lines are difficult to observe from the ground, CO is more easily detected as its rovi- brational transitions can be observed as absorption in the UV and near- infrared (NIR) and emission in the NIR, far-infrared (FIR), and submillime- ter. The rotational energy level spacings of a diatomic molecule depend in- versely on its moment of inertia, and the relatively large moment of inertia of CO compared to H2 yields rotational levels with small energy separa- tion. The lowest rotational transition of CO has a wavelength of ⇠ 2.6 mm which gives an excitation temperature of just ⇠ 5.5 K. Therefore, CO can be collisionally excited to high rotational levels in photon-dominated regions (PDRs), in shocks, and in other moderately energetic environments. A number of CO pure rotational lines starting at j = 14 13, where j ! is the rotational quantum number, have been detected with the Long Wave- length Spectrometer (LWS)[83] aboard the Infrared Space Observatory [ISO, 84]. These include the FIR spectra of all the lines up to j = 39 38 in ! the carbon-rich circumstellar envelope IRC +10216 [85], transitions up to j = 25 24 in VY Canis Majoris and other oxygen-rich circumstellar en- ! velopes [86], transitions up to j = 24 23 from the carbon-rich planetary ! NGC 7027 [87], and transitions up to j = 19 18 of the Herbig Haro ! objects HH 52 - 53 - 54 and IRAS 12496 - 7650 in a nearby star-forming re- gion [88]. The KOSMA 3m and IRAM 30m telescopes in Switzerland and Spain, respectively, performed observations of the low-j transitions of CO in the Rosette Molecular Complex [89], while the Odin Orbital Observatory [90] detected j = 5 4 emission from the photon-dominated region (PDR) ! of Orion KL [91]. More recently, the Herschel Space Observatory [92] observed the j = 9 8 line towards Monoceros R2 [93] and high-j CO lines in the ! NGC 1333 low-mass star-forming region [94] with the Heterodyne Instru- ment for the Far-Infrared (HIFI) [95]. The Spectral and Photometric Imag- ing REceiver (SPIRE) [96] aboard Herschel has also probed the submillimeter molecular interstellar medium of M82 from j = 4 3 up to j = 13 12 ! ! [97]. The chemical and physical conditions where CO resides are deduced from observed spectral lines, but it is often not appropriate to assume lo- cal LTE when modeling these regions. In low-density environments where the level populations routinely depart from LTE, collisional excitation rate coefficients with the dominant species — mainly H2, H, He, and electrons — are necessary to accurately predict spectral line intensities. Although the effect of electron collisions is minor, the role of collisional excitation of CO by H, especially in environments such as diffuse molecular clouds [98] and

Preliminary version – January 29, 2016 3.1 introduction 35 cool mixed atomic and molecular hydrogen gas [33], cannot be neglected. Rate coefficient calculations for pure rotational excitation of the first 15 ro- tational levels of the He-CO system were performed by Cecchi-Pestellini et al. [20] for the temperature range 5 to 500 K. Yang et al. [27] calculated reliable H2-CO rate coefficients for rotational transitions in CO induced by both para- and ortho-H2 collisions for deexcitation from j = 1 - 40 to all lower j0 levels for temperatures between 1 and 3000 K. Although the H-CO system has been studied extensively with the most recent rate coefficients reported by Yang et al. [15], there is unsatisfactory agreement between the results obtained upon the various PESs. The earliest analysis of H-CO collisions was performed by Chu & Dal- garno [99], who used a short-range, semi-empirical potential joined with a long-range Buckingham potential to calculate cross sections via two close- coupling formulations. The resulting cross sections were found to be com- parable to those of H2-CO for small j. The Maxwellian-averaged rate coeffi- cients were calculated for the temperature range of 5 - 150 K. The following year, quantal calculations were carried out by Green & Thaddeus [100] over the same temperature range, but with a new semi-empirical potential and strikingly different results. These collisional rate coefficients were typically an order of magnitude smaller than those of Chu & Dalgarno [99], and small j transitions, i.e., |j| = 1 or 2, were reported to have much larger rate coefficients when compared to large j transitions. The small magni- tude of the H-CO rate coefficients obtained by Green & Thaddeus [100] led to the neglect of H as a collider in CO emission modeling. The next interaction PES for the H-CO system followed a more sophisti- cated level of theory using ab initio calculations for the surface [34], hereafter called BBH. These calculations employed the Dunning & Hay [101] valence double-zeta contractions of the Huzinaga [102], (9s,5p) sets of carbon- and oxygen-centered primitive Gaussians and used the (4s,2s) contraction with a scale factor of 1.2 for hydrogen. RHF calculations were carried out, fol- lowed by configuration interaction computations including all singly- and doubly-excited configurations. The resulting empirically corrected potential was used in dynamics calculations using a cubic spline interpolation. Sub- sequent coupled-channel scattering calculations were carried out on this PES by Lee & Bowman [79] who first noticed the propensity for even-j transitions for inelastic scattering. Another surface was constructed by Werner et al. [17], hereafter WKS, to see if new experimental results could be more closely reproduced. The ab initio electronic structure calculations were performed with the MOLPRO program package [57] using the internally contracted multireference config- uration interaction (icMRCI) method. Resonance energies and widths from the WKS surface agreed with experimental data [18], but striking differences were apparent in the collision cross sections obtained using the BBH and WKS surfaces [49]. The cross sections on the newer WKS surface, especially for low-j pure rotational transitions, yielded higher values than those calcu- lated on the BBH surface. Balakrishnan et al. [16] obtained similar results

Preliminary version – January 29, 2016 36 h-co rate coefficients for pure rotational transitions

when they used both the BBH and the WKS surfaces to compute collisional rate coefficients for temperatures in the range 5 to 3000 K. The discrepancy in the CO(j = 1 0) transition was the largest. However, Balakrishnan et ! al. [16] performed explicit quantum-mechanical scattering calculations with the WKS PES for the first eight pure rotational transitions (using the close- coupling (CC) framework) and first five vibrational levels (in the infinite or- der sudden (IOS) approximation), and this complete set of rate coefficients has been extensively used in astrophysical models. Shepler et al. [19] revisited H-CO scattering by introducing two new rigid- rotor PESs and computing cross sections on each surface at collision ener- gies of 400 and 800 cm-1. One ab initio surface was calculated using the coupled cluster method with single and double excitations and a pertur- bative treatment of triple excitations [CCSD(T)] employing the frozen core approximation [103]. RHF orbitals were used for the open-shell calculations and the spin-restrictions were relaxed in the solution of the coupled clus- ter equations [R/UCCSD(T)] [54]. The other PES was computed using the CASSCF [51] and icMRCI [104] method with the aug-cc-pVQZ basis set for H, C, and O by Woon & Dunning [105] as implemented in the MOLPRO program. Their results for both potentials are more similar to Green & Thad- deus [100] and it was recommended that the rate coefficients calculated by Balakrishnan et al. [16] using the inaccurate WKS surface be abandoned. Based on this recommendation, astrophysical models using the overesti- mated rate coefficients should also be reexamined. The most recent set of H-CO calculations were performed by Yang et al. [15] and extend the cal- culations of Shepler et al. [19] on the MRCI rigid-rotor PES. State-to-state rotational deexcitation cross sections and rate coefficients from initial states j = 1 - 5 in the ground vibrational state to all lower j0 levels were computed. In this chapter, we report new quantum scattering calculations for rota- tional deexcitation transitions of CO induced by H using the 3D interac- tion PES of Song et al. [106]. This 3D ab initio PES uses the spin-unrestricted open-shell single and double excitation coupled-cluster method with per- turbative triples [UCCSD(T)] with molecular orbitals from RHF calculations. Electronic structure calculations were performed with the MOLPRO 2000 package [57] and 3744 interaction energies were included in the fit. When compared to previous PESs, the new 3D surface uses the most sophisticated level of theory and is the most accurate available for scattering calculations. State-to-state cross sections for collision energies from 10-5 to 15, 000 cm-1 are computed for CO(v = 0, j) deexcitation from initial states j = 1 - 45 to all lower j0 levels. While close-coupling and coupled-states calculations, in full-dimension, are performed for j = 1 - 5, 10, 15, 20, 25, 30, 35, 40, and 45, scaling approaches are used to estimate the rate coefficients for all other intermediate rotational states. Rate coefficients for temperatures rang- ing from 1 to 3000 K are evaluated and compared with previous scattering results obtained on earlier surfaces. The astrophysical implications of new rotational deexcitation rate coefficients are illustrated.

Preliminary version – January 29, 2016 3.2 computational approach 37

3.2 computational approach

We performed inelastic scattering calculations on the most current 3D PES 2 of the H-CO complex in the ground state (X A0)[106]. Computations were carried out using the quantum mechanical CC method [107] for kinetic ener- gies below 1000 cm-1. From 1000 cm-1 to 15e, 000 cm-1 the coupled state (CS) approximation of McGuire & Kouri [108] was utilized. We treated hydro- gen as a structureless atom and allowed the bond length to vary for CO. The full 3D surface is used in the calculations; at no point is the CO bond length fixed nor a rigid-rotor approximation made. The interaction poten- tial was expressed as V(R,r,✓), where r is the CO intramolecular distance, R the distance from the CO center of mass to the H nucleus, and ✓ the angle between the vector R and the CO bond axis, where linear C-O-H has an angle of zero degrees and ✓ = 180 for linear H-C-O. The potential was expanded according to

max V(R, r, ✓)= (R, r)P(cos ✓),(3.1) X=0 where P are Legendre polynomials of order . The radial dependence of the potential used a 20-point Gauss-Hermite quadrature over r to represent CO stretching. The angular dependence of the potential was expanded to max = 20 with a 22-point Gauss-Legendre quadrature. The quantum-mechanical CC calculations were performed using the mixed- mode OpenMP/MPI version of the nonreactive scattering program MOLSC- AT [109] modified by Valiron & McBane [110] and Walker [111]. The mod- ified log-derivative Airy propagator of Alexander & Manolopoulos [112] with variable step size was used to solve the coupled-channel equations. The propagation was carried out from R = 1 a0 to a maximum distance of R = 100 a0, where a0 is the atomic unit of length (the Bohr radius). For calculations with initial state j<25, the basis set included 31 rotational lev- els in the ground vibrational state of CO. For initial states j > 25 the basis sets included at least 5 - 10 closed rotational levels. Only pure rotational transitions within the v = 0 vibrational level are reported here. A conver- gence study was performed with excited vibrational levels in the basis set, but the addition of higher vibrational levels yielded results within 2%. A sufficient number of angular momentum partial waves were included to en- sure convergence to within 10% for the largest state-to-state cross sections. For the low energy range, 10-5–10 cm-1, the number of maximum partial waves, Jmax, was increased each decade resulting in the values 2, 4, 8, 12, 14, and 20 being added to the value of initial j, respectively, while converged cross sections for energies from 10 - 49, 50 - 95, 100 - 900, 1000 - 9000, and 10, 000 - 15, 000 cm-1 were calculated by adding the values of 26, 30, 40, 50, and 60 to that of the initial state, respectively.

Preliminary version – January 29, 2016 38 h-co rate coefficients for pure rotational transitions

The degeneracy-averaged-and-summed integral cross section for a rota- tional transition from an initial state j to a final state j0 in the CC formalism is given by

Jmax J+j J+j0 ⇡ J 2 j j (Ej)= (2J + 1) |jj ll - S (Ej)| ,(3.2) 0 2 0 0 jj0ll0 ! (2j + 1)kj XJ=0 l=X|J-j| l0=X|J-j0| where j is the rotational angular momentum of the CO molecule, l is the orbital angular momentum of the collision complex, and J = j + l is the total J angular momentum. S is an element of the scattering matrix, Ej is the jj0ll0 relative kinetic energy of the initial channel, and kj = 2µ(E - ✏j)/h is the wave-vector of the initial channel, where µ is the reduced mass of the H-CO p system (0.97280 u), E is the total energy, ✏j is the rotational energy of CO, and h is the reduced Planck constant. The CS approximation, on the other hand, reduces the computation expense by neglecting the Coriolis coupling between different values of ⌦, the projection of the angular momentum quantum number of the diatom along the body-fixed axis, and within this formalism the integral cross section is given by

Jmax ⌦max ⇡ J⌦ 2 j j (Ej)= (2J + 1) (2 - ⌦0)|jj - S (Ej)| ,(3.3) 0 2 0 jj0 ! (2j + 1)kj XJ=0 ⌦X=0

where ⌦max is equal to 0, 1, 2, ..., max(J,j). State-to-state cross sections for collision energies from 10-5 to 15, 000 cm-1 were computed for CO(v = 0,j) deexcitation from initial state j = 1 - 5, 10, 15, 20, 25, 30, 35, 40, and 45 to all lower j0 levels. Deexcitation rate coef- ficients ranging from 1 to 3000 K were obtained by averaging the cross sections over a Boltzmann distribution of collision energies,

1/2 8kbT 1 kj j (T)= j j (Ej) exp(-Ej/kbT)EjdEj,(3.4) ! 0 ⇡µ (k T)2 ! 0 ✓ ◆ b Z01

where j j is the state-to-state rotationally inelastic cross section, Ej = ! 0 E - ✏j is the center of mass kinetic energy, and kb is the Boltzmann constant.

3.3 results and discussion

Before scattering calculations were carried out, we compared the RHF-UCC- SD(T) 3D ab initio PES of Song et al. [106] with the 2D rigid rotor MRCI and CCSD(T) surfaces of Shepler et al. [19]. The lower order Legendre terms of the RHF-UCCSD(T) PES agree well with the CCSD(T) terms; the similar high level of theory in both CCSD(T) calculations is expected to yield a more accurate potential than the lower-level MRCI calculations (see Fig. 3.1). The higher-order -terms are typically small and decrease with . For example, for the RHF-UCCSD calculations shown in Fig. 3.1, 3 ⇠ 0/100. Further, as the cross sections for a direct transition are driven by where |j| = ,

Preliminary version – January 29, 2016 3.3 results and discussion 39

10 10.0 λ = 0 λ = 1 0 7.5

−10 5.0 ) -1 −20 2.5 (cm λ V

−30 0.0

−40 RHF-UCCSD(T), Song et al.(2013) -2.5 CCSD(T), Shepler et al.(2007) MRCI, Shepler et al.(2007) −50 -5.0 67891011126789101112 15.0 0.2 λ = 2 λ = 3 12.5

10.0 0.1

7.5 ) -1 0.0

(cm 5.0 λ V 2.5

-0.1 0.0

-2.5 -0.2 678910111289101112

R(a0) R(a0)

Figure 3.1: The first four Legendre expansion terms v(R) near the van der Waals well of the H-CO potential energy surface with the CO intermolecular distance fixed at r = 2.20 a0 for MRCI and the equilibrium bond length re = 2.1322 a0 for the CCSD(T) and RHF-UCCSD(T) surfaces. they are expected to decrease with |j| and as a consequence the uncertainty in the cross sections are expected to increase with |j|, though multi-step transitions due to smaller terms will also contribute. The values of the Legendre expansion terms of the MRCI potential are all greater than those of the RHF-UCCSD(T) and CCSD(T) PESs. The feature seen in the MRCI results near R = 10 a0 is due to the poor quality of fitting where the long–range and short–range regions are joined. Since the depth and location of the van der Waals well may strongly influence rotationally inelastic scattering, the behavior of the three interaction potentials was closely examined in the vicinity of the van der Waals well. Figure 3.2 shows a contour plot of the long-range behavior of the interaction potential of Song et al. [106] with the CO bond length fixed at the equilibrium distance of re = 2.1322 a0. The positions and values of the local and van der Waals minima on the three interaction PESs were calculated and are shown in Table 3.1. The van der Waals well of the MRCI potential is significantly different from that of the CCSD(T) and RHF-UCCSD(T) being both shallower and located at a larger internuclear distance. The differences in van der Waals well depth and anisotropies of the surfaces lead to differences in the scattering results. Figure 3.3 shows state-to-state pure rotational H-CO rate coefficients from initial state CO(j = 5) to indicated lower states j0. The largest rate coef- ficients of CO(5 3) as calculated by Chu & Dalgarno [99], Green & ! Thaddeus [100], and Balakrishnan et al. [16] are shown. Rate coefficients

Preliminary version – January 29, 2016 40 h-co rate coefficients for pure rotational transitions

Table 3.1: Minima of the interaction PESs. -1 PES rCO(a0) Vmin(cm ) R(a0) ✓ (degrees) Global minimum:a MRCIb 2.20 -6934.052 3.01 145.5 CCSD(T)b 2.1322 -5817.161 3.00 144.7 RHF-UCCSD(T)c 2.1322 -5832.994 3.01 144.8 RHF-UCCSD(T)c 2.20 -7256.805 3.02 144.8 van der Waals minimum: MRCI 2.20 -19.934 7.21 110.8 CCSD(T) 2.1322 -34.677 6.87 108.1 RHF-UCCSD(T) 2.1322 -35.269 6.86 108.0 RHF-UCCSD(T) 2.20 -36.296 6.84 108.5 a The term "global" is used loosely here since rCO has been frozen at an unoptimized distance to ease direct comparison with the previous two-dimensional surfaces. b Shepler et al. [19] c Song et al. [106]

from Yang et al. [15] on the MRCI PES of Shepler et al. [19] are also shown. The rate coefficients in this chapter agree well with those of Chu & Dal- garno [99] and Yang et al. [15]. On average, rate coefficients of Green & Thaddeus [100] are less than those of this chapter by around an order of magnitude, while the values from Balakrishnan et al. [16] are around twice as much as ours. Although the CO(5 3) transition is highlighted here, ! the results are typical for other state-to-state transitions as well. Figures 3.4 and 3.5 present sample results from our computations of cross sections and rate coefficients, respectively, from initial state j = 10 to all final states j0. In general, the cross sections for j0 = 0 are smallest and then increase with increasing j0 up to the dominating transition where |j| = |j0 - j| = 2 after which they decrease. Since CO is near-homonuclear, odd-j transitions are suppressed and the transitions follow an even-j propensity. At 1000 cm-1, the difference between the CC and CS cross sections is less than ⇠ 5%, except for the largest changes in j where the differences are typically ⇠ 10%. While the quenching from selected high rotational states is explicitly cal- culated, a zero-energy scaling technique (ZEST) is used to predict state-to- state rate coefficients for all intermediate states. Figure 3.6 shows the total quenching and state-to-state cross sections calculated at 10-5 cm-1 as a function of increasing j. The total quenching rate coefficient in the ultra- cold limit (collision energy of 10-5 cm-1) is fairly constant for j>10. Note that j =-2 dominates for low initial j, as expected, but at j ⇠ 22, j =-1 begins to dominate. From these ultracold state-resolved quenching rate co- efficients, we estimated the rate coefficients for all temperatures following

Preliminary version – January 29, 2016 3.3 results and discussion 41

180

150

120 -35

90 -30 -25 -20 -15 -10 θ (degrees) 60

30 -15 -5 -25 -20 -10 0 0 678910

R(a0)

Figure 3.2: The long-range behavior of the interaction potential of Song et al. [106] -1 with VHCO in cm and r fixed at the equilibrium bond length re = 2.1322 a0. the procedure introduced in Yang et al. [113], but modified here. The un- known rate coefficient for a transition from j to j - j for any temperature T, given by k(j, j - j; T), can be estimated based on the rate coefficients from calculated transitions for other initial rotational states, for example, larger j (i.e., above, jA) or smaller j (i.e., below, jB)[k(jA, jA - j; T) and k(jB, jB - j; T), respectively] and calculated ultracold (T ⇠ 0) rate coeffi- cients for the transitions above, below, and the desired according to

k(j, j - j; T)=k(j, j - j; T = 0) ⇥ w k(j , j - j; T)+w k(j , j - j; T) (3.5) A A A B B B . wAk(jA, jA - j; T = 0)+wBk(jB, jB - j; T = 0)

Weights wA and wB for the calculated rate coefficients were determined by considering the change in initial j in the desired transition when compared to the states above and below according to the equations

j - jB jA - j wA = wB = .(3.6) jA - jB jA - jB For example, the predicted j = 16 rate coefficients used calculated transi- tions from j = 20 and j = 15 and weights wA = 0.2 and wB = 0.8. The cross sections at ultracold energies are easily calculated and simply multiplied by the velocity of the system (obtained from the kinetic energy of the collision)

Preliminary version – January 29, 2016 42 h-co rate coefficients for pure rotational transitions

3 10−10 4 2 ) -1 s

3 1 10−11

0

10−12

j=5→3, Chu & Dalgarno(1975) j=5→3, Green & Thaddeus(1976) Rate Coefficient 10(cm −13 j=5→3, Balakrishnan et al.(2002) j=5→0 j=5→1 j=5→2 Solid lines - This work j=5→3 Dashed lines - Yang et al.(2013) j=5→4 10−14 100 101 102 103 Temperature (K)

Figure 3.3: State-to-state pure rotational rate coefficients due to H collisions from initial state CO(j = 5) to indicated lower states j0 from this chapter using the Song et al. [106] potential energy surface, and from Yang et al. [15] on the MRCI potential energy surface of Shepler et al. [19]. Previous rate coefficient calculations of Chu & Dalgarno [99], Green & Thaddeus [100], and Balakrishnan et al. [16] on other surfaces are also shown for the dominant transition j = 5 3. !

to obtain the T ⇠ 0 rate coefficients. A comparison of this zero-energy scal- ing technique with explicit calculations is given in Fig. 3.7 for initial state j = 4. The predicted rate coefficients agree well with the calculated values, although in this case transitions from j = 3 and j = 5 are equally weighted and a slight underestimate of the j = 4 - 2 transition occurs. Given the high accuracy of the current PES, any uncertainty remaining in the computed cross section is likely due to the adoption of the CS approximation, the truncation in the Legendre expansion of the PES to max = 20 and the un- certainty in for large values of , and limited basis set sizes for collision energies above 1000 cm-1. Otherwise, the interpolation formula introduces the largest source of uncertainty.

3.4 astrophysical applications

The cross sections and/or rate coefficients calculated in this chapter are available online.1 These high-j pure rotational rate coefficients are especially

1Rate coefficient data in the Leiden Atomic and Molecular Database [LAMDA, Schoöier et al. 30] format, as well as cross section data can be obtained at www.physast.uga.edu/amdbs/excitation/.

Preliminary version – January 29, 2016 3.4 astrophysical applications 43

3 To t al 10 j=10→0 j=10→1 j=10→2 Total j=10→3

) j=10→4 2 j=10→5 8 j=10→6 cm j=10→7 9

-16 j=10→8 6 j=10→9 7 100 5 4 3 2 Cross Section (10 Section Cross 1 −3 10 0

10−5 10−4 10−3 10−2 10−1 100 101 102 103 104 Energy (cm-1)

Figure 3.4: State-to-state pure rotational deexcitation cross sections due to H colli- sions from initial state CO(j = 10) to all lower states j0.

10−9 Total 10−10 8 )

-1 −11 9

s 10 6 3 7 10−12 5 4 −13 10 3 To t al j=10→0 2 j=10→1 −14 j=10→2 10 j=10→3

Rate Coefficient (cm 1 j=10→4 j=10→5 j=10→6 −15 10 j=10→7 0 j=10→8 j=10→9 10−16 100 101 102 103 Temperature (K)

Figure 3.5: State-to-state pure rotational deexcitation rate coefficients due to H collisions from initial state CO(j = 10) to all lower states j0.

Preliminary version – January 29, 2016 44 h-co rate coefficients for pure rotational transitions

104 Total

∆j=-2 ∆j=-3 ) 3 2 10

cm ∆j=-4 -16 ∆j=-1

102

1

Cross Section (10 Section Cross 10

E = 10-5 cm-1 100 010203040 Initial j

Figure 3.6: State-resolved cross sections at 10-5 cm-1 as a function of increasing j for the H-CO system. Circles indicate state-to-state cross sections while diamonds indicate total quenching cross sections. Each series (circles) correspond to different values of j, decreasing from the top at the far right as j =-1, -2, -3, -4, ... (red, green, blue, magenta).

3-2 10−10 3-1 4-3 est. 4-2 est. ) 4-3 calc. -1

s 4-2 calc. 3 5-4 5-3 ∆ j = -2 10−11

Rate Coefficient (cm ∆ j = -1 −12 10 Zero-Energy Scaling Calculated: j = 3,4,5 Estimated: j = 4

10−4 10−3 10−2 10−1 100 101 102 103 Temperature (K)

Figure 3.7: A comparison of the zero-energy scaling technique with explicit calcu- lations of the largest deexcitation rate coefficients for initial state j = 4. Solid lines indicate j =-1 while dashed lines are j =-2. Initial states are: j = 3 (black), j = 4 estimated from Equation 3.5 (green with symbols), j = 4 explicitly calculated (red), and j = 5 (blue).

Preliminary version – January 29, 2016 3.4 astrophysical applications 45

7 10 j = 30

6

) 10 -3

105

104

Critical Density (cm Density Critical 3 10 j = 1

102 101 102 103 Temperature (K)

Figure 3.8: Critical densities for CO(v = 0, j) due to H collisions as a function of gas temperature T. useful as input data for codes developed to solve for level populations. One such code is RADEX [31], which can perform a non-LTE analysis of interstel- lar line spectra. The reliable rate coefficients calculated in this chapter also will help extract more accurate astrophysical conclusions from current mod- els. Recently, Thi et al. [9] modeled protoplanetary disks (PPDs) adopting the rate coefficients computed by Balakrishnan et al. [16]. These rate coeffi- cients are based on an inaccurate interaction surface, so the conclusions of Thi et al. [9] may be compromised. Furthermore, Thi et al. [9] extrapolated the rate coefficients to higher temperatures. Our rate coefficients extend the range in temperature from previous studies up to 3000 K; therefore extrap- olation in this case would not be necessary. It is expected that our reliable and comprehensive rate coefficients would lead to more accurate astrophys- ical models of PDRs, PPDs, and other molecular environments. In the protoplanetary disk (PPD) models of Thi et al. [9], they note that pure rotational transitions of CO probe the entire disk, while the warm in- ner regions are probed by rovibrational transitions. Further, the abundance of atomic hydrogen is high in the CO line-emitting region. In particular, it is typically greater than 106 cm-3 throughout the disk, except near the mid-plane and at the disk surface. Assuming that H collisions and spon- taneous emission, with transition probability Aj j , dominate the CO ro- ! 0

Preliminary version – January 29, 2016 46 h-co rate coefficients for pure rotational transitions

tational populations, the critical density for each rotational level j can be estimated following Osterbrock & Ferland [114] with the relation

Aj j ! 0 cr j0

3.5 conclusion

The 3D H-CO potential energy surface of Song et al. [106] was used to perform quantum scattering calculations for rotational deexcitation transi- tions of CO induced by H. State-to-state cross sections for collision energies from 10-5 to 15, 000 cm-1 and rate coefficients for temperatures ranging from 1 to 3000 K were computed for CO pure rotational deexcitation. Not surprisingly, previous scattering results using a PES based on the CCSD(T) level of theory produced similar rate coefficients, but considered only low rotational excitation states. The MRCI surface of Shepler et al. [19] with a less sophisticated level of theory also produced similar results as presented in Yang et al. [15], while the scattering results of Balakrishnan et al. [16] are deemed to be unreliable due to an inaccurate surface. The H-CO pure rotational rate coefficients presented here can be used to aid astrophysical modeling and they both extend the temperature range and angular momen- tum quantum number of reported H-CO rate coefficients. Further, while calculations of deexcitation from excited rovibrational levels are presented in Chap. 4 and 5, experimental data on low-temperature rotational and vi- brational inelastic rate coefficients would be highly desirable.

Preliminary version – January 29, 2016 H-CO SCATTERING CALCULATION FOR RO-VIBRATIONAL4 TRANSITIONS

This chapter focuses on quantum-mechanical scattering calculations for ro- vibrational relaxation of CO in collision with hydrogen atoms. Collisional cross sections of CO ro-vibrational transitions from v = 1, j = 0 - 30 to v0 = 0, j0 are calculated using the close coupling method for collision en- ergies between 0.1 and 15000 cm-1 based on the three-dimensional poten- tial energy surface described in Chap. 2. Cross sections of transitions from v = 1, j > 3 to v0 = 0, j0 are reported for the first time at this level of the- ory. Also calculations by the more approximate coupled states and infinite order sudden methods are performed in order to test the applicability of these methods to H-CO ro-vibrational inelastic scattering. Vibrational de- excitation rate coefficients of CO (v = 1) are presented for the temperature range from 100 K to 3000 K and are compared with the available experimen- tal and theoretical data. All of these results and additional rate coefficients reported in Chap. 5 are important for including the effects of H-CO colli- sions in astrophysical models.

Lei Song, N. Balakrishnan, Ad van der Avoird, Tijs Karman, and Gerrit C. Groenenboom, Quantum scattering calculations for ro-vibrational de-excitation of CO by Hydrogen atoms, J. Chem. Phys., 142, 204303 (2015)

Preliminary version – January 29, 2016 48 h-co scattering calculation for ro-vibrational transitions

4.1 introduction

Carbon monoxide (CO) is the second most abundant molecule in the uni- verse. Its v = 1 ro-vibrational bands around 4.7µm are commonly ob- served in protoplanetary disks using the CRyogenic high-resolution In- fraRed Echelle Spectrograph (CRIRES) on the Very Large Telescope (VLT). [115, 116] These CO bands now rank among the best tracers of the physical conditions in these disks, thus contributing to the understanding of the gas kinematics. Most analyses of CO ro-vibrational bands assume LTE. How- ever, more accurate studies of CO emission lines, such as those concerning the efficiency of IR/VU fluorescence, require non-LTE methods. [9] Accurate non-LTE modeling requires reliable collision rate coefficients of CO with its dominant collision partners H, He, and H2. These rate coefficients can be obtained from scattering calculations based on the interaction potentials of CO with its partners. Besides the accurate CC method, more approximate quantum mechanical approaches such as the CS [108, 117] and IOS [118] methods are commonly used for such scattering calculations. The H-CO collision dynamics has been studied in several experiments [119, 120, 121] and calculations. [17, 18, 19, 48, 49, 99, 122] Until now, experi- mental collision rate coefficients are available only for v = 1 0 transitions, ! without resolution of the transitions between individual rotational levels. The majority of the calculations focus on pure rotational or pure vibra- tional inelastic collision processes. Progress in ro-vibrational cross section calculations has been slow. The difficulty of reliable H-CO ro-vibrationally inelastic scattering calculations is related to the fact that HCO is a chemi- cally bound species, in contrast with other astrophysically important colli- sion complexes such as He-CO and H2-CO that are bound only by weak van der Waals forces. This is illustrated by the depth of the well in the 3D H-CO PES of Song et al. [106], which is 0.835 eV (6738 cm-1), whereas He- -1 -1 CO and H2-CO are bound only by about 23 cm and 94 cm , respectively. [123, 124, 125, 126] Another essential difference with He-CO and H2-CO important for inelastic collision processes is that the H-CO potential has a barrier for dissociation of HCO into H + CO with a height of 0.141 eV -1 (1138 cm ) above the H + CO limit, while He-CO and H2-CO have no barriers. In principle, the strong interaction in H-CO gives rise to strong vibrational coupling and highly efficient translation-vibration (T-V) energy transfer, but in order to get into the region of the deep well, the H atom ap- proaching the CO molecule must first pass over or tunnel through the bar- rier. As one will see below, this leads to a steep rise of the ro-vibrationally inelastic cross sections when the collision energy surpasses the threshold of 1138 cm-1. When the incoming CO molecule is in a rotationally excited state, it may use some of its rotational energy in crossing the barrier. This involves rotation-translation (R-T) coupling, which depends sensitively on the anisotropy of the PES in the long range region. Also pure rotationally inelastic cross sections are mostly determined by the long range anisotropy, while the ro-vibrationally inelastic transition cross sections are sensitive to

Preliminary version – January 29, 2016 4.1 introduction 49 the long and short range anisotropy of the H-CO potential, as well as to its dependence on the CO bond length. A consequence of these complexities is that in H-CO scattering calcula- tions, a large basis set is required to converge the cross sections. Green et al. [48] verified this point in their scattering calculations for ro-vibrational excitation of CO (v = 0, j = 0) by H atoms. The convergence of the cross sections was tested in a series of CS calculations at a total energy of 1 eV (8065 cm-1) with basis sets of increasing size. For ro-vibrational transitions to v0 = 1 and j0 values up to 30 as considered in the present chapter, they found that convergence of the cross sections to within about 10% requires basis sets containing vibration functions up to v = 8 and rotation functions up to j = 90. Even for qualitative results, converged within a factor of 2, basis sets with two or more vibrational levels above the level of interest should be included. In the full quantum CC approach, the CPU time for the rate-limiting step scales roughly with the third power of the basis size. Therefore, the required large basis sets result in heavy computational ef- forts. An additional complication was shown by Stepler et al. [19] in their CC rigid-rotor calculations for CO (v = 0, j = 0) excitation by H atoms based on four different PESs. Their studies revealed that the long-range behavior of the PES strongly affects the inelastic collision cross sections. So it is neces- sary in the scattering calculations to propagate to a relatively long distance to obtain reliable inelastic cross sections, which increases the computational effort as well. The incomplete knowledge about ro-vibrational rate coefficients for H- CO collisions obliges astronomers to use scaling laws to fill the gaps. In a recent paper, Thi et al. [9] modeled CO ro-vibrational emission from Her- big Ae discs by using extrapolated rate coefficients for H-CO based on the values reported by Balakrishnan et al. [16]. In the work of Balakrishnan et al., pure rotational rates were calculated for CO rotational levels up to j = 7 for a range of gas temperatures of 5 K 6 T 6 3000 K, while pure vibrational transition rates were computed in the IOS approximation up to v = 4 for a range of gas temperatures of 100 K 6 T 6 3000 K. The missing rotational rates for j values higher than 7 and vibrational rates below 100 K result in large uncertainties in the extrapolation with the scaling laws. Based on the information from this chapter, we report in Chap. 5 rate coefficients for ro- vibrational v = 1, j = 0 - 30 v = 0, j and v = 2, j = 0 - 30 v = 1, j ! 0 0 ! 0 0 transitions and vibrational rates up to v = 5 for a range of gas temperatures 10 K 6 T 6 3000 K. Furthermore, we introduce a new, more reliable method to extrapolate all ro-vibrational transition rates for CO in initial states with v 6 5, j 6 30. The accuracy of calculated inelastic cross sections depends strongly on the quality of the PES used. The 3D BBH [34] and WKS [18] potentials were used in scattering calculations in previous work. The BBH potential, how- ever, is insufficiently accurate to reproduce stimulated emission pumping (SEP) spectra. [18, 37] This was the reason why Keller et al. [18] constructed the WKS potential, but they did not include a sufficiently large number of

Preliminary version – January 29, 2016 50 h-co scattering calculation for ro-vibrational transitions

ab initio data points at large H-CO distances to make their potential ac- curate also in the long range. It was demonstrated by Shepler et al. [19] that this seriously affects the accuracy of rotationally inelastic cross sec- tions. The 3D PES of Song et al. [106] was constructed with a focus on the long range behavior. In bound state calculations of HCO, it reproduces multiple experimental data, which show that it is accurate also in the well region. In the present work, we adopt this (SAG) potential [106] to calcu- late state-to-state ro-vibrational de-excitation cross sections for initial states with v = 1, j = 0 - 30 to final states v0 = 0, j0. In Sec. 4.2 we briefly describe the theoretical methods and computational details. The different basis sets and approximation methods are discussed in Sec. 4.3. In the same section we present the calculated ro-vibrational (de-)excitation cross sections and vibrational quenching rate coefficients. Our conclusions follow in Sec. 4.4.

4.2 calculations

4.2.1 Theoretical methods

2 The Hamiltonian in BF coordinates for H-CO(X A0) can be written (in atomic units) as e 1 @2 Jˆ2 + ˆj2 - 2jˆ Jˆ Hˆ =- R + · + V (r, R, ✓)+Hˆ ,(4.1) 2µR @R2 2µR2 int CO

where µ = mHmCO/(mH + mCO) is the reduced mass of the H-CO complex. The vector R with length R, connecting the CO center of mass with the H atom, is embedded along the z axis. The vector r, with length r, points from C to O. The angle ✓ is the angle between the vectors R and r. The operator jˆ is the rotational angular momentum of the CO molecule, while Jˆ is the total angular momentum operator of the H-CO complex. The interaction energy between CO and H is written as Vint(r, R, ✓). The CO Hamiltonian Hˆ CO is given by 2 ˆ2 ˆ 1 @ j HCO =- 2 r + 2 + VCO(r),(4.2) 2µCOr @r 2µCOr

where µCO is the reduced mass of CO and VCO(r) is the potential of free CO. To solve the time-independent Schrödinger equation with the Hamilto- nian of Eq. 4.1, the wave functions are expanded as

1 = |n (R),(4.3) R i n n X

where n(R) are functions of the radial coordinate, while the channel basis functions |n are given by i 1/2 2J + 1 1 (J) |n = |v, j, J, K, M = (r)Y (✓, )D (↵, , 0)⇤.(4.4) i i 4⇡ r v,j j,K M,K 

Preliminary version – January 29, 2016 4.2 calculations 51

They consist of the j-dependent vibrational wave functions v,j(r) of free CO and a BF angular basis set. The spherical harmonics Yj,K(✓, ) and Wigner (J) D-functions DM,K(↵, , 0)⇤ form the angular basis set, in which the angles (✓, ) are the polar angles of the CO axis with respect to the BF frame and the Euler angles (↵, , 0) define the orientation of the BF frame with respect to a SF frame. The quantum numbers v and j label the vibrations and rotations of CO. Other quantum numbers are the total angular momentum J of the H-CO complex and its projections K and M on the BF and SF z-axes, respectively. Angular basis functions with parity p =(-1)J+m symmetrized under inversion are given by

1 (J) m (J) Yj,KD ⇤ +(-1) Yj,-KD ⇤ with m = 0, 1 for K>0 p2 M,K M,-K h (J)i Yj,0DM,⇤0 with m = 0 for K = 0. (4.5)

Substituting the expansion of Eq. 4.3 into the time-independent Schrödinger equation and projecting with n’| gives the coupled channel equations h @2 (R)= n’| Wˆ |n (R).(4.6) @R2 n’ h i n n X The operator Wˆ is introduced as

ˆ2 ˆ2 ˆ ˆ ˆ ˆ J + j - 2j J W =-2µ E - HCO - 2 · - Vint(r, R, ✓) ,(4.7) " 2µR # where E is the total energy. The eigenvalues of the Hamiltonian Hˆ CO are the asymptotic channel energies written as ✏v,j. The difference (E - ✏v,j) is the collision energy Ev,j. Solving the coupled channel equations with the appropriate boundary conditions produces the scattering S matrix, which includes all the information needed to deduce the state-to-state scattering cross sections v,j v ,j (Ev,j). Computational details are given in Sec. 4.2.2. ! 0 0 This approach, with the channel basis chosen sufficiently large to converge the solutions of the coupled channel equations, is the accurate CC scattering method used for most calculations in this chapter. If the off-diagonal part of the Coriolis coupling is neglected, i.e., if the operator jˆ Jˆ is replaced by · jˆzJˆz so that K, the eigenvalue of both jˆz and Jˆz, becomes a good quantum number, we obtain the CS approximation. When the exact diagonal Coriolis coupling term is retained, this approach is more accurate than the com- monly applied CS approximation [108, 117] in which the centrifugal kinetic energy term (Jˆ2 + ˆj2 - 2jˆ Jˆ)/2µR2 is approximated by a term L(L + 1)/2µR2 · with some effective value of the quantum number L. When using the CS ap- proximation, the matrix n’| Wˆ |n becomes block-diagonal, with a block for h i each K, and the coupled channel equations of Eq. 4.6 can be separated into subsets for each K to reduce the computational effort. Another method ap- plied in this chapter is the IOS approximation. In this method, which ignores

Preliminary version – January 29, 2016 52 h-co scattering calculation for ro-vibrational transitions

the rotational energy spacings in comparison with the collision energy [118] and neglects the CO rotational kinetic energy term in the Hamiltonian, it is assumed that the angle ✓ of the CO molecule remains fixed during the collision. The IOS calculations were performed with the molecular scatter- ing program MOLSCAT [109] to obtain the total vibrational v = 1 v = 0 ! 0 quenching cross section. Coupled channels equations in a pure vibrational basis set are solved in the IOS approximation for a fixed angle ✓ and the results are averaged over 50 Gauss-Legendre quadrature points ✓i with the proper weights. The basis set includes the lowest 15 vibrational lev- els, which is sufficient to converge the v = 1 v = 0 vibrational transition ! 0 cross sections. Rate coefficients for ro-vibrational energy transfer can be obtained by integrating the corresponding state-to-state cross sections over a Boltzmann distribution of collision energies at a certain temperature T,

1 8kBT 2 1 Ev,j rv,j v ,j (T)= v,j v ,j (Ev,j) exp(- ) Ev,j dEv,j, ! 0 0 ⇡µ 2 ! 0 0 k T ✓ ◆ (kBT) Z01 B (4.8) where kB is the Boltzmann constant. The state-to-state cross sections v,j v ,j ! 0 0 (Ev,j) can be obtained with the CC method or the CS approximation. The to- tal vibrational quenching cross section from a specific ro-vibrational initial state v, j to final state v0 is the sum of cross sections over all final j0 states of the v0 state, i.e., v,j v (Ev,j)= v,j v ,j (Ev,j).(4.9) ! 0 ! 0 0 Xj0 The corresponding rate coefficients are written as rv,j v (T). Then, by ther- ! 0 mally averaging the rate coefficients over different initial j states, we get the vibrational de-excitation rate coefficients based on the CC method or the CS approximation as follows

✏v,j j gjexp(- k T )rv,j v (T) B ! 0 rv v (T)= ✏ ,(4.10) ! 0 g exp(- v,j ) P j j kBT P where gj = 2j + 1 is the degeneracy of the j rotational level.

4.2.2 Computational details

Having described the theoretical formalism, we turn to the numerical de- tails in the calculations of the integral cross sections. The CC and CS calcula- tions were performed with our own scattering codes written in Fortran and Scilab [127], which include some features developed to save computational resources. As usual in scattering calculations, we expand the anisotropic interaction potential Vint(r, R, ✓) in Legendre polynomials Pl(cos ✓) and cal- culate its matrix elements over the angular basis of Eq. 4.4 analytically in terms of 3j-symbols.[118] We prepare the angular integrals and R-dependent angular expansion coefficients of the anisotropic potential in advance and

Preliminary version – January 29, 2016 4.2 calculations 53 save them on disk. The coupled channel equations are solved with the renor- malized Numerov propagator [12, 13]. During the propagation, instead of keeping the W-matrices in core, our program recomputes the W-matrix from the prepared angular integrals and potential expansion coefficients for every R point, which considerably saves memory. The rate-limiting steps are matrix multiplications and inversions in the renormalized Nu- merov algorithm; these are run on several cores in parallel with optimized MKL LAPACK library routines. Also the fact that the expression for the W-matrix elements of the potential becomes simpler in the BF angular basis than in the usual SF basis saves both computer time and storage. Moreover, when applying the CS approximation, we divide the problem into smaller blocks for each K quantum number. At the end of the propagation, the ra- dial solutions (R) associated with the BF angular basis |n = |v, j, J, K, M n i i are transformed into the corresponding ones for the SF basis |v, j, L, J, M i with the end-over-end rotational quantum number L, in order to apply the scattering boundary conditions formulated in the SF frame. The exact solu- tion of the SF scattering Schrödinger equation at the outer boundary where the interaction energy Vint(r, R, ✓) has vanished, is a linear combination of spherical Bessel functions labeled with the quantum number L. The match- ing of the propagated radial wave functions to spherical Bessel functions was carried out at the boundary Rmax with values of 40 to 18 a0 adapted to -1 a range of collision energies from Ev,j = 0.1 to 15000 cm . The propagation step size varied from 0.032 to 0.055 a0. The parallellization and memory saving features of the code reduce the requirement on the computational resources dramatically, which made it possible to use hundreds of simple personal computers connected over a regular ethernet network for a large number of small jobs. As usual in scattering calculations, we exploit the property that J, M, and the parity p are good quantum numbers and we collect the correspond- ing contributions to the integral cross sections from separate calculations. The maximum value of J needed to converge the cross sections depends on the collision energy and the initial j value; values of J up to 100 were used. Using the CC method, we calculated the state-to-state de-excitation cross sections from v = 1, j = 0 - 30 to v0 = 0, j0. Initially, we adopted two basis sets B3(61, 52, 40, 22) and B9(79, 70, 59, 45, 45, 30, 30, 25, 25, 25) for the calculations with initial states v = 1, j = 0 - 10, where the notation Bn(j0, j1, j2, ..., jn) indicates a basis with the highest vibrational level n and the highest rotational level ji included for vibrational level v = i. We no- tice in Table 4.1 that the number of channels in the B9 and B3 basis sets increases dramatically with increasing J. The number of channels for B3 is less than half of that for the B9 basis. The third power of the ratio of the numbers of channels for B9 and B3 is around 15 to 18, which implies that the calculation with the B3 basis is 15 - 18 times faster than the one with the B9 basis. To further reduce the computational effort, we adapted the renormalized Numerov algorithm so that we could gradually omit closed channels during the propagation.[128] This saved about 30% of CPU time

Preliminary version – January 29, 2016 54 h-co scattering calculation for ro-vibrational transitions

and did not deteriorate the accuracy of the results by more than 0.001%. This truncation is based on the idea that the coupling with higher energy levels diminishes when the potential becomes smaller for larger R. The cri- terion used to truncate the channel basis considers the channel energies and is an exponential function of R dependent on the decay of the potential with increasing R. The accuracy of the CS approximation is tested by comparison with full CC calculations performed for the initial states v = 1, j = 0 - 10 with basis B3(61, 52, 40, 22). The propagation was done separately for the different K blocks of the W-matrix and the propagated results for all K with -J 6 K 6 J were combined and transformed to the SF basis when applying the L-dependent outer boundary conditions.

4.3 results and discussion

4.3.1 Convergence test

By comparison with the results from an even larger basis set B9⇤(84, 75, 64, 50, 50, 35, 35, 30, 30, 30) we could show that the B9(79, 70, 59, 45, 45, 30, 30, 25, 25, 25) basis yields cross sections of the transitions v = 1, j = 0 - 10 -!1 v0 = 0, j0 converged to within 10% at collision energies Ev,j 6 5000 cm , which is in agreement with the results of convergence tests by Green et al. [48]. An accuracy of 10% is sufficient for the applications in astronomy that we are aiming at. Even our B9 basis is still large, however, and produces up to 11, 000 scattering channels. In order to reduce our computational ef- forts, we tested a smaller B3(61, 52, 40, 22) basis set for initial states with v = 1, j = 0 - 10. Figure 4.1 shows that the state-to-state cross sections from the B9 and B3 bases are similar, except that some fine resonance struc- tures are lost with the B3 basis set which may affect the rate coefficients. The root mean square relative difference of the rate coefficients from the two bases is less than 25%, which makes the values from the B3 basis suffi- ciently accurate for applications in astronomy. In addition, since resonances become less important with increasing initial rotational quantum number j —which will be illustrated in Sec. 4.3.3— use of the smaller basis is ex- pected to have a much smaller influence on the rate coefficients for the transitions from initial v = 1 states with j>10. Our final choice was to em- ploy a B4(70, 60, 49, 36, 15) basis for the transitions from v = 1, j = 11 - 30 and the B9(79, 70, 59, 45, 45, 30, 30, 25, 25, 25) basis for transitions from states with lower initial j values. Table 4.1 shows that the use of CPU time with the B4(70, 60, 49, 36, 15) basis is 6 or 7 times less than with the B9 basis.

4.3.2 Tests of the CS and IOS approximations

One of the options to calculate cross sections more efficiently is the CS ap- proximation. To test its accuracy, we performed scattering calculations for initial states v = 1, j = 0 - 10 with the B3(61, 52, 40, 22) basis. Figure 4.2 shows a comparison of total quenching cross sections from v = 1, j = 0, 5, 10

Preliminary version – January 29, 2016 4.3 results and discussion 55

−1 10

−2 (a) v=1, j=0 → v’=0, j’=5 10 ) 2

cm −3 10 − 16

−4 10

−5 10

−6 Cross section (10 10 B9 B3 −7 10

−3 (b) v=1, j=5 → v’=0, j’=0 10 ) 2

cm −4 10 − 16

−5 10

−6 10

Cross section (10 −7 10 B9 B3 −8 10

−2 → 10 (c) v=1, j=10 v’=0, j’=5 ) 2

cm −3 10 − 16

−4 10

−5 10

Cross section (10 −6 10 B9 B3 −7 10 −1 0 1 2 3 4 10 10 10 10 10 10 Collision energy (cm−1)

Figure 4.1: Comparison of state-to-state cross sections for H+CO calculated with the B9(79, 70, 59, 45, 45, 30, 30, 25, 25, 25) and B3(61, 52, 40, 22) basis sets.

Preliminary version – January 29, 2016 56 h-co scattering calculation for ro-vibrational transitions

0 10 (a) v=1, j=0 → v’=0 )

2 −1 10 cm

− 16 −2 10

−3 10

−4 10 CC−B3

Cross section (10 −5 10 CS−B3 IOS

−6 10 0 10 (b) v=1, j=5 → v’=0

) −1 2 10 cm −2

− 16 10

−3 10

−4 10

−5 10 Cross section (10 −6 CC−B3 10 CS−B3

−7 10

−1 → 10 (c) v=1, j=10 v’=0 ) 2

cm −3 10 − 16

−5 10

−7 10

Cross section (10 −9 10 CC−B3 CS−B3 −11 10 −1 0 1 2 3 4 10 10 10 10 10 10 Collision energy (cm−1)

Figure 4.2: Comparison of total quenching cross sections for H+CO based on the accurate close coupling method, the coupled states approximation, and the infinite order sudden approximation. The line labeled "IOS" shows the cross sections of the vibrational v = 1 v = 0 transition from the ! 0 infinite order sudden approximation.

Preliminary version – January 29, 2016 4.3 results and discussion 57

Table 4.1: Number of channels with the B3, B4, and B9 basis sets for total J = 0 (parity p = 1) and J = 5, 10, 20, 50, 80 (parity p =-1).

B9 3 B9 3 JB3B4B9B3 B4 0179235443157

5101413352508157

10 1570 2075 3880 15 7

20 2740 3660 6760 15 6

50 4273 6061 10700 16 6

80 4342 6326 11390 18 6

to v0 = 0 based on the CC method and the CS approximation. The total quenching cross sections are obtained by summation of the state-to-state cross sections over all final rotational states, see Eq. 4.9. In comparison with the accurate CC calculations, the CS approximation produces reliable cross sections for collision energies above 1000 cm-1. The deviation of the CS cross section from the CC results at lower energies becomes larger for higher quantum numbers j. The CS approximation is obtained from the CC method by neglecting the off-diagonal Coriolis coupling terms in the kinetic energy operator. The terms neglected are more or less proportional to j(j + 1), which explains why the CS approximation becomes worse with increasing p j. So, we must conclude that the CS approximation is not suitable for the calculation of the H-CO ro-vibrational de-excitation rates at low tempera- tures. The CS approximation has the tendency to perform well for processes dominated by short-range repulsive interactions; in the present study of H- CO, we do not find ourselves in that situation, however. In the top panel of Fig. 4.2, we also present the cross section for vibrational v = 1 v = 0 ! 0 transitions based on the IOS approximation. The IOS results for this tran- sition agrees fairly well with the CC values for v = 1, j = 0 v = 0 ! 0 quenching. They also agree well with the total vibrational v = 1 v = 0 ! 0 quenching cross section from CC calculations obtained by averaging the v = 1, j v = 0 cross sections over initial j levels, see Eq. 4.10. The latter ! 0 are not explicitly shown in Fig. 4.2, but one can observe in this figure that the v = 1, j = 0 v = 0 cross sections hardly depend on the initial j values ! 0 —apart from resonance structures, which are most pronounced for j = 0—, so that the averaging over j has only a minor effect.

Preliminary version – January 29, 2016 58 h-co scattering calculation for ro-vibrational transitions

−1 −1 ) 10 ) 10 2 2 (a) v = 1, j = 0 → v’ = 0, j’ (b) v = 1, j = 5 → v’ = 0, j’ cm cm 10−3 −3 − 16 − 16 10

−5 10 −5 j’ = 0 10 j’ = 0 j’ = 5 j’ = 5 −7 j’ = 10 j’ = 10 10 −7 j’ = 15 10 j’ = 15 j’ = 20 j’ = 20 −9 Cross section (10 10 Cross section (10 10−1 100 101 102 103 104 10−1 100 101 102 103 104 Collision energy (cm−1) Collision energy (cm−1)

−1 ) ) 10 2 2 −2 (c) v = 1, j = 15 → v’ = 0, j’ (d) v = 1, j = 25 → v’ = 0, j’

cm 10 cm 10−3 − 16 − 16 −4 10 −5 j’ = 0 10 j’ = 0 j’ = 5 j’ = 5 −6 10 j’ = 10 10−7 j’ = 10 j’ = 15 j’ = 15 j’ = 20 j’ = 20 −8 −9 Cross section (10 10 Cross section (10 10 10−1 100 101 102 103 104 10−1 100 101 102 103 104 Collision energy (cm−1) Collision energy (cm−1)

Figure 4.3: State-to-state cross sections of CO in collision with H atoms from ini- tial states v = 1, j = 0, 5, 15, 25 to individual final states v0 = 0, j0 = 0, 5, 10, 15, 20. For initial states v = 1, j = 0, 5 the values are calculated with the B9(79, 70, 59, 45, 45, 30, 30, 25, 25, 25) basis.

4.3.3 Cross sections

Examples of energy dependent state-to-state integral cross sections of the transitions from initial states v = 1, j = 0, 5, 15, 25 to individual final states v0 = 0, j0 = 0, 5, 10, 15, 20 are shown in Fig. 4.3. One observes in this figure that the cross sections from the same initial v, j states have a similar energy dependence. Figure 4.3(a) shows that the cross sections from initial state v = 1, j = 0 have abundant structures, due to resonances. Panels (b)–(d) of this figure illustrate that these structures vanish with increasing initial j. All of these state-to-state cross sections are relatively small for low col- lision energy, stay more or less constant with increasing energy, and then increase by three orders of magnitude at energies in the range from 100 to 1000 cm-1, followed by a flatter behavior in the high energy region. The increase of the cross sections is steepest for the initial CO state with j = 0, where it starts at collision energies around 1000 cm-1. The explana- tion of this dramatic increase is the existence of a barrier with a height of 1138 cm-1 in the entrance channel of the H-CO potential surface. Vibration- translation (V-T) and vibration-rotation (V-R) energy transfer mainly takes place at short H-CO distances in the chemical bonding and repulsive re- gion of the H-CO potential, inside of the barrier, where the potential de- pends most sensitively on the CO bond length. For higher initial rotational

Preliminary version – January 29, 2016 4.3 results and discussion 59

0 10

−1

) 10 2 cm

− 16 −2 10

−3 10

Cross section (10 −4 j = 1 10 j = 8 j = 15

−5 j = 20 10 j = 30

−1 0 1 2 3 4 10 10 10 10 10 10 Collision energy (cm−1)

Figure 4.4: Total quenching cross sections of CO in collision with H atoms from v = 1, j = 1, 8, 15, 20, 30 to v0 = 0. For initial states v = 1, j = 1, 8, the values are calculated with the B9(79, 70, 59, 45, 45, 30, 30, 25, 25, 25) basis. quantum number j of CO, the increase of the cross sections becomes more gradual and starts at lower collision energy. This can be explained by a release of the initial rotational energy of CO through rotation-translation (R-T) energy transfer when the system approaches the barrier. The addi- tional translational energy helps the system to overcome the barrier, so that this can happen at lower collision energies. R-T energy transfer is caused by anisotropic H-CO interactions, which makes it important that the long range anisotropy of the potential is accurate. This interplay of short and long range interactions in H-CO leads to a much more complex behavior than in a weakly bound system as He-CO, where ro-vibrational energy transfer processes [129] are all determined by van der Waals interactions. Due to this dramatic increase of the cross sections at collision energies of about 1000 cm-1, their high-energy contribution to the rate coefficients is much more important than usual. Therefore, cross sections at relatively high energies needed to be computed and included in the calculation of the H-CO rate coefficients by Boltzmann averaging over collision energies, even for lower temperatures. This is another reason why the calculation of rate coefficients for H-CO collisions is so expensive. In Fig. 4.4 we present the total quenching cross sections from initial states v = 1, j = 1, 8, 15, 20, 30 to final v0 = 0. The general behavior of these cross sections is similar to that of the state-to-state cross sections, but the struc- tures in the cross sections are washed out to some extent in the sum of state-to-state cross sections over all final j0 values. Apart from resonances, the cross sections from higher initial j states are always larger at collision energies below 1000 cm-1. For higher energies, the total quenching cross sections for different initial j values become almost equal and gently rise together.

Preliminary version – January 29, 2016 60 h-co scattering calculation for ro-vibrational transitions

−4 x 10

(a) E = 0.1 cm−1 vj 1.5 ) 2 cm − 16 1

0.5 Cross section (10 j = 0 (x10) j = 10(x10) j = 20 j = 30 0

0.05 (b) E = 5000 cm−1 vj

) 0.04 2 cm − 16 0.03

0.02 j = 0

Cross section (10 j = 10 0.01 j = 20 j = 30

0 0 10 20 30 40 50 60 70 J

Figure 4.5: Contributions to the total quenching cross sections for H+CO from different total angular momenta J for the transitions from v = 1, j = -1 0, 10, 20, 30 to v0 = 0 at collision energies (a) Ev,j = 0.1cm and (b) -1 -1 Ev,j = 5000 cm (5210cm for j = 0).

The contributions from individual partial waves to the total quenching cross sections are shown in Fig. 4.5 for four different initial states with j = 0, 10, 20, and 30. In the top panel, for a collision energy of 0.1 cm-1, the contributions are all sharply peaked with maxima appearing at J = j. This reveals the dominant contribution of s-wave (L = 0) scattering at this low energy. For a collision energy of 5000 cm-1, the contributions from different total J values are much more broadly distributed, as illustrated in the bottom panel of the figure. For the higher initial j values of 20 and 30, the maxima still correspond to J = j - 1 or j , while for lower initial j values of 0 and 10, they appear at J values higher than j. The higher the collision energy, the larger the number of partial waves needed to obtain converged cross sections. Moreover, the contributions from different partial waves shift to higher J for higher initial j. For both reasons the number and size of the computations increase dramatically for higher collision energy.

Preliminary version – January 29, 2016 4.3 results and discussion 61

−5 x 10

(a) E = 0.1 cm−1 1 vj ) 2 cm 0.75 − 16

0.5

Cross section (10 j = 0 0.25 j = 10 j = 20 j = 30(x0.5) 0

0.05 (b) E = 5000 cm−1 vj

0.04 ) 2 cm − 16 0.03

0.02 j = 0 j = 10

Cross section (10 j = 20 0.01 j = 30

0 0 10 20 30 40 50 60 70 j’

Figure 4.6: Distributions of final rotational levels j0 in the v0 = 0 vibrational level from de-excitation of initial states v = 1, j = 0, 10, 20, 30 at collision -1 -1 -1 energies (a) Ev,j = 0.1cm and (b) Ev,j = 5000 cm (5210cm for j = 0).

The distributions of the final rotational levels j0 in ro-vibrationally inelas- tic v = 1, j v = 0, j collisions with collision energies E = 0.1 cm-1 and ! 0 0 v,j 5000 cm-1 are presented in Fig. 4.6. At the low collision energy, the final rotational level distributions are dominated by j = 0 transitions for initial j values of 0 and 10. For the higher initial j values, they become broader. At the higher collision energy, the distributions of the final rotational lev- els j0 are broader for all initial j values, but they still show maxima for j = 0 transitions for the higher initial j values of 20 and 30. At both col- lision energies, one observes oscillations in the j0 distributions which are due to the competition of rotationally inelastic processes induced by the terms with even and odd l in the Legendre expansion of the anisotropic H-CO interaction potential. This explanation is basically the same as the explanation given for the propensity of even/odd j transitions in the v = 0 ground state, [106] but the nature of the oscillations becomes less simple when vibrational inelasticity is involved.

Preliminary version – January 29, 2016 62 h-co scattering calculation for ro-vibrational transitions

10−9

10−10 ) − 1 s 3 10−11

10−12

CC−SAG potential 10−13 IOS−SAG potential Rate coefficient (cm IOS−WKS Balakrishnan 10−14 Exp. Glass & Kironde Exp. Rosenberg−Kozlov Astro. Neufeld & Hollenbach 10−15 Astro. Hollenbach & McKee

500 1000 1500 2000 2500 3000 T(K)

Figure 4.7: Comparison of vibrational v = 1 v = 0 de-excitation rate coeffi- ! 0 cients of CO by collision with hydrogen atoms from different calcula- tions and measurements.

4.3.4 Vibrational quenching rate coefficients

The vibrational v = 1 v = 0 quenching rate coefficients of CO in colli- ! 0 sions with H atoms can be calculated directly from the IOS approximation, but can also be obtained from state-to-state cross sections calculated with the CC method through the use of Equs. 4.8–4.10. In Fig. 4.7, we plot these rates in the temperature range from 100 K to 3000 K. The CC results are plot- ted as a green line with cross markers, the IOS values calculated from the same (SAG) H-CO potential of Song et al. are shown in the figure as a dark red dashed line. The difference between the CC and IOS rate coefficients is small, which implies that the IOS method is a good approximation for the calculation of pure vibrational de-excitation rate coefficients. The results of IOS calculations of Balakrishnan et al. [16] on the semi-empirically adjusted WKS potential are slightly larger than our values from the SAG potential at temperatures below 1500 K. This is a consequence of the dissociation bar- rier of HCO to H + CO on the modified WKS potential being somewhat lower than on the SAG potential. In Fig. 4.7 we also plot experimental re- sults measured by Glass et al. [120] and by von Rosenberg et al. [119]. The solid lines with square and circular markers are fits to the measured data, while the dashed lines are extrapolations of these fits to temperatures not accessed by the measurements. Other dashed lines in Fig. 4.7 illustrate some empirical formulas used in astrophysical models. [130, 131] The differences between the experimental data from Refs. [119] and [120] are large. von Rosenberg et al. [119] only speculated on the occurrence of H-CO collisions affecting their measurements for other species, and they estimated the ef-

Preliminary version – January 29, 2016 4.4 conclusions 63 fect of collisions with hydrogen atoms on the CO relaxation. Similar results were obtained by Kozlov et al. [121] Glass et al. [120] performed H-CO scat- tering experiments at various temperatures and pressures, so we believe that their data are more reliable. Our CC and IOS vibrational quenching rate coefficients agree well with these data [120], which implies that the 3D SAG potential is suitable for the calculation of vibrational quenching rate coefficients. The IOS calculations on the modified WKS potential reproduce the experimental data as well, which indicates that the vibrational energy is released in the well region where the two potential surfaces are similar and that the vibrational quenching rate coefficients are less sensitive to the long range part of the potential. However, as mentioned in previous publi- cations [19, 15] and also here, the modified WKS potential is less successful in predicting rotationally inelastic scattering cross sections, owing to its less accurate long range part, and this will probably hold also for the state-to- state ro-vibrational cross sections and rate coefficients.

4.4 conclusions

State-to-state ro-vibrational v, j v , j transition cross sections for CO in ! 0 0 collision with H atoms are obtained by means of the full CC method with the 3D (SAG) potential of Song et al. [106]. Ro-vibrational basis sets with maximum v and j values of 9 and 79, respectively, and up to 11, 000 chan- nels are required to converge the ro-vibrational transition cross sections and corresponding rate coefficients to a level appropriate for astronomical applications. Comparison with results obtained from CS calculations shows that the CS approach is less suitable in this case. Pure vibrational v v ! 0 transition cross sections and rate coefficients are obtained from the CC re- sults by summation over final rotational levels j0 and averaging over initial levels j and they were computed also in the IOS approximation. The CC and IOS results are in good agreement, which shows that the IOS method works well for v v transitions. The computed v v rate coefficients also agree ! 0 ! 0 with the measured data of Glass et al. [120]. We find, in agreement with earlier work [19], that pure rotational j j ! 0 transition cross sections are mainly sensitive to the anisotropy of the H-CO potential in the long range. Ro-vibrational transitions require the system to enter the short range region, where the H-CO potential has a deep well -1 (De = 6738 cm ) at a H-CO distance of 3.021 a0 and the coupling to the CO vibration is most effective. A barrier of 1138 cm-1 is present in the H+CO entrance channel, however, which explains why the calculated v, j v , j ! 0 0 cross sections increase by three orders of magnitude when the collision energy surpasses 1000 cm-1. Overcoming the barrier is facilitated through R-T energy transfer when CO is initially in a higher rotational level j;we observe that the rise of the v, j v , j cross sections starts at lower collision ! 0 0 energies and becomes less steep in that case. Both the anisotropy of the H- CO potential and its dependence on the CO bond length are important for

Preliminary version – January 29, 2016 64 h-co scattering calculation for ro-vibrational transitions

these combined processes, and this also explains why such large basis sets are needed to converge the ro-vibrationally inelastic cross sections. In Chap. 5, we present ro-vibrational transition rate coefficients for CO in collision with H atoms for temperatures up to 3000 K. The dramatic rise of the v, j v , j cross sections when the collision energy gets above 1000 ! 0 0 cm-1 made it necessary to sample a relatively large set of high collision energies in the Boltzmann averaging, even for lower temperatures. In com- bination with the large basis sets required, this made the calculations com- putationally very demanding. We described some special features of our close-coupling scattering code that facilitated the computations.

Preliminary version – January 29, 2016 H-CO RATE COEFFICIENTS FOR RO-VIBRATIONAL TRAN-5 SITIONS

In this chapter, we present calculated rate coefficients for ro-vibrational tran- sitions of CO in collisions with H atoms for a gas temperature range of 10 K 6 T 6 3000 K, based on the recent three-dimensional ab initio H-CO interaction potential of Song et al. [106], described in Chap. 2. Rate co- efficients for ro-vibrational v = 1, j = 0 - 30 v = 0, j transitions were ! 0 0 obtained from scattering cross sections previously computed with the close- coupling method by Song et al. [132], described in Chap. 4. Combining these with the rate coefficients for vibrational v = 1 - 5 v

L. Song, N. Balakrishnan, K. M. Walker, P. C. Stancil, W. F. Thi, I. Kamp, A. van der Avoird and G. C. Groenenboom, Quantum calculation of inelastic CO collisions with H. III. Rate coefficients for ro-vibrational transitions, Astro- phys. J., 813, 96 (2015)

Preliminary version – January 29, 2016 66 h-co rate coefficients for ro-vibrational transitions

5.1 introduction

As the second most abundant molecule in the universe, carbon monoxide is commonly detected in a variety of astrophysical environments. As early as the 1970’s, the infrared ro-vibrational bands of CO were observed in late-type stars [133, 134, 135]. Later, Scoville et al. [136] and Ayres [137] de- tected CO ro-vibrational lines in the spectra of young stellar objects and the . In a sample of nine infrared sources identified by Mitchell et al. [138], eight sources show evidence of a hot gas component with temperatures of 120-1010 K and CO ro-vibrational emission. Recently, v = 1 ro-vibrational transitions of CO near 4.7µm have been observed in star-forming regions, protoplanetary disks, and external galaxies using high-resolution spectrom- eters [32, 115, 116, 139, 140, 141, 142, 143, 144, 145, 146]. Analysis of these CO spectral lines provides the physical conditions and chemical composi- tion of the interstellar gas, in particular the kinetic temperature, column density, volume density, and molecular abundances. Many of the analy- ses assumed that the CO ro-vibrational emission is based on LTE popula- tions, which, however, applies only in high-density regions where collisions dominate the excitations. Non-LTE modeling is necessary when IR and UV fluorescent excitation and radiative de-excitation compete with molecular collisions [147], and it is required for a detailed understanding of the CO ro-vibrational lines in terms of spatial location and efficiency of the IR/UV fluorescence [9]. The non-LTE analysis requires accurate collision rate coef- ficients of CO with its main collision partners, H, H2, He, and electrons, as input. The lack, or limited reliability, of such rate coefficient data hin- ders non-LTE modeling of molecular spectra for many astrophysical envi- ronments. In particular, rate coefficients for the H-CO system are known to be highly uncertain [19] and difficult to calculate due to the existence of a chemical bond between H and CO. The appearance of a deep well and a dissociation barrier in the H-CO potential make it substantially more dif- ficult to compute converged scattering cross sections than for typical van der Waals systems such as He-CO and H2-CO. In addition, it was shown in Walker et al. [148] that standard approaches for scaling He-CO rate coeffi- cients to obtain values for H2-CO or, especially, H-CO are not valid. For pure rotational transitions, rate coefficients of H-CO were first given by Chu & Dalgarno [99] about four decades ago for CO excitation from j = 0 - 4 to j0 = 1 - 5 at temperatures of T = 5 - 150 K. One year later, Green & Thaddeus [100] reported their values for transitions from initial states with j = 0 - 7 at temperatures of T = 5 - 100 K. Rate coefficients for pure rotational transitions up to j = 7 for a broad range of gas temperatures T = 5 - 3000 K, were calculated by Balakrishnan et al. [16]. Compared with the earlier values of Green & Thaddeus [100], their results differed by a factor of 30 for rate coefficients at temperatures below 100 K. More recently, Shepler et al. [19] performed scattering calculations with two new ab initio potential energy surfaces and suggested that the pure rotational rate coef- ficients obtained by Balakrishnan et al. [16] were incorrect, due to the fact

Preliminary version – January 29, 2016 5.1 introduction 67 that the WKS interaction potential [18] they used for H-CO was inaccurate in its long range part. In a recent paper, Yang et al. [15] calculated the rota- tional quenching rate coefficients of low-lying (j = 1 - 5) rotational CO lev- els for temperatures ranging from 1 to 3000 K. They found pure rotational quenching rate coefficients similar to the values of Green & Thaddeus [100], which confirms the inaccuracy of results obtained using the WKS potential [16]. Very recently, Walker et al. [149] performed calculations of the pure rotational quenching rates for initial states up to j = 45 at a temperature range of T = 1 - 3000 K based on the new 3D H-CO potential of Song et al. [106]. For vibrational rate coefficients the situation is still unsatisfactory, how- ever. Experimental values are available only for the rotationally unresolved v = 1 0 transition. The most reliable results were measured by Glass & ! Kironde [120] using a discharge-flow shock tube at temperatures between 840 and 2680 K. Other experimental values reported by von Rosenberg et al. [119] and Kozlov et al. [121] are based on estimates of the efficiency of H atoms in vibrational relaxation of CO derived from data involving other gases. Quantum scattering calculations of rate coefficients for transitions between vibrational levels up to v = 4 were performed by Balakrishnan et al. [16] with the IOS approximation at temperatures ranging from 100 to 3000 K. Their computed results agreed with the experimental data mea- sured by Glass & Kironde [120] in the high-temperature range. Vibrational rate coefficients below 100 K were not available, however. Rate coefficients for ro-vibrational transitions were even less complete. Yang et al. [122] pre- sented quenching rate coefficients for initial states v = 1, j = 0, 1, 2 to v = 0 at a temperature range of T = 10-5 - 300 K. They performed CC calculations, but their channel basis may have been too small to ensure convergence and they adopted the WKS potential [18], known to be inaccurate in the long range. Both factors may lead to uncertainties in the ro-vibrational rate co- efficients and the lack of a comprehensive set of H-CO ro-vibrational rate coefficients obliged astronomers to use scaling laws to extrapolate the data (e.g., Ref [9]). Recently, Thi et al. [9] modeled CO ro-vibrational emission from Her- big Ae discs using the H-CO rate coefficients calculated by Balakrishnan et al. [16]. They extrapolated the pure rotational de-excitation rate coefficients to initial states with j>7and vibrational transition rate coefficients to tem- peratures below 100 K. Assuming complete decoupling of rotational and vibrational motion, they estimated the state-to-state ro-vibrational rate co- efficients from the corresponding rates for pure rotational transitions in the ground vibrational level. This method was described in detail and applied to the H2O-H2 system by Faure & Josselin [150]. However, extrapolations based on only a few calculated rate coefficients may cause large errors in the deduced rates. Moreover, the above mentioned inaccuracy of the pure ro- tational rate coefficients derived from the WKS potential [16] will introduce an additional error. In the present chapter, we report explicitly calculated state-to-state ro-vibrational de-excitation rate coefficients for v =-1 tran-

Preliminary version – January 29, 2016 68 h-co rate coefficients for ro-vibrational transitions

sitions in H-CO for initial states with v = 1 and 2 and j values up to 30 for temperatures ranging from 10 to 3000 K. A new extrapolation method is devised and applied to obtain state-to-state ro-vibrational de-excitation rate coefficients for initial states with v up to 5. A Leiden Atomic and Molecular Database (LAMDA)-type file [30] is provided with ro-vibrational de-excitation rate coefficients of H-CO for v = 1 - 5, j = 0 - 30 v , j ! 0 0 transitions at temperatures of T = 10 - 3000 K.

5.2 theory

5.2.1 Equations for rate coefficients

The CC calculations were performed using our scattering code described in Chap. 4, while the IOS calculations were carried out with the non-reactive scattering program MOLSCAT [109]. The scattering methods used to obtain cross sections are reported in detail in Chap. 4. Rate coefficients for specific ro-vibrational transitions were calculated by averaging the corresponding cross sections over a Maxwell-Boltzmann distribution of translational ener- gies of the colliding particles,

1 8k T 2 1 E r(T)= B (E )exp - k E dE ,(5.1) ⇡µ 2 k k T k k ✓ ◆ (kBT) Z01 ✓ B ◆

where µ is the reduced mass of H-CO, kB is the Boltzmann constant and T is the gas temperature. Cross sections were calculated over an energy range from 0.1 to 15000 cm-1, see Chap. 4. The integral over collision energies was computed numerically with the trapezoidal rule after cubic spline interpo- lation of the cross sections on a logarithmic energy scale. The cross section (Ek) as a function of the collision energy Ek can be the state-to-state ro- vibrational cross section v,j v ,j (Ek), the total vibrational quenching cross ! 0 0 section v,j v (Ek) for the transitions from an initial v, j state to a final v0 ! 0 state, or the vibrational transition cross section v v (Ek). The quantum ! 0 numbers v and j refer to the vibration and rotation in the initial state, while v0 and j0 refer to the final state. The rate coefficients for the reverse transi- tions can be obtained by detailed balance

2j + 1 ✏v0,j0 - ✏v,j rv ,j v,j(T)= exp rv,j v ,j (T),(5.2) 0 0! 2j + 1 k T ! 0 0 0 ✓ B ◆

where ✏v,j and ✏v0,j0 are the energies of the ro-vibrational levels. The state-to- CC state ro-vibrational cross sections (Ek) are obtained from full CC cal- v,j v0,j0 culations; the corresponding rate coefficients! can be calculated from Eq. 5.1. By summation of the state-to-state ro-vibrational cross sections over all fi-

Preliminary version – January 29, 2016 5.2 theory 69

nal j0 levels in the v0 state, we obtain the total vibrational quenching cross section from a specific ro-vibrational initial state v, j to final state v0

CC CC v,j v (Ek)= v,j v ,j (Ek).(5.3) ! 0 ! 0 0 Xj0 CC The corresponding rate coefficients are denoted by rv,j v (T). Averaging the ! 0 rate coefficients rCC (T) over a thermal population of initial j states yields v,j v0 the vibrational transition! rate coefficient based on the CC approach

✏v,j CC gjexp(- )r (T) CC j kBT v,j v0 r T ! v v ( )= ✏v,j ,(5.4) ! 0 g exp(- ) P j j kBT P where gj = 2j + 1 is the degeneracy of the rotational level j. IOS The vibrational transition cross sections (Ek) are obtained directly v v0 from scattering calculations in the IOS approximation.! The corresponding rate coefficients can again be calculated from Eq. 5.1.

5.2.2 Extrapolation method

If we wish to calculate all ro-vibrational rate coefficients of interest for as- trophysical modeling, the full CC method is prohibitively expensive. The incompleteness of the available ro-vibrational rate coefficients forced as- tronomers to use extrapolated data, commonly based on a complete de- coupling of vibration and rotation [9, 150, 151]. Here, we introduce a new extrapolation method for state-to-state ro-vibrational rate coefficients in which we assume the coupling in v, j v , j transitions with v

rv,j v ,j (T)=Pvv (T)r1,j 0,j (T),(5.5) ! 0 0 0 ! 0 where the factor Pvv0 (T) is defined as ✏ r (T) g exp(- v,j ) v v0 j j kBT rv v (T) P (T)= ! = ! 0 .(5.6) vv0 ✏v,j r (T) j [gjexp(-Pk T ) j r1,j 0,j (T)] 1 0 B 0 ! 0 ! The difference withP the extrapolationP method used by Thi et al. [9] and Faure & Josselin [150] lies in the replacement of r0,j 0,j by r1,j 0,j . Al- ! 0 ! 0 though this change seems minor, the improvement in the estimated rate coefficients is substantial. We will show this in detail in Sec. 5.3, where we compare the extrapolated rates of v = 2, j v = 1, j transitions with ! 0 0 results obtained directly from CC calculations. The ro-vibrational rate co- efficients r1,j 0,j (T) have been calculated using the full CC method, while ! 0 the vibrational rate coefficients rv v (T) were obtained from the IOS approx- ! 0 imation. All other ro-vibrational rate coefficients rv,j v ,j (T) can then be ! 0 0

Preliminary version – January 29, 2016 70 h-co rate coefficients for ro-vibrational transitions

obtained from the extrapolation formula, Eq. 5.5. Another advantage of our method is that it also yields the values for v, j v , j = j transitions; these ! 0 0 could not be obtained with the previous extrapolation method since the data for rotationally elastic j j transitions were not tabulated. And even ! if they had been available, they most likely would have given much too large extrapolated results.

−10 10 (a) v = 1 → v’ = 0 −11 10 ) − 1

s −12

3 10

−13 10

−14 10

Rate coefficient (cm −15 10 CC IOS −16 Exp. 10

−10 10 (b) v = 2 → v’ = 1 −11 10 ) − 1

s −12

3 10

−13 10

−14 10

Rate coefficient (cm −15 10 CC

−16 IOS 10

1 2 3 10 10 10 T (K)

Figure 5.1: Comparison of close-coupling and infinite order sudden rates for vibra- tional quenching v = 1 v = 0 and v = 2 v = 1. The experimental ! 0 ! 0 rates are from Glass & Kironde [120].

5.3 results

5.3.1 Rate coefficients from quantum scattering calculations

Cross sections for ro-vibrational v = 1, j = 0 - 30 v = 0, j transitions ! 0 0 from full CC calculations are reported in Chap. 4. We also concluded in Chap. 4 that the CS approximation is only suitable for collision energies above 1000 cm-1 for the H-CO system and it is therefore not applied ⇡ here. Table 5.1 lists the ro-vibrational transition rate coefficients for a gas temperature range of 10 K 6 T 6 3000 K calculated from these cross sectio-

Preliminary version – January 29, 2016 5.3 results 71 13 12 13 12 12 13 e- e- e- e- e- e- 573 148 462 016 509 696 ...... 13 7 13 1 13 8 13 1 12 1 13 9 e- e- e- e- e- e- 316 865 258 609 283 091 ...... 13 6 13 9 13 7 13 8 12 1 13 8 e- e- e- e- e- e- 999 046 900 837 026 375 ...... 13 8 13 5 13 1 13 6 13 4 13 6 e- e- e- e- e- e- 622 965 330 838 353 562 ...... transitions in the temperature transitions in the temperature 13 4 13 4 13 3 13 5 13 4 13 7 0 0 e- e- e- e- e- e- j j , , 488 465 087 623 096 628 ...... 0 1 = = 13 2 13 3 13 2 13 2 13 4 13 2 0 0 e- e- e- e- e- e- ) v v 260 418 068 793 298 081 10 ...... ! ! 14 1 14 1 14 1 14 1 14 2 14 1 30 30 e- e- e- e- e- e- - - ) 334 466 504 489 173 790 ...... 0 0 10 15 4 14 7 15 5 15 5 14 9 14 6 = = e- e- e- e- e- e- j j , , 1 2 025 111 571 767 754 466 ...... = = 16 6 15 1 15 9 15 9 15 1 15 1 v v e- e- e- e- e- e- 020 440 297 654 996 499 ...... 17 8 17 1 17 1 17 1 16 2 17 2 e- e- e- e- e- e- T (K) T (K) 777 135 265 222 011 688 ...... 17 5 17 5 17 3 17 9 17 1 17 6 e- e- e- e- e- e- 674 840 455 789 932 256 ...... with the exponent denoting power of with the exponent powers of 17 3 17 2 17 7 17 5 17 4 17 7 1 1 e- e- e- e- e- e- - - s s 881 900 516 433 770 746 ...... 3 3 17 1 17 4 17 3 17 2 17 6 17 6 e- e- e- e- e- e- 514 920 032 172 892 478 ...... 17 3 17 2 17 1 17 5 17 5 17 3 e- e- e- e- e- e- 449 679 267 139 645 588 ...... 17 2 17 1 17 1 17 5 17 4 17 3 K (in units of cm K (in units of cm e- e- e- e- e- e- 019 388 152 467 970 412 ...... 3000 3000 17 2 17 1 17 1 17 4 17 3 17 3 6 6 e- e- e- e- e- e- T T 733 281 170 888 475 441 ...... 6 6 17 1 17 1 17 1 17 3 17 3 17 3 K K e- e- e- e- e- e- 10 10 585 346 338 445 243 823 ...... 17 1 17 1 17 1 17 3 17 3 17 3 e- e- e- e- e- e- 600 627 751 269 537 962 ...... is published in its entirety in the electronic edition of the Astrophysical Journal. A portion is shown here for guidance regarding its form and content. is published in its entirety in the electronic edition of the Astrophysical Journal. A portion is shown here for guidance regarding its form and content. range of range of Rate coefficients from close-coupling calculations for ro-vibrational Rate coefficients from close-coupling calculations for ro-vibrational 1 2 : : . . 17 1 17 1 17 3 17 4 17 1 17 3 5 5 1 2 . . e- e- e- e- e- e- 5 5 10 20 30 40 50 6010 70 20 80 30 90 40 100 200 50 300 60 500 70 700 80 1000 1500 90 2000 100 2500 200 3000 300 500 700 1000 1500 2000 2500 3000 949 427 768 814 241 974 ...... Table Table 0 0 jj jj 001 012 022 Note 003 015 027 Note Table Table

Preliminary version – January 29, 2016 72 h-co rate coefficients for ro-vibrational transitions

Table 5.3: CO vibrational energy levels

-1 -1 -1 v✏v(cm ) v✏v(cm ) v✏v(cm ) 010825115341021331 132256135461123213 253427155311225069 374328174901326899 494969194241428704

ns with the use of Eq. 5.1. The highest final j0 = 27 - 42 values for which the rates are given in this table depend on the initial j quantum number. Transitions for even larger final j0 are not reported, either because they are negligibly small, or because they were not completely converged with the largest number of partial waves, i.e., the highest total angular momentum J included. However, we add all final j0 states, including those that may not be fully converged, to calculate the vibrational quenching cross sections CC CC (Ek) and the vibrational transition rate coefficients r (T). Hence, v,j v0 v v0 the!CC vibrational quenching rate coefficients shown in Fig. 5!.1 are slightly different from the values that would be obtained by summing and averag- ing only the reported rates rCC (T). v,j v0,j0 In order to test our proposed! extrapolation scheme, we also calculated cross sections and rate coefficients for ro-vibrational v = 2, j = 0 - 30 v = ! 0 1, j0 transitions with the CC method. We used a channel basis B4(75, 65, 54, 41, 20), where the notation Bn(j0, j1, j2, ..., jn) represents a basis with the highest vibrational level n and the highest rotational level ji for vibrational level v = i [see Chap. 4]. The state-to-state ro-vibrational rate coefficients were computed for a temperature range of 10 K 6 T 6 3000 K; they are listed in Table 5.2. In addition, we obtained vibrational quenching cross sections from scat- tering calculations in the IOS approximation with a v = 0 - 14 basis for CO. This basis of 15 vibrational levels, with energies given in Table 5.3, is suf- ficiently large to converge all v = 1 - 5 v

Preliminary version – January 29, 2016 5.3 results 73 applications. The IOS rate coefficients for v = 4 v = 1, 2, 3 transitions in ! 0 Table 5.4 at temperatures T = 40 and 50 K were obtained by interpolation of the rate coefficients at other temperatures. The original IOS data show a sharp peak around these temperatures due to resonance effects in the cross sections. However, the resonances in the IOS cross sections do not precisely coincide with the resonances in the cross sections from CC calculations [see Fig. 4.2 of Chap. 4].

10−10

−13 → ) ) → 10 (a) v = 1, j = 5 v’ = 0, j’ = 0 (b) v = 1, j = 5 v’ = 0, j’ = 6 − 1 − 1

s s −13 3 3 10 10−15

10−16 10−17

CC −19 CC −19 10 10 old extrapolation old extrapolation

Rate coefficient (cm in past modeling Rate coefficient (cm in past modeling 10−21 10−22 101 102 103 101 102 103 T(K) T(K)

10−10 10−11 ) ) − 1 − 1

s s −20 3 −16 3 10 10

−30 10−21 10 (c) v = 1, j = 5 → v’ = 0, j’ = 10 (d) v = 1, j = 5 → v’ = 0, j’ = 15

CC −40 CC 10−26 10 old extrapolation old extrapolation

Rate coefficient (cm in past modeling Rate coefficient (cm in past modeling 10−31 10−50 101 102 103 101 102 103 T(K) T(K)

Figure 5.2: Comparison of ro-vibrational transition rate coefficients from close- coupling calculations with extrapolated data obtained with the old ex- trapolation method of Thi et al. [9] and Faure & Josselin [150] (using vibrationally elastic data) and with the data used in astrophysical mod- eling [9].

5.3.2 Extrapolated rate coefficients

First, we test the validity of the extrapolation method used by Thi et al. [9] and Faure & Josselin [150]. Combining the IOS vibrational rate coefficient for the v = 1 v = 0 transition with pure rotational (vibrationally elastic) ! 0 rate coefficients from Walker et al. [149], both of them based on the 3D H-CO potential of Song et al. [106], we can obtain the state-to-state ro-vibrational rate coefficients for v = 1, j = 0 - 30 v = 0, j transitions by extrapolation. ! 0 0 Examples of the comparison of our CC rate coefficients, rate coefficients from extrapolation based on the method of Faure & Josselin [150] and Thi et al. [9], and the corresponding rate coefficients used by Thi et al. [9] are illustrated in Fig. 5.2. For transitions with final j0 less than or close to the

Preliminary version – January 29, 2016 74 h-co rate coefficients for ro-vibrational transitions 11 12 11 11 11 11 12 11 11 11 11 11 11 11 11 e- e- e- e- e- e- e- e- e- e- e- e- e- e- e- 107 200 376 912 632 328 394 899 243 053 406 144 674 879 148 ...... 12 6 11 5 11 3 11 2 11 1 12 3 11 2 11 6 11 3 12 1 11 5 11 3 11 2 11 5 11 5 e- e- e- e- e- e- e- e- e- e- e- e- e- e- e- 477 428 254 185 049 445 674 963 717 346 630 179 666 970 746 ...... 11 4 11 3 11 2 12 1 12 2 11 4 11 2 12 8 11 4 11 3 11 1 12 4 11 3 11 4 11 1 e- e- e- e- e- e- e- e- e- e- e- e- e- e- e- 355 502 677 511 548 811 005 476 498 421 209 877 833 554 141 ...... 11 2 11 1 12 7 13 1 11 2 12 5 11 3 11 2 12 1 12 2 11 3 11 2 11 3 12 1 11 3 e- e- e- e- e- e- e- e- e- e- e- e- e- e- e- 220 692 129 606 075 613 291 023 323 624 586 543 768 339 689 ...... 12 1 12 4 13 8 12 3 11 2 12 1 12 7 13 1 11 2 12 1 12 1 11 2 12 6 11 2 12 1 e- e- e- e- e- e- e- e- e- e- e- e- e- e- e- 704 718 070 974 112 164 169 163 481 838 116 206 145 858 398 ...... transitions in the temperature range 13 2 13 2 12 1 12 8 12 3 13 5 12 1 12 8 12 5 12 8 12 1 12 2 12 1 12 6 13 1 e- e- e- e- e- e- e- e- e- e- e- e- e- e- e-

Preliminary version – January 29, 2016 5.3 results 75

→ → ) (a) v = 2, j = 5 v’ = 1, j’ = 0 ) (b) v = 2, j = 5 v’ = 1, j’ = 10 −13 −13 − 1 − 1

s 10 s 10 3 3

10−15 10−15

10−17 10−17

CC CC 10−19 10−19 new extrapolation new extrapolation Rate coefficient (cm Rate coefficient (cm in past modeling in past modeling 10−21 10−21 101 102 103 101 102 103 T(K) T(K)

→ −13 → ) (c) v = 2, j = 5 v’ = 1, j’ = 16 ) 10 (d) v = 2, j = 5 v’ = 1, j’ = 30 −13 − 1 − 1

s 10 s 3 3 10−17 10−15

−21 −17 10 10

CC −19 10−25 10 new extrapolation CC Rate coefficient (cm in past modeling Rate coefficient (cm new extrapolation 10−21 10−29 101 102 103 101 102 103 T(K) T(K)

Figure 5.3: Comparison of ro-vibrational transition rate coefficients from close- coupling calculations with extrapolated data obtained with our new extrapolation method (using vibrationally inelastic data) and with the data used in astrophysical modeling [9]. initial j, i.e., the v = 1, j = 5 v = 0, j = 0, 6 transitions, the CC and ! 0 0 extrapolated rate coefficients agree quite well with each other, as shown in Figs. 5.2(a) and (b). Large discrepancies appear, however, between CC and extrapolated rate coefficients for transitions with larger j = j0 - j, espe- cially at lower temperatures, see Figs. 5.2(c) and (d). These large discrepan- cies result because the pure rotational (vibrationally elastic) transitions in v = 0 are endoergic for j0 >j, while the ro-vibrational quenching processes v = 1, j v = 0, j are exoergic for all j <33even when j = 0. In addition, ! 0 0 0 by analogy with Eq. (5.6), the vibration-related scale factor P (T)(v = v ) vv0 6 0 can be rewritten in the extrapolation method of Thi et al. [9] and Faure & Josselin [150] as rv v (T)/r0 0(T). Clearly, the validity of scaling inelastic ! 0 ! vibrational transition rate coefficients with vibrationally elastic data is ques- tionable. The rate coefficients adopted in the modeling of Thi et al. [9] are even more discrepant, as shown by the green lines with triangle markers in Fig. 5.2. In their extrapolations, the pure rotational rates were adopted from Balakrishnan et al. [16] who used the WKS potential, known to be in- accurate at long range. Moreover, the pure rotational rates for initial states with j>7and the vibrational rates for temperatures below 100 K were not available and had to be obtained by extrapolation. Let us now discuss the results from our new extrapolation method based on vibrationally inelastic rates from quantum scattering calculations. Us- ing the data in Tables 5.1 and 5.4, the ro-vibrational rate coefficients for

Preliminary version – January 29, 2016 76 h-co rate coefficients for ro-vibrational transitions

v = 2 - 5, j = 0 - 30 v

5.4 astrophysical models

In order to show the relevance of these new collision rate coefficients in the astrophysical context, we chose a twofold approach: (1) simple 1D slab models with constant temperature, density and CO abundance and (2) the 2D radiative thermo-chemical disk model from Thi et al. [9]. In the follow- ing, we compare the results obtained with the old H-CO collision rates to those obtained with the new data. Since this chapter presents new collision rates for v 6 5 and j 6 30, we restricted the calculations below to the same range of quantum numbers.

5.4.1 ProDiMo slab and disk models

ProDiMo is a 2D radiation thermo-chemical disk code [152]. The code solves the 2D continuum radiative transfer to obtain the dust temperature and radiation field throughout the disk and provides the self-consistent solution for the chemistry and gas heating/cooling balance. The gas temperatures can be used iteratively to find a vertical hydrostatic equilibrium solution. However, we use for this work a fixed parametrized gas scale height and assume that gas and dust are well mixed. Details of the numerical methods, the chemical network and a list of heating/cooling processes can be found in Woitke et al. [152]. The CO molecular data used are described in detail above and in Thi et al. [9]. The code can be used in 1D slab mode, to calculate the emission emerging from a fixed total gas column with a constant gas temperature and constant - volume densities of collision partners (H, H2, He and e ). The non-LTE level populations are calculated using the escape probability method taking into account also IR pumping by thermal dust emission. Since we focus here on the effect of the new collisional rates, we minimize the effect of IR pumping

Preliminary version – January 29, 2016 5.4 astrophysical models 77 by fixing the dust temperature at 20 K. The turbulent width is fixed to 2 2 vturb = 1.0 km/s and the total broadening is given by b = vth + vturb, where vth is the thermal velocity of the CO molecules. The slabq models provide a powerful means to exploit a very wide parameter space under which CO emission could arise in space. In addition, we use the standard model for a disk around a 2.2 M Herbig star (L = 32 L ) taken from Thi et al. [9]. The disk has a radial size of ⇤ 300 AU and a mass of 10-2 M . More details on the dust opacities and vertical disk structure can be found in Table 3 of Thi et al. [9]. Such a disk model does not provide any freedom as to the conditions under which CO is emitting. Its abundance, excitation and emission are calculated self- consistently given the density distribution of gas within the disk and the irradiating stellar and interstellar radiation field.

5.4.2 Slab model results

We ran four series of models, each with three different total gas column 19 21 23 -2 densities of N H = 10 , 10 , and 10 cm . The CO abundance is fixed h i at 10-4 with respect to the total hydrogen number density. Series 1 uses a fixed gas temperature T of 800 K and high densities of collision partners, -3 log nH = log nH2 = 9, log nHe = 8, log ne = 5 (in cm ). Note that the gas in these slab models is not fully molecular, i.e., the H/H2 abundance ratio is 1. Series 2 uses a lower gas temperature of 200 K and the same collision partner densities. Series 3 uses T = 800 K and a lower density of collision -3 partners, log nH = log nH2 = 6, log nHe = 5, log ne = 2 (in cm ). An ad- ditional series 4 was calculated to isolate the effect of H-collisions. In this series, the temperature and density of H are the same as in series 3, while the collision partner densities of H2, He and electrons are assumed to be negligible. Figure 5.4 illustrates the change in the non-LTE fluxes of the v = 1 - 0 band resulting from the old and new collisional rates for H-CO in the slab models. We note that changes are typically smallest for the slab with the largest line optical depth log N H =23 (log NCO =19). For slabs at T =800 K, the line fluxes with the new ratesh i are smaller than those with the old rates. Exceptions are the smaller and the highest j levels; the flux changes are generally largest for the lowest four levels. It is also evident from these comparisons that the implementation of the new rates will have an effect on the shape of the vibrational band structure. In most cases of high T (800 K) slabs, the band structure (flux vs wavelength) will get flatter for low j and drop faster at high j compared to the band structure with the old collision data. The total cooling rate of the v = 1 - 0 band (j 6 30) is typically 20-30% lower with the new collisional rates. This will have an impact on the gas temperature whenever CO cooling dominates the energy balance. Examples are (1) the surface layers of the inner 10 pc of active ⇡ galactic nucleus (AGN) disks [153], (2) the inner 10 AU of protoplanetary ⇡ 5 -3 disks [152] and (3) intermediate density (n H 10 cm ) J-type shocks h i ⇡

Preliminary version – January 29, 2016 78 h-co rate coefficients for ro-vibrational transitions

100 R-branch P-branch 100 R-branch P-branch v=1-0 P4 log N=19 v=1-0 P4 log N=19 log N=21 log N=21 [%] log N =23 [%] log N =23 50 50 J,old J,old )/F )/F J,old J,old

-F 0 -F 0 J,new J,new (F (F -50 -50

4.5 4.6 4.7 4.8 4.9 5.0 4.5 4.6 4.7 4.8 4.9 5.0 λ (µm) λ (µm)

100 R-branch P-branch 100 R-branch P-branch v=1-0 P4 log N=19 v=1-0 P4 log N=19 log N=21 log N=21 [%] log N =23 [%] log N =23 50 50 J,old J,old )/F )/F J,old J,old

-F 0 -F 0 J,new J,new (F (F -50 -50

4.5 4.6 4.7 4.8 4.9 5.0 4.5 4.6 4.7 4.8 4.9 5.0 λ (µm) λ (µm)

Figure 5.4: Changes in the modeled CO ro-vibrational emission from a series of slab models using the new and previously available H-CO rate coeffi-

cients. Top left (series 1): T = 800 K, log nH,H2 = 9, log nHe = 8, log ne = 5 -3 in units of cm . Top right (series 2): T =200 K, log nH,H2 =9, log nHe =8,

log ne = 5. Bottom left (series 3): T = 800 K, log nH,H2 = 6, log nHe = 5, log ne = 2. Bottom right (series 4): T = 800 K, log nH = 6, collisions with other partners negligible. The total gas column density of the slab N H is given in the legend. h i

[154]. If H collisions are omitted entirely in slab series 1, the fluxes in the most optically thick case are lower by a factor of 100; hence H collisions are more important than H2 collisions and they are key in modeling the CO ro-vibrational emission. The gas temperature T is a free parameter and we can explore the effect of the new collision rates at lower temperatures. The comparison between series 1 and 2 (top left and right panel of Fig. 5.4) clearly shows that the changes in the line fluxes increase substantially for T = 200 K. However, at these low temperatures, the populations in the CO vibrational levels are very low and hence also the emerging line fluxes will be low. An exception may be the case of fluorescent excitation, where the vibrational levels are populated by an external radiation field to non-LTE values. Examples are the inner disks of Herbig Ae stars, where the stellar radiation field can substantially pump the vibrational levels to vibrational temperatures of a few thousand K [143, 155]. The bottom panels of Fig. 5.4 illustrate the effects for low density environ- ments (non-LTE cases). Here, line-to-line differences are larger than in the high density environment. The overall picture does not change very much if we neglect the other collision partners, H2, He, and electrons.

Preliminary version – January 29, 2016 5.4 astrophysical models 79

Besides the application to protoplanetary disks discussed in the next sec- tion, these new rate coefficients will also be important for proper modeling of CO infrared emission from other interstellar regions with large H/H2 transition zones such as dissociative J-type shocks [32, 156] or dense X-ray irradiated gas found near Active Galactic Nuclei.

20 R-branch P-branch v=1-0 P4

10 [%] J,old

)/F 0 J,old -F J,new (F -10

-20 4.5 4.6 4.7 4.8 4.9 5.0 λ (µm)

Figure 5.5: Changes in the modeled CO ro-vibrational emission from the disk around a Herbig Ae star using the new and previously available H-CO rate coefficients.

5.4.3 Disk model results

Figure 5.5 shows the change in the CO ro-vibrational emission of the v = 1-0 band at 4.7µm resulting from the old and the new (this work) H-CO collision rate coefficients. The collision rate coefficients with H2, He, and electrons remain unchanged. Differences are small, on the order of 5%. ± The line fluxes with the new rates show a clear pattern with respect to the old ones: the new rates lead to lower line fluxes at low j, higher fluxes in the intermediate range and then again lower fluxes for j & 25. The upturn at the highest j is a boundary effect, since the rotational states of CO are artificially cut off in the model at j = 30. As noted in Thi et al. [9], the CO emission arises from a hot surface layer with T 1000 K, where the ⇡ gas is predominantly atomic (see their Fig.12). At these high temperatures, differences between the old and new collision rates are smaller than at low temperatures (Fig. 5.2) and this partially explains the relatively small changes observed in the disk model. In addition, IR pumping is competing with collisions, especially in the inner disk, where dust continuum emission peaks in the NIR (1000 K

Preliminary version – January 29, 2016 80 h-co rate coefficients for ro-vibrational transitions

5.5 conclusions

We present ro-vibrational de-excitation rate coefficients from full CC quan- tum scattering calculations for v = 1, j = 0 - 30 v = 0, j and v = ! 0 0 2, j = 0 - 30 v = 1, j transitions of CO in collision with H atoms in the ! 0 0 temperature range of 10 K 6 T 6 3000 K. Also vibrational quenching rate coefficients from IOS calculations are given for v = 1 - 5 v

Preliminary version – January 29, 2016 A r -CO SCATTERING FOR PURE ROTATIONAL TRANSI- 6 TIONS

We present a novel means of analyzing velocity-map images of angular mo- mentum polarization in inelastic scattering. In this approach, linear com- binations of angular distributions obtained by integrating select regions of images obtained with two probe laser polarizations directly yield the alignment-free differential cross sections and the differential alignment mo- ments. No fitting is needed in the analysis. The method relies on the fact that the angular distribution for out-of-plane scattering is encoded in the distribution along the relative velocity vector, and this may be recovered quantitatively owing to the redundancy of the in-plane and out-of-plane scattering for the horizontal polarization case.

Arthur G. Suits, Chandan Kumar Bishwakarma, Lei Song, Gerrit C. Groe- nenboom, Ad van der Avoird, David H. Parker, Direct Extraction of Align- ment Moments from Inelastic Scattering Images, J. Phys. Chem. A, 119, 5925 (2015)

Preliminary version – January 29, 2016 82 ar-co scattering for pure rotational transitions

6.1 introduction

Despite two decades of active investigation, imaging studies of inelastic scattering continue to offer surprising new insights into fundamental as- pects of atomic and molecular interactions.[157, 158, 159, 160, 161, 162, 163] These studies allow stringent tests of the quality of electronic structure cal- culations and stimulate efforts to combine the intuition afforded by classi- cal treatments with the rigors of a full quantum description. Among the more intriguing such examples in recent years has been the direct detection of collision-induced polarization in scattering.[29, 164, 165, 166, 167, 168, 169, 170] The power of imaging in these experiments is that it combines a spectroscopic probe (i.e., resonance-enhanced ionization) with angular distribution measurements yielding a molecular-frame description of the scattering event.[171, 172] The study of collision-induced angular momen- tum polarization is an example of 3-vector correlation: the initial relative velocity vector, the final relative velocity vector, and the angular momen- tum projection and their correlations are simultaneously measured.[166, 173, 174, 175, 176, 177] Meyer reported early investigations of rotational alignment in state-resolved scattering of NH3 and NO[164, 178] which was followed by imaging studies by Chandler, Cline and coworkers on the argon-NO system [165, 168]. Studies by Chandler and Cline and cowork- ers also showed that such collisions could give rise locally to rotational orientation, and the sense of this orientation and its angular distribution could be measured using circularly polarized probe lasers. This continues to be an active area of investigation.[163, 167] Very recently, Chandler and coworkers have studied scattering of electronically excited NO, even exam- ining alignment in that system[162, 179], and Brouard and coworkers have revisited the issue of rotational alignment in the argon-NO system and com- bined this with both full quantum calculations and with a quasi-quantum treatment.[29, 169, 180] In essentially all of these inelastic scattering alignment studies, the anal- ysis is performed by taking difference images obtained with probe polar- ization parallel to the detector (“H”) or perpendicular to the detector (“V”). These difference images are then fitted using various means to extract the alignment moments. For rotational alignment with one-photon probe there ( 2 ) ( 2 ) are two such moments, A 0 and A 2 + (we will employ the Hertel-Stoll renormalized polarization-dependent differential cross sections[181, 182]) whose angular distributions are reported and related to classical treatments or to full time-independent quantum scattering calculations. Although these approaches to the analysis are clearly successful, we are naturally led to wonder if there might be a more direct way to extract these distributions from high quality scattering data rather than through fitting methods. With the remarkable resolution of velocity map imaging and the simple clarity of these images such an objective seems attainable. If we are successful, any structures in the distributions will appear automatically and we need not wonder if our analysis has overlooked some or artificially imposed others.

Preliminary version – January 29, 2016 6.2 experimental methods 83

In this chapter we show just such a direct means of extracting the three ( 2 ) key distributions: the differential cross sections (DCS) and the A 0 and ( 2 ) A 2 + moments, simply through linear combinations of angular distribu- tions obtained by integrating various regions of images obtained for H and V polarization. We will illustrate this using a single experimental image for final j CO = 9, obtained for scattering of CO with argon at a collision energy of 700 cm- 1 , and we compare our extracted distributions with quantum scattering calculations for this single rotational level. This level is chosen as it has broad scattering and is spectroscopically well-isolated. A full presen- tation of all the scattering results for the system and a detailed comparison to theory will be reported in a future publication.

6.2 experimental methods

The data were collected in a rotatable crossed molecular beam machine using 1+10 VUV REMPI ionization of the scattered products with velocity map imaging detection. We use two skimmed supersonic beams, one of neat argon produced from a Nijmegen pulsed valve[183] and the other of 5% CO seeded in argon produced from a Jordan pulsed valve. The backing pressure for the expansion was 1 bar for both colliding partners. The pulse duration of the Nijmegen pulsed valve is about 50 µs and the pulse dura- tion of the Jordan valve is about 60 µs. The source chamber that houses the primary beam (CO) was differentially pumped and the molecular beam passed through a skimmer of aperture diameter 2.5 mm positioned 7 cm away from the nozzle. The secondary molecular beam (Ar) source was mounted in the differentially pumped rotatable chamber positioned 3 cm from the skimmer of aperture diameter 2.5 mm. Turbo molecular pumps pumped all the three chambers. The two molecular beams and the VUV laser were coplanar parallel to the detector and were optimally aligned to ensure the best signal. We have used four-wave difference mixing in xenon to produce tunable VUV light.[184, 185] We used two dye lasers: one (Scan- mate) was fixed near the xenon resonance at 249.618 nm (!1) and the other dye laser (Fine Adjustment) tuned around 650 nm (!2) to produce tunable VUV (!VUV) light around 155 nm. The 355 and 532 nm output from an Nd:YAG laser (Continuum Powerlite 9010) were used to pump the two dye lasers. The two beams were combined at a 248 dichroic mirror transparent to visible light and both beams subsequently focused into the Xe gas (30 mbar) contained in a stainless steel cell with a f = 100 mm plano-convex lens. A lens of focal length f = 1500 mm is placed 1000 mm away from the stainless steel cell to collimate only the visible light so that 249 nm and 650 nm will focus at the same point inside the cell. In order to study alignment effects we have used horizontal and vertical polarization of the VUV laser. The VUV was focused onto the interaction region by a magnesium fluoride lens. Its orientation was adjusted so that it did not alter the polarization of the visible and UV beams, and photodissociation images of OCS were used to confirm the desired polarization of VUV laser.[186] Separate images

Preliminary version – January 29, 2016 84 ar-co scattering for pure rotational transitions

were recorded independently for H and V polarization using a wave plate to rotate the polarization of the visible beam , which directly determines the VUV beam polarization. We chose the Q branch for our data collec- tion, as it is well resolved and it shows greater sensitivity to alignment effects compared to P and R branch transitions. The laser and both molecu- lar beams were operated at 10 Hz. The scattered product was ionized and was extracted by a velocity map imaging ion optics lens.[172] After 85 cm free flight along the time of flight tube, the 3-D ion sphere was projected onto a 2-D detector, which is read by a CCD camera. Application of gating enables us to select our required mass and filter out the background ions of other masses. Scattered CO molecules were probed by the VUV laser on the A1⇧ - X1⌃+ (0-0) transitions.

6.3 theoretical methods

Integral and differential state-to-state cross sections for CO-Ar collisions and the alignment of CO after the collision were computed with the close- coupling method. The intermolecular potential used in the calculations was obtained by Pedersen et al. from ab initio electronic structure calculations with the coupled cluster method [187]. We refer to this reference for de- tails of these calculations and the analytic fit of the potential. Their three- dimensional (3D) CO-Ar potential depends on the two intermolecular Ja- cobi coordinates, as well as on the CO bond length r. We used the 2D potential also given in Ref. [187] that was obtained by averaging the 3D potential over the ground vibrational (v = 0) state of CO. The close-coupling equations were solved with the renormalized Nu- merov propagator, with R ranging from 4 to 40 a0 in 900 equal-sized steps. All rotational states of CO up to j = 25 were included and all partial wave contributions up to a total angular momentum of J = 200 were taken into account. The experimental collision energy was estimated to be 700 cm-1, but there is a certain spread in this energy. Hence, the calculations were per- formed for three different collision energies, 650, 700, and 750 cm-1, and the cross sections were averaged with weights of 25, 50, and 25%, respec- tively. The state-to-state cross sections and CO alignment moments (discussed below) were computed for initial CO states with j = 0 and j = 1 and for all final j states open at the given collision energies. Our measurements showed that the incoming CO beam contained 80% of j = 0 and 20% of j = 1, and so we averaged the calculated results over the initial j = 0 and j = 1 states of CO with these percentages. The results discussed below refer in particular to a final CO state with j = 9. (2) (2) The alignment moments A0 and A2+ are the irreducible components of the scattering density matrix [29]. The diagonal elements of this density (2) matrix appear in the moment A0 and represent the populations in the different magnetic quantum levels, while the off-diagonal elements appear-

Preliminary version – January 29, 2016 6.4 results and discussion 85

Figure 6.1: CO Q(9) images for CO-Ar scattering and indicated probe polarization. For the H image, the associated Newton diagram is shown. The image intensities are arbitrarily scaled relative to each other.

(2) ing in the A2+ moment reveal information about the coherences existing between different eigenstates, which are essential for the azimuthal polar- ization [182]. The scattering angle dependent density matrix is expressed in the scattering amplitudes related to the S-matrix, after undoing the angular momentum coupling used in the close-coupling calculations for different to- tal angular momenta J [170, 188]. This uncoupling transformation is needed to resolve the different magnetic (mj) sublevels of the final j states, as well as the scattering angle dependence of the DCSs and alignment moments.

6.4 results and discussion

6.4.1 Overview

Our objective is to extract the DCS and the rank 2 alignment moments directly from inelastic scattering data. Our approach is to use Q-branch im- ages for perpendicular probe transitions and consider horizontal (H) and vertical (V) polarization, where H refers to the probe polarization lying par- allel to the plane of the image, and V, normal to that plane. The images that we will use for demonstration are given in Fig. 6.1. These were obtained via the Q(9) transition of CO via the A state as described in the Experimental section. The difference in the appearance of these two images is a direct manifestation of angular momentum polarization. In its absence, these im- ages would be identical. The asymmetry across the relative velocity vector arises owing to the fact that we detect product density, so the products that are slower in the laboratory frame are detected more efficiently. This is the familiar density-to-flux issue that is always faced in such studies. We find it can readily be corrected for the case of inelastic scattering by dividing the pixel intensity by (↵ + v), where v is the lab frame velocity and ↵ is an adjustable parameter that accounts for the finite size of the interaction

Preliminary version – January 29, 2016 86 ar-co scattering for pure rotational transitions

Figure 6.2: Coordinate frames employed. The alignment moments are defined in the Collision Frame.

volume and the finite duration of the collision event. However, as discussed in the Sensitivity section below, we determine the alignment moments with the uncorrected data. The relevant coordinate frames we employ are shown in Fig. 6.2. Our probe laser propagation direction is in the plane of the beams, nominally perpendicular to the relative velocity vector. This defines the Laboratory Frame in which the Z-axis coincides with the CO beam di- rection in the center-of-mass, the X-axis is parallel to the laser propagation direction, and the Y-axis coincides with the time-of-flight axis. The align- ment moments are obtained by analyzing the distributions in the Collision Frame, in which Zcoll again coincides with the the CO beam direction in the center of mass frame, but Xcoll is now in the plane of the scattered CO, perpendicular to Zcoll. Ycoll is then perpendicular to Zcoll and Xcoll. For CO scattering in the plane of the beams, the Laboratory and Collision Frames coincide. For scattering of CO out of the plane, the Collision Frame effec- tively represents a rotation of the Laboratory Frame about the Z axis. In the conventional approach, images are recorded with H or V probe laser po- larizations and the difference between these is fitted using various means to obtain the alignment moments. Our strategy instead relies on consider- ing just three limiting distributions. We first describe intuitively how this works and how we do this analysis in practice, then we present the under- lying mathematical justification.

Preliminary version – January 29, 2016 6.4 results and discussion 87

Figure 6.3: Relation of the lab frame and collision frame for the four angular dis- tributions obtained from the images.

The analysis is based upon the fact that the alignment-free differential cross section (DCS) and alignment moments may be obtained rigorously from the Collision Frame angular distributions for each of three orthogonal probe polarizations: One with probe laser polarization along Zcoll, paral- lel to the relative velocity vector, (termed H, in-plane or “HIP”), one with the probe laser polarization parallel to Ycoll: This is the in-plane scattering for V polarization (V,-in-plane, “VIP”). And finally from the scattering for probe polarization along Xcoll. But for the in-plane scattering, Xcoll also corresponds to the laser propagation direction, so this polarization geome- try is not accessible. However, this geometry is equivalent to the Collision Frame which is rotated 90o from the Laboratory Frame for Vertical polar- ization as illustrated in Fig. 6.3. If we can obtain the angular distribution for the products scattered perpendicular to the plane of the beams (and the detector) for V polarization, we obtain the missing geometry needed to extract the parameters directly from the distributions. We have found that by analyzing a narrow rectangular stripe along the relative velocity vector and converting this to an angular distribution, we are able to obtain the distribution corresponding to the third geometry, which we designate V, out-of-plane (“VOOP”). To obtain this distribution quantitatively, we first analyze and compare the equivalent image for H polarization (“HOOP”). In this case the two distributions HOOP and HIP should be identical. We typically adjust the limits of the region of integration (i.e., the box position and size) so that the features in the distributions are coincident. To convert from the rectangular segment to the HOOP and VOOP distributions we first integrate pixel intensities in the segment along the width (perpendicular to the relative velocity vector), then convert the values at each point alogn the

Preliminary version – January 29, 2016 88 ar-co scattering for pure rotational transitions

relative velocity vector to angle by taking the inverse cosine of the distance from the center-of-mass origin divided by the nominal recoil velocity in pixels. These are then interpolated onto the same linear grid as the in-plane distributions for further manipulation. We note that to achieve reasonable angular resolution at the poles high resolution scattering data is required. For event-counted data such as is presented here, with images at 100 pixel radius the velocity resolution at the poles for the HOOP and VOOP distri- butions is 8 degrees. Lower resolution images may be employed of course, albeit with a sacrifice of detail for the first few points near the poles.

Figure 6.4: CO Q(9) images for indicated probe polarizations and the four distri- butions extracted from them. These images have been density-to-flux corrected, but as we note below, this is used in determining the full DCS but for the alignment moments we use the uncorrected images. Although the overall intensity scale for the angular distributions is ar- bitrary, the two plots are on the same scale.

We then adjust the scaling so that the HOOP distribution has the same integral as the HIP distribution. An example of this analysis is shown in Fig.6.4. When we obtain the appropriate segment of the image that gives two nearly-identical angular distributions for H images, we employ the same rectangular image segment, integration and scaling for V images obtained under otherwise identical conditions. This yields three effective angular distributions: IH(✓), IVIP(✓) and IVOOP(✓). The alignment-free dif- ferential cross sections and the polarization moments are then obtained simply as appropriate linear combinations of these three distributions. The

Preliminary version – January 29, 2016 6.4 results and discussion 89 sensitivity of the reported distributions to various aspects of the analysis is examined below. We note also that this strategy, which rests on the recog- nition that the angular distribution may be accurately obtained simply by integrating the region along the relative velocity vector, may find other uses, for example in photodissociation or reactive scattering particularly involv- ing polarization studies.

6.4.2 Mathematical foundation

To obtain the form of these distributions, we loosely follow the approach of Brouard et al.[180], which in turn rests on a celebrated paper of Fano and Macek[189], although many have made contributions in this area. In addi- tion to the alignment-free DCS, there are two angle-dependent parameters that embody the description of the alignment distribution for collisions of (2) (2) (2) unpolarized beams: A0 and A2+. A0 embodies the v-J correlation, with limiting values of -1, corresponding to v J and +2 corresponding to v (2) ? J. A2+ reflects the extent to which the J vectors align parallel to Ycoll k (2) (2) (A2+ =-1) or Xcoll (A2+ =+1).[181, 182] We write the signal appearing in a given angular range in the image as a product of the DCS and the detection probability:

d I(✓)= (✓)P(✓; , ⇥, ) (6.1) d✓ The polarization-dependent transition probability P(✓; , ⇥, ) depends both on the angular momentum polarization and the probe sensitivity to that polarization[180]:

k k P(✓; , ⇥, )=C 1 + A( )(✓)F( )(, ⇥, ) (6.2) 2 q q 3 Xkq 4 k 5 In this expression C is a constant and the F( )(, ⇥, ) functions are geomet- q± ric factors that include the probe sensitivity as determined by the angles (, ⇥, ) characterizing the linear probe laser polarization: ⇥ and are the polar and azimuthal angles of klaser in the Collision Frame, and is the third Euler angle needed to specify the laser polarization direction in that frame. These functions take the general form[180]

2 1 F( )(, ⇥, )= h(2)(j , j )c (j )(3 sin2 ⇥ cos 2 -[3 cos2 ⇥ - 1]) (6.3) 0 4 i f 2 i 2 p3 F( )(, ⇥, )= h(2)(j , j )c (j )(2 sin ⇥ cos sin 2 (6.4) 1+ 4 i f 2 i +2 sin ⇥ cos ⇥ sin cos 2 - sin 2⇥ cos )

2 p3 F( )(, ⇥, )= h(2)(j , j )c (j )([1 + cos2 ⇥] cos 2 cos 2 (6.5) 2+ 4 i f 2 i -2 cos ⇥ sin 2 sin 2 - sin2 ⇥ cos 2) .

Preliminary version – January 29, 2016 90 ar-co scattering for pure rotational transitions

Table 6.1: Geometric factors for indicated distributions.

HIP HOOP VIP VOOP

⇡ ⇡ ⇡ ⇡ ⇡ ⇡ ⇡ ⇡ (, ⇥, ) (⇡, 2 , 0)(⇡, 2 , 2 )(2 , 2 , 0)(2 , 2 , 2 ) (2) 1 1 F0 11- 2 - 2 (2) p3 p3 F2+ 00- 2 2

They are functions of the laser polarization direction and the line strength factor h(2), which for Q-branch transitions is always unity.[189] We will confine our treatment to scattering directly in the detector plane or perpen- dicular to it. We then have only two polarization geometries for each of these two collision frame distributions, characterized by particular values of (, ⇥, ) and given in Table I. Substituting these values into eq. (6.2)we obtain the following expressions for the angular distributions in the differ- ent probe geometries:

d 2 I (✓)= (✓)C[1 + A( )(✓)] (6.6) H d✓ 0

d 1 2 p3 2 I (✓)= (✓)C[1 - A( )(✓)- A( )(✓)] (6.7) VIP d✓ 2 0 2 2+

d 1 2 p3 2 I (✓)= (✓)C[1 - A( )(✓)+ A( )(✓)] (6.8) VOOP d✓ 2 0 2 2+ We give only one H result as HIP and HOOP are identical. Table 6.1 and these equations show a number of key features of the polarization sensitiv- ity. First, it is seen that the scattering signal for the H polarization has no (2) sensitivity to the A2+ parameter. This is expected given the cylindrical sym- metry of this geometry. Furthermore, if we compare V and H in-plane, for example by taking a difference image as is commonly done, we can elim- (2) inate the population contribution but we cannot disentangle the A0 and (2) A2+ contributions. However, by averaging these three angular distributions together we obtain a distribution directly proportional to the DCS, free of any polarization modulation. We can then use this to normalize the other distributions to isolate the alignment moments. By taking the difference of the VOOP and VIP distributions we obtain a signal that is directly propor- (2) tional to the A2+ parameter (with the same proportionality as the DCS). (2) Finally, the A0 moment is obtained simply from the H distribution and the alignment-free DCS. The relevant expressions are

d 1 (✓)= [I (✓)+I (✓)+I (✓)] (6.9) d✓ 3 H VIP VOOP

Preliminary version – January 29, 2016 6.4 results and discussion 91

0.4

0.4 0 0

-0.4 -0.4

-0.8 -0.8

0 0

-0.2 -0.4

-0.4 -0.8

-0.6 0 45 90 135 180 0 45 90 135 180 CM Angle (deg) CM Angle (deg)

Figure 6.5: DCS and alignment moments obtained from the analysis (left) and from theory (right).

(2) p3[IVOOP(✓)-IVIP(✓)] A2+(✓)= (6.10) IH(✓)+IVIP(✓)+IVOOP(✓)

(2) 3IH(✓)-[IH(✓)+IVIP(✓)+IVOOP(✓)] A0 (✓)= (6.11) IH(✓)+IVIP(✓)+IVOOP(✓) The results of this analysis for the images in Fig. 6.1 are shown in Fig. 6.5 and compared to theory. The overall trends are reproduced quite well but one should take note of the different scales. We have assumed the laser propagation direction is perpendicular to the relative velocity vector. Devi- ations from this will primarily attenuate the observed alignment, and this may in part account for some of the discrepancy between the experiment and theory. A number of qualitative features are seen in the theory and reproduced in the experimental analysis, and the details of this will be discussed in a (2) future publication. However, we make a few observations here. The A0

Preliminary version – January 29, 2016 92 ar-co scattering for pure rotational transitions

moment rises from a highly negative value in the backward direction to a positive value in the forward direction, as is widely seen and rationalized in the kinematic apse (KA) picture. In this view the collision is sudden and the angular momentum change is perpendicular to the direction of linear mo- mentum transfer. The result is that for the directly backscattered product, (2) the angular momentum has no projection on the recoil direction (A0 =-1, J v), while in the forward direction there is a tendency to positive align- ? ment, J v. For the pure forward scattering the projection of J onto the k recoil direction must vanish by symmetry, but as is usually the case, this sharp transition is not detected experimentally. The other notable feature (2) is that the overall magnitude of the A0 moment is consistently lower in experiment than in the theory in the backward direction. This behavior is seen in our studies involving CO scattering, and in virtually all NO stud- ies reported previously as well. One possible explanation is the truncation of the expansion at the second moment and neglect of higher moments (2) that might make important contributions. The A2+ moment shows inter- esting structure in the theory that is clearly captured in the experiment. It increases from a minimum at 45o to zero for the most forward scattered (2) (2) product. This sharp increase in A2+, along with the rise in the A0 moment in the same region, combine (owing to opposite signs) to give rise to the gap in the forward scattering for the V image (and to the forward notch in the difference images commonly shown).

6.4.3 Sensitivity

We also examined the sensitivity of the inferred moments and the DCS to various aspects of the analysis. One point we note is that to obtain accu- rate DCSs that exhibit some of the structure in the theoretical results, we must include the slow (in the lab frame) portion of the angular distribu- tion. However, when we employ this strategy to generate the alignment moments, substantial deviations from expected values are obtained and it is difficult to obtain HIP and HOOP distributions that agree. It may be that these images are partly “sliced” so that the out-of-plane scattering is undercounted. Or there may be some nonlinearities in the detection that are equalized for the HOOP and the HIP obtained from the fast (in the lab frame) side of the scattering image. A final possibility is that stray magnetic fields or collisions perturb the alignment for the slower products. We plan future experiments to address these questions. The general sensitivity analysis results are compiled in Fig. 6.6. In each case we simply superimpose the results for a range of different parameter sets to convey the range of the distributions obtained. Figure 6.6A shows the consequence of changing the width of the rectangular “box” used for the VOOP integration from 8 to 30 pixels. It is clear that the distributions are quite insensitive to this value. Similarly, Fig. 6.6B shows the results for changing the width of the ring integrated to give the VIP and HIP distri- butions over a range of 5 to 23 pixels. There is some slight variation in the

Preliminary version – January 29, 2016 6.4 results and discussion 93

Figure 6.6: Results of sensitivity analysis. Variations in box width (A), ring width (B), box origin (C), relative image intensity (D) and relative image posi- tion (E) were examined (see text).

DCS and the q = 0 moment, and greater associated variation in the q = 2 moment. Fig. 6.6C shows the result of changing the origin of the box by up to four pixels. At this point the HIP and HOOP distributions become notice- ably dissimilar, yet only the forward scattered edge of the q = 2 moment shows any significant sensitivity to this. Fig. 6.6D shows the result of reduc- ing the V image intensity by 5 and 10%. There is surprisingly little variation in any of the extracted distributions in this case, which is heartening given that it is always difficult to ensure that the intensities are well controlled as the polarization is changed. The last panel, Fig. 6.6E, shows the result of translating the V image by 2 and 3 pixels in X and Y. This has a large effect (2) (2) on the A2+ moment and a significant effect on the A0 moment but again little impact on the inferred DCS. This would only occur if the image focus- ing conditions or camera moved in the course of changing the polarization. In all, the results presented in Fig. 6.6 show that this approach is remark- ably robust. As a final comment, we note that other approaches to analysis of alignment data often incorporate full modeling of the instrument prop- erties such as beam velocity spreads and intersection angles. We do not do this here, so this approach is restricted to well-defined experimental con- ditions and high-resolution images. An interesting possibility would be to combine the approach presented here with forward convolution modeling in an iterative procdure.

Preliminary version – January 29, 2016 94 ar-co scattering for pure rotational transitions

6.5 conclusion

We have outlined a simple approach to extract alignment moments and the alignment-free DCS directly from state-resolved images of inelastic scat- tering. The approach is demonstrated in application to velocity-mapped images of CO scattering off argon into J = 9, probed via the A state. The re- sults are compared to theoretical calculations of these moments, and good agreement is seen in many features of the distributions, although the magni- tude of the alignment for the most backscattered product is underestimated as is generally found to be the case experimentally. The approach may have other potential applications to photodissociation and reactive scattering.

Preliminary version – January 29, 2016 H e -CO SCATTERING FOR PURE ROTATIONAL TRANSI- 7 TIONS

A joint theoretical and experimental study of state-to-state rotationally in- elastic polarization dependent differential cross sections (PDDCSs) for CO ( v = 0 , j = 0 , 1 , 2 ) molecules colliding with helium is reported for colli- sion energies of 513 and 840 cm- 1 . In a crossed molecular beam experi- ment, velocity map imaging (VMI) with state-selective detection by ( 2 + 1 ) and ( 1 + 1 0 ) resonance enhanced multiphoton ionization (REMPI) is used to probe rotational excitation of CO due to scattering. By taking account of the known fractions of the j = 0, 1, and 2 states of CO in the rotationally cold molecular beam (T 3 K), close-coupling theory based on high- rot ⇡ quality ab initio potential energy surfaces for the CO-He interaction is used to simulate the differential cross sections for the mixed initial states. With polarization-sensitive 1 + 1 0 REMPI detection and a direct analysis proce- dure described by Suits et al. (J. Phys, Chem. A 2015, 119, 5925), alignment moments are extracted from the images and the latter are compared with images simulated by theory using the calculated DCS and alignment mo- ments. In general, good agreement of theory with the experimental results is found, indicating the reliability of the experiment in reproducing state- to-state differential and polarization-dependent differential cross sections

Lei Song, Gerrit C. Groenenboom, Ad van der Avoird, Chandan Kumar Bishwakarma, Gautam Sarma, David H. Parker, Arthur G. Suits Inelastic Scattering of CO with He: polarization dependent differential state-to- state cross sections, J. Phys. Chem. A, 119, 12526 (2015)

Preliminary version – January 29, 2016 96 he-co scattering for pure rotational transitions

7.1 introduction

Rotational excitation of diatomic molecules by collisions with atoms has been studied in intricate detail over the past few years, particularly for col- lisions of argon with the molecule NO [29, 165, 190], where state-to-state measurements of collision induced orientation and alignment of the NO final rotational angular momentum as a function of scattering angle has been determined. Recently, these measurements on NO-Ar have been ex- tended to study the effects of preorientation of the NO molecular axis be- fore collision. [191] Because these are nearly "perfect" experiments where all reactant and product variables are controlled, they form a suitable chal- lenge for ab initio quantum theory. In practice, the systems studied so far are simple enough that theory at present has proven to be more accurate than experiment, even for the open-shell NO molecule. In this case, imper- fections in experiment, noted as deviations from theory, can then be iden- tified, accounted for, and/or improved. This bootstrap process is necessary to prepare for experimental studies of scattering processes involving more than 3 - 4 atoms, which offer significant challenges for a fully quantum theoretical treatment. In this direction, encouraging agreement has already been found between advanced experiment and semiclassical or reduced di- mensional quantum theory for the systems OH-NO, [192]H2 O-H2 ,[193] NH3 -H2 [194] and CH3 -N2 ,[195] for example. In this chapter we describe a joint experimental and theoretical study at a similar level of detail as the recent studies on NO-He/Ar, but now for collisions of He with the closed-shell molecule CO. There is a long history of CO-He inelastic scattering studies, as outlined later in this section. An experimental study with angular momentum polarization-sensitive state- selective detection of CO is, however, far more challenging than for NO because the excited electronic states that can be used as a resonance step for CO lie deep in the vacuum ultraviolet (VUV). Although CO inelastic scat- tering is less challenging for ab initio full quantum theory than NO, predic- tion of state-to-state PDDCS for CO-He elastic and inelastic scattering, as de- scribed here, has not been reported. Two ab initio potential energy surfaces are compared in this chapter, one from the more recent study of Peterson et al. [196] named CBS+corr, and the other a symmetry-adapted perturbation theory (SAPT) potential from 1997 taken from Heijmen et al. [123]. Both potentials have been proven to be highly accurate in simulating the colli- sion energy dependence of integral cross sections for low-energy scattering resonances of CO+He [197]. In our experiment we utilize two different laser ionization detection schemes with REMPI of CO in the ( 2 + 1 ) and ( 1 + 1 0 ) configurations, couple this with VMI detection [172], and measure (PD)DCSs describing collisions of CO ( j = 0 , 1 , 2 ) with helium at collision energies 513 and 840 cm- 1 . Here, j refers to an initial rotational state of CO whereas j 0 refers to a nascent state populated by collision. ( 2 + 1 ) REMPI detection of CO via the E-state [198] is highly efficient, convenient, and for our conditions, polariza-

Preliminary version – January 29, 2016 7.1 introduction 97 tion insensitive, and thus useful for extracting the pure DCS. We have shown in a recent paper [199] that the ( 1 + 1 0 ) detection method for CO-Ar col- lisions is polarization sensitive and produces high-resolution images. Tak- ing advantage of the sharp and polarization-sensitive images, we demon- strated a simple scheme for extracting the polarization moments describing collision-induced alignment for the CO-Ar (final CO state j 0 = 9) system directly from the measured VMI images. We test this method here for CO- He collisions for a range of final rotational ( j 0 ) states of CO. CO-He has served as a benchmark for energy-transfer dynamics, and also as a model for the CO-H and CO-H2 collision systems. Earlier work has been done to identify the key aspects of the potential energy surface. Early molecular beam experiments on CO-He were reported by Butz et al. in 1971 [200], and Nerf et al. in 1975 presented the first experimental evi- dence of sensitivity to anisotropy of the CO-He potential [201]. Antonova et al. [158] measured relative state-to-state integral cross sections for rota- tional excitation of CO in collision with helium at collision energies of 583 and 720 cm-1. They compared their experimental data with results from two high-quality potential energy surfaces, with the ab initio SAPT surface of Heijmen et al., [123] and with the XC(fit) surface of LeRoy et al. [202] and found better agreement for the SAPT surface [123]. Lorenz et al. mea- sured state-to-state differential cross sections for rotational excitation of CO in collision with Ne at a collision energy near 511 cm-1, and they have com- pared their experimental data with results based on the CCSD(T) surface of McBane and Cybulski and the surface of Moszynski et al. [203, 204, 205]. In 2004, Smith et al. [206] reported state-to-state rotational transition rate constants for CO-He obtained from infrared double resonance measure- ments and scattering calculations based on the SAPT potential of Heijmen et al. [123]. Their work provided a nice demonstration of the quality of the SAPT potential. Even better ab initio CO-He potentials were calculated more recently and tested against experimental data in scattering calculations by Peterson and McBane [196]. In this article comparison of the differential cross sections from our measurement with the calculations on the SAPT po- tential [123] and the best potential of Peterson et al. [196] demonstrate the accuracy of these potentials. Carbon monoxide is of high astrochemical relevance as the second most abundant molecule in the universe after molecular hydrogen. Radiation emitted by CO is commonly used as a tracer to extract astrophysical param- eters, as rotational transitions of the asymmetric molecule CO provide the best means to determine the physical conditions of the interstellar medium. Many models of the processes occurring in such environments are based on the assumption of LTE. However, when hot stars, low-density chromo- spheres, and coronae of solar-type stars are investigated, the LTE approx- imation breaks down and non-LTE modeling is required [26, 32, 100]. In this case, successful modeling of the observed CO spectral lines requires the knowledge of accurate collision rate coefficients of CO with its domi- nant collision partners: He, H2, H and electrons [32, 33]. Rate coefficients

Preliminary version – January 29, 2016 98 he-co scattering for pure rotational transitions

Figure 7.1: Left: Newton diagram with velocity vectors for He and CO molecular beams crossed at an angle of 85, center of mass (long dash) and relative velocity vectors (short dash). The circle represents elastic scattering (j0 = 0 final state), corresponding to a collision energy of 840 40 cm-1. An ± image for the CO j0 = 7 final state is inset in the circle, and a color bar is shown to decode the signal level. Positions of forward, sideways, and backward scattering are indicated. Right: Energy level diagram with CO rotational energy spacings and energetics for excitation of j = 0 to -1 -1 j0 = 7 at 840 cm collision energy. Measurements at 513 cm collision energy where j0 = 15 is the highest energy final state are also reported here.

are commonly obtained from theoretical calculations, the quality of which must be tested by experiment. DCS measured in crossed-molecular beam experiments have been shown to provide the most stringent test for state- of-the-art theoretical calculations [207, 208]. To introduce the energetics of the CO-He imaging experiment, a Newton diagram and a typical scattering image for the j0 = 7 final state of CO is shown in Fig. 7.1. The images are two-dimensional "crushed" projections of the three-dimensional velocity spheres, having the form of a ring displaced from the center of the image (zero velocity position) by the velocity of CO relative to the center-of-mass velocity. The variation of the intensity around the ring provides the nature of scattering, e.g., forward or backward scatter- ing, which in turn reveals the nature of the interaction between the colliding species.

7.2 experimental method

The data were collected in a crossed molecular beam machine with a vari- able beam crossing angle and REMPI ionization of the scattered products with velocity map imaging detection as shown schematically in Fig. 7.2.

Preliminary version – January 29, 2016 7.2 experimental method 99

Figure 7.2: Schematic of the crossed-beam VMI setup showing two molecular beams formed by Nijmegen pulsed valves(NPVs), crossing at the center of the imaging ion optics. The angle between the two molecular beams can be changed by rotating one beam about the scattering center in the plane of collision. The scattered product is excited by VUV radiation res- onant with a CO X (v = 0, j) A (v = 0, j ) transition. VUV is generated ! 0 from four-wave difference frequency mixing in a cell containing Xe us- ing 249 nm and 650 nm radiation from two tunable dye laser systems. CO A-state molecules are then ionized by the remaining dye laser ra- diation. The VUV polarization is set in or perpendicular to the collision plane by rotation of the 650 nm laser polarization. The nascent CO+ im- age is projected by velocity map imaging ion optics and mass selected by time-of-flight onto a two dimensional (2D) imaging detector, then recorded by a CCD camera.

The basic apparatus was described in Ref. [209], details specific to the 1 + 10 REMPI measurements are given here. We used two skimmed supersonic beams, one of neat helium produced from a Nijmegen pulsed valve [183] and the other of 2% CO seeded in argon produced from a second Nijmegen pulsed valve. Backing pressure for the expansion was 3 bar for both molecular beams. The pulse duration of both valves at the crossing point is about 50 µs. The source chamber that houses the secondary beam (He) was differentially pumped and the molecular beam passes through a skimmer of aperture diameter 2.5 mm positioned 3 cm away from the nozzle. The primary molecular beam source was mounted in the differentially pumped rotatable chamber positioned 7 cm from the skimmer of aperture diameter 2.5 mm. All vacuum chambers were pumped by turbo molecular pumps. The laser and molecular beams

Preliminary version – January 29, 2016 100 he-co scattering for pure rotational transitions

were coplanar and parallel to the imaging detector face and were optimally aligned to ensure the best signal. Rotational populations of the parent CO/Ar beam were obtained from the (2 + 1) REMPI spectrum of the E1⇧-X1⌃+ transition of CO [198, 210]. Similar to the results of Antonova et al. [158], the population of j = 0 was about 65% and of j = 1 about 33%, corresponding to a rotational temper- ature of 3 K, and a small fraction of higher states with a close to ther- mal distribution are present. By comparing the PGOPHER [211] simulated spectrum with the experimental one, we can determine the most favorable REMPI line positions for detection of higher rotational states, which is the prerequisite information for state-to-state inelastic scattering studies. Under collision conditions, depletion of the CO j = 0 ground state was measured to be less than 10%, which should ensure a single inelastic col- lision [207, 208]. The scattered product was ionized using either (2 + 1) or (1 + 10) REMPI, and the ionic velocity sphere extracted by a velocity map imaging electrostatic lens [172]. After 85 cm free flight along the time-of- flight tube, the 3-D ion sphere was crushed onto 2-D detector, which is read by a CCD camera. Application of time gating at the detector enables selec- tion of the required mass and filters out background ions of other mass. The laser and both molecular beams were operated at 10 Hz. Typically, 20 - 50 000 laser shots were averaged for each image, under interleaved con- ditions with and without the He beam in time overlap with the CO beam. The second condition was used for background subtraction; difference im- ages are taken when the final-state population in the parent is significant and overlaps the scattered-in final-state image. This subtraction method is not ideal; it leads to problems especially in the forward scattering region, as discussed later in the text. The laser frequency was fixed during each image collection, rather than scanning the wavelength over the full width of the Doppler profile of the REMPI transition (0.07 cm-1 for inelastic scattering of CO from He), because the laser bandwidth with both (2 + 1) and (1 + 10) REMPI is 0.1 cm-1. For (2 + 1) REMPI the 2 mJ energy per pulse power ⇡ ⇡ broadening of the transitions was also sufficient to make Doppler scanning unnecessary. For (2 + 1) REMPI on the E1⇧-X1⌃+ transition, well-resolved and strong S-branch transitions around 215 nm were used for data collection. The laser wavelengths were generated by frequency tripling the output of a tunable pulsed dye laser (Fine Adjustment) operating with DCM dye using two BBO crystals. The dye laser was pumped by a Nd:YAG laser (Continuum Surelite) operating at 532 nm and with 10 Hz repetition rate. UV laser en- ergies were typically 2 mJ per pulse in pulses of duration 4 - 6 ns. The ⇡ probe beam was focused with a 20 cm lens onto the intersection region of the molecular beams. For (2 + 1) REMPI of CO the E1⇧ state was used as the resonant state instead of the commonly used B1⌃+ state where the low-j levels are over- lapped. Detection on the S-branch of the E1⇧ state should in principle be polarization sensitive but repeated measurements (under the same exper-

Preliminary version – January 29, 2016 7.2 experimental method 101

imental conditions as (1 + 10) REMPI) did not reveal any sensitivity to the laser polarization. Due to the low ionization efficiency and small probe vol- ume it was necessary to use 2 mJ laser pulse energy, which is likely to be too high to retain polarization sensitivity. Antonova et al. [158] used 100 µJ per pulse, for example, and estimated that the maximum polarization effect in the integral cross section for CO-He scattering with E-state detection was less than 10%. Experiments employing (1 + 10) REMPI of CO were described briefly in Ref. [199]. Radiation at 249 nm tuned resonant with the two-photon ⇡ 5p6 5p56p[1/2] transition at 80119 cm-1 in Xe was combined in a cell ! containing 30 mbar Xe with tunable radiation around 650 nm to generate tunable 154 nm light by difference frequency generation. This VUV light is resonant with the CO A1⇧-X1⌃+(v = 0) transition. For simplicity we label the ionization process as (1 + 10) REMPI. Although the 154 nm light drives the resonant excitation, the process driving ionization of the A-state is not yet identified. Analysis of the size and shape of the parent CO+ beam spot shows that ionization is close to threshold (i.e., ion recoil is small), indicat- ing that ionization occurs by the combination of 248 nm + 650 nm photons. A photoelectron study is presently underway to better identify the ioniza- tion process. The collision energy was 840 cm-1 for the geometry used in the (1 + 10) REMPI detection experiments. This is higher than in the (2 + 1) REMPI detection experiments due to a different mounting of the pulsed valves, which led to a higher equilibrium temperature after warming up. The collision energy was estimated for both REMPI detection methods by combining data from both the initial beam velocities and from the size of the images as a function of the final rotational state j0. The PGOPHER program [211] was used to fit the CO A-X (1 + 10) REMPI transition, which shows P, Q, and R branches of similar intensities, and (as with the (2 + 1) REMPI spectrum) a rotational temperature of 3K. This ⇡ similarity indicates that the ionization step in the (1 + 10) process does not significantly perturb the spectrum. The measured images for the same j0 are different when a P, R or Q branch transition is used. For the same transition branch the images differ with H (VUV laser polarization direction in the scattering plane) versus V polarization where the VUV polarization is directed perpendicular to the scattering plane, along the time-of-flight axis, as shown later. As described in Ref. [199], all H and V polarization images were obtained for Q-branch transitions. The raw H and V images were first converted from density to flux using the IMSIMM program [212], although for the polarization analysis method this was not essential. The asymmetry in the raw images disappears to a large extent upon density to flux correction. Because the effective angu- lar resolution varies widely with scattering angle, the images still do not look fully symmetric after correction. The density to flux correction is a secondary effect in that at any scattering angle the alignment moments are obtained by differences divided by the DCS. For each j0 final state, two angular distributions were extracted from the H and V images by integrat-

Preliminary version – January 29, 2016 102 he-co scattering for pure rotational transitions

Figure 7.3: Definition of the laboratory frame coordinate system for the He-CO scattering process. See text and Fig. 2 of Ref. [199] for details.

ing over the outside annulus of the image, or over a stripe through the middle of the image along the initial relative velocity vector, as described in Ref. [199]. This yields four polarization angular distributions: HIP, hori- zontal in-plane; VIP, vertical in-plane; HOOP, horizontal out-of-plane; and VOOP, vertical out-of-plane polarization. An important difference between the present CO-He experiment and the CO-Ar experiment described in Ref. [199] is the direction of laser propaga- tion, klaser. Using the standard reference frame as sketched in Fig. 7.3, the z axis is parallel to the initial relative velocity vector k, xz is the scattering plane containing k and the product relative velocity vector k0, and the y axis is parallel to k k . Using linearly polarized light, the electric field of the ⇥ 0 laser can be directed either along z (labeled horizontal or H polarization) or along the ion time-of-flight axis y (vertical, V polarization) when klaser is directed along the x axis. In the laboratory frame, klaser is the same as in the CO-Ar experiment, fixed at an angle of 45 with the CO initial velocity vector vCO. In the CO-Ar case, this angle and the angle of 85 between the CO and Ar beams, cf. Fig. 7.1, implied that the laser beam was perpendic- ular to the initial relative velocity vector k. Also in the CO-He experiments the CO and He beams cross at an angle of 85 and for the collision energy of 840 cm-1, where the CO final-state detection is polarization sensitive, the velocities of CO and He are such that the angle between klaser and k is 120. The implications of this different geometry are discussed later in the text.

Preliminary version – January 29, 2016 7.3 theoretical method 103

5.5 6.0 6.5 7.0 7.5 8.0 8.5

(a0)

Figure 7.4: Comparison of the first four coefficients v(R) in the Legendre expan- sions of the CBS+corr potential [196] and the SAPT potential [123].

7.3 theoretical method

Integral, differential and polarization-dependent differential state-to-state cross sections for CO-He collisions were computed using the quantum CC method [107]. The most accurate potential from the work of Peterson et al. [196], CBS+corr, was used; the SAPT potential was taken from Heijmen et al. [123]. Both potentials were calculated as full 3D CO-He potentials; we used the 2D potentials obtained by averaging them over the ground-state vibrational wave function of CO. The CO molecule was then treated as a -1 rigid rotor, with the experimental rotational constant B0 = 1.9225 cm . The 2D CO-He potential is a function of the Jacobi coordinates R, the length of the vector R pointing from the CO center of mass to the He nucleus, and , the angle between R and the CO bond axis ( = 0 for linear CO-He). The coupled channel equations were solved with the renormalized Nu- merov propagator, with R ranging from 3 to 50 a0 with a step size of 0.06 a0. The potential was expanded in normalized Legendre polynomials P(cos ) with running from 0 to 12. The R-dependent expansion coefficients were computed by 13-point Gauss-Legendre quadrature in the angle . The chan- nel basis consisted of functions for CO with a maximum value of j = 30; total angular momenta up to Jtot = 90 were included. With these parameters all of the calculated inelastic cross sections are converged to better than 1%. Both integral and differential cross sections were computed for transitions from initial states with j = 0, 1, 2 to all final states up to j0 = 15. The cross sections calculated for the different initial CO (j) states were weighted by the populations of these states in the mixture to compare the calculated val-

Preliminary version – January 29, 2016 104 he-co scattering for pure rotational transitions

ues with the experimental data. Because of the spread of about 40 cm-1 ± in the experimental collision energies of 513 and 840 cm-1, we computed the (PD)DCSs at 473, 513, and 553 cm-1 and at 800, 840, and 880 cm-1 and averaged the results with a weight ratio of 25 : 50 : 25 %.

7.4 results and discussion

7.4.1 Theoretical results

The potentials of Heijmen et al. [123] and Peterson et al. [196] used in the scattering calculations are compared in Fig. 7.4 by showing the main co- efficients in the Legendre expansion of these potentials as functions of R. The first three anisotropic terms with = 1, 2, 3 are even larger than the isotropic term with = 0; the term with = 2 is dominant. The coeffi- cients of the SAPT potential are slightly smaller than those of the CBS+corr potential [123]. These potentials are also compared in Fig. 7.5 in calculations of the differ- ential cross sections from the initial state with j = 0 to the final states j0 = 1, 4, 8, and 12 at a collision energy of 513 cm-1. The similarity of the result- ing DCSs illustrates that the already 18 year old SAPT potential is of nearly the same high quality as the more recent CBS+corr potential, a fact was also mentioned in Ref. [196]. The DCSs show the expected trend that the cross section decreases and that more backscattering occurs as the value of j0 increases. Although the peaks in the DCS for the CBS+corr potential are slightly shifted to smaller scattering angles (corresponding to a higher collision energy or a less attractive PES), the shifts are quite small and, as the reader will see below, the differences between the DCSs from the two potentials are smaller than the deviations from the experimental data. Figure 7.6 presents the elastic and inelastic integral cross sections (ICSs) from the initial states with j = 0, 1, 2 to the final states with j0 up to 15. Elas- tic (j = j0) integral cross sections are typically 1 - 2 orders of magnitude larger than the inelastic cross sections. The inelastic cross sections obviously depend strongly on j, and in general, they decrease for larger initial j val- ues because of the larger energy gap between the higher rotational states. The fact that the cross sections are largest for j = 2 follows from the dom- inance of the term with = 2 in the Legendre expansion of the potential, Fig. 7.4. As mentioned in the Experimental section, the present experiment does not achieve pure initial-state selection. The parent CO beam contains a mix- ture of j = 0, 1, and 2 states, which due to their different internal energy and quantum numbers show widely varying scattering behavior. This is il- lustrated in Fig. 7.7, where DCSs are shown for the three initial states j = 0, 1, 2 excited to the same final state j0 = 4. We took the population ratio of these states as 65 : 33 : 2 %, respectively. As expected, the initial-state averaged DCS (lower panel) is mainly determined by the 65% j = 0 contri- bution but note that the scattering cross section increases with increasing j

Preliminary version – January 29, 2016 7.4 results and discussion 105

Figure 7.5: Comparison of state-to-state differential cross sections from j = 0 to j0 = 1, 4, 8, 12 calculated on the CBS+corr and SAPT potentials, at a collision energy of 513 cm-1.

Preliminary version – January 29, 2016 106 he-co scattering for pure rotational transitions

Figure 7.6: Integral cross sections for elastic and inelastic scattering from the ini- tial j = 0, 1, 2 states to the j0 = 0 - 15 final states at collision energy 513 cm-1.

(decreasing j) and that the outer maximum in the DCS for j = 0 j = 4 ! 0 is much weaker for j = 1 j = 4 excitation. ! 0

7.4.2 Vector Correlations

Investigations of vector correlations in inelastic scattering processes reveal valuable information on the underlying dynamics [180]. The angular mo- mentum distributions in inelastic scattering are generally presented in terms of renormalized PDDCSs, but different expressions for these PDDCSs have been used in previous studies. Brouard et al. [180] listed the relationship between the polarization moments and alignment parameters used in var- ious papers. We present our alignment results in terms of the Hertel-Stoll (k) renormalized Aq (✓) polarization parameters, which are the irreducible components of the scattering density matrix related to the classical prob- ability density function. They are defined in the scattering frame with the (0) z-axis along the initial relative velocity vector k. A0 (✓) is directly related to the conventional DCS, which reveals the vector pair correlation between the initial and final relative velocity vectors k and k0, and ✓ is the angle between k and k0 in the collision plane, as shown in Fig. 7.3. Polarization- (2) (2) dependent differential cross sections, represented by the A0 (✓), A1+(✓), (2) and A2+(✓) polarization moments, describe the triple correlation between k, k0, and the final rotational angular momentum j0. In the case of inelastic

Preliminary version – January 29, 2016 7.4 results and discussion 107

Figure 7.7: Comparison of state-to-state differential cross sections from j = 0, 1, -1 and 2 to j0 = 4 at collision energy 513 cm .

Preliminary version – January 29, 2016 108 he-co scattering for pure rotational transitions

(2) atom-molecule scattering, the A0 (✓) moment for the j0, k correlation is re- lated to the population of the j0 substates with different mj0. It ranges from -1 when j0 is perpendicular to k, to +2 when j0 is parallel to k. Its value is sometimes expressed as the degree of "frisbee" (j k, k ) versus "propeller" 0? 0 (j0 k, k0) alignment of the final angular momentum j0 with respect to the k (2) (2) k, k0 scattering plane. The A1+(✓) and A2+(✓) moments contain the off- diagonal elements of the density matrix; they range from -1 to +1.[165] Introduction of the linearly polarized laser beam propagation along klaser in Fig. 7.3 breaks the cylindrical symmetry along k and allows probing of the propensities for alignment of j0 relative to the X, Y, Z axes of the colli- sion frame [199], where k lies along Zcoll. In the laboratory frame, Fig. 7.3, the present CO-He collision study differs from the Ar-CO collision study described in Ref. [199]. In the Ar-CO experiment the laser beam crossed k at 90, i.e., it was parallel to Xlab in Fig. 7.3. Due to the different relative velocity, this was not possible for CO-He and the laser now propagates at an angle of 120 with k and thus with Zlab. In the lab-to-collision frame transformation described in Ref. [199], the frame rotation angle ⇥ becomes ⇥ = 120 instead of 90, as is discussed in more detail later in the text. Algebraic combination of the four experimental HIP, HOOP, VIP, and VOOP distributions yields the polarization moments for the angular distri- (2) (2) (2) (0) butions [199]: A0 (✓), A1+(✓) and A2+(✓), as well as A0 (✓), which is the DCS.

7.4.3 Experimental results

Angular distributions extracted from velocity map images for collision in- duced alignment from rotational energy-transfer processes can be expressed in the form

d k k I(✓)=C (✓) 1 + A( )(✓)F( )(, ⇥, ) ,(7.1) d⌦ 2 q q 3 Xkq 4 5 d where d⌦ (✓) is the DCS, C is a constant factor that includes the effect of the density to flux correction, and the quantity in square brackets describes the (k) renormalized alignment distribution. The function Fq (, ⇥, ) contains all information on the experimental variables needed. [180, 199] In the case of (2 + 1) REMPI detection, described first in this section, the experimental (k) conditions are such that the Fq (, ⇥, ) values in Eq. 7.1 are effectively zero (i.e., there is no alignment sensitivity in the experiment) and the DCS can be determined directly from the density-flux corrected image and compared with theory. Next, the angular distributions determined using polarization sensitive (1 + 10) REMPI detection are described.

Preliminary version – January 29, 2016 7.4 results and discussion 109

Figure 7.8: Raw experimental two-dimensional velocity map images of the j0 = 3 - 11 final states using (2 + 1) REMPI detection at collision energy 513 40 cm-1, on the lefthand side. The direction of the initial rel- ± ative velocity vector is shown as a white arrow in the image for the j = 3 state. The color code is shown in Fig. 7.1. Images simulated us- ing the IMSIMM program are shown (middle column) for comparison with the experimental image. On the right-hand side experimental DCSs are compared with theoretical predictions for each final stat, where the experimental curve is scaled by a constant factor to match theory.

Preliminary version – January 29, 2016 110 he-co scattering for pure rotational transitions

(k) Table 7.1: Frame transformation angles and Fq (, ⇥, ) factors

HIP HOOP VIP VOOP (, ⇥, ) (⇡, 2⇡/3, 0) (⇡, 2⇡/3, ⇡/2)(⇡/2, 2⇡/3, 0)(⇡/2, 2⇡/3, ⇡/2) (2) F0 0.625 0.625 -0.5 -0.5 (2) F1+ 0.375 -0.375 0.375 0.375 (2) F2+ 0.217 -0.217 -0.87 0.87

7.4.3.1 Differential cross sections at 513 cm-1 collision energy using (2 + 1) REMPI detection

Raw images taken at 513 cm-1 collision energy with (2 + 1) REMPI detec- tion are shown on the left-hand side of Fig. 7.8 for final states in the range j0 = 3 - 11. Simulated images from the IMSIMM program which accounts for the divergence and velocity spread of the molecular beams and their rel- ative timing, as well as the laser ionization timing, volume, and propagation direction, are shown in the middle column. All final states from j0 = 3 to 11 were measured except for j0 = 5, in which case the S(5) and R(11) branches are fully overlapped. The S(6) transition is also overlapped with R(15), but the latter final state is very close to the energetic limit (see Fig. 7.1); its pop- ulation is thus negligible. As mentioned previously, detection with (2 + 1) REMPI under our experimental conditions showed no measurable sensitiv- ity to the direction of the laser polarization. The experimental DCSs from the IMSIMM program are shown on the right-hand side of Fig. 7.8 and compared with theoretical predictions.

7.4.3.2 Polarization dependent differential cross sections at 840 cm-1 collision energy In previous work on collision-induced alignment in NO-Ar collisions [29, 180], reasonable assumptions were made to extract the DCS and alignment functions from the data. The DCS was extracted from the image data I(✓) (k) by assuming the Aq (✓) values from the kinematic apse model [29, 178] or from a full quantum mechanical (QM) scattering calculation. For determin- (k) ing the Aq (✓) values, the DCS was taken from the QM calculations. Note that three polarization parameters can be predicted from the kinematic apse (2) model and all three are measurable. The A1+(✓) parameter, however, is not accessible when klaser is perpendicular to the initial velocity vector k, as in the CO-Ar experiment described in Ref. [199]. The frame transformation an- (k) gles and functions Fq (, ⇥, ) for the geometry in the present experiment are given in Table. 7.1. (k) Following the procedures outlined in Ref. [199], the Fq (, ⇥, ) factors are combined with the polarization moments to yield the angular distribu- tion for each polarization geometry as

Preliminary version – January 29, 2016 7.4 results and discussion 111

d 2 2 2 I (✓)= (✓)C 1 + 0.625A( )(✓)+0.375A( )(✓)+0.217A( )(✓) HIP d⌦ 0 1+ 2+ h i d 2 2 2 I (✓)= (✓)C 1 + 0.625A( )(✓)-0.375A( )(✓)-0.217A( )(✓) HOOP d⌦ 0 1+ 2+ h i (7.2) d 2 2 2 I (✓)= (✓)C 1 - 0.5A( )(✓)+0.375A( )(✓)-0.87A( )(✓) VIP d⌦ 0 1+ 2+ h i d 2 2 2 I (✓)= (✓)C 1 - 0.5A( )(✓)+0.375A( )(✓)+0.87A( )(✓) . VOOP d⌦ 0 1+ 2+ h i

In the setup of Ref. [199] with ⇥ = 90 the HIP and HOOP distribution must be identical, so that the position and scaling of the rectangular re- gion used to obtain the HOOP distribution could be adjusted to match the HIP distribution. The same position and scaling could then be used for the VOOP distribution. Here, we have ⇥ = 120 and it is apparent from the above equations that HIP and HOOP are not identical, so that this scaling method used is not applicable. A solution to this problem is to measure also at a different klaser direction; for example, at ⇥ = 0 where VIP=HOOP and HIP=VOOP. As this proved inconvenient, theory was again called on, and the relative cross sections from theory were used to calibrate the experimen- tal scaling factor relating the HIP and HOOP curves. The same scaling was then applied to the experimental VIP and VOOP distributions. In this ap- proach, the length of the integration region for HOOP and VOOP was fixed at the average radius of the annulus used for the HIP and VIP integrations. Linear combination of the expressions in Eq. 7.2 yields the differential cross section d 1 (✓)C I (✓)+I (✓)+I (✓)+I (✓) d⌦ ⇡ 4 HIP HOOP VIP VOOP h i (7.3) d 1 3 = (✓)C 1 + A(2)(✓)+ A(2)(✓) d⌦ 16 0 16 1+  where, considering the small values of their scaling coefficients and their tendency to cancel each other, the last two terms in Eq. 7.3 are ignored. This approximation was confirmed to be valid by comparing the exact and approximate forms of Eq. 7.2 using the polarization moments and DCS from theory.

Preliminary version – January 29, 2016 112 he-co scattering for pure rotational transitions

Other combinations yield the polarization parameters:

4 I (✓)+I (✓) A(2)(✓)= HIP HOOP - 2 0 5 d " d⌦ (✓)C #

2 4 I (✓)-I (✓) 2 A( )(✓)= HIP HOOP - 0.43A( )(✓) 1+ 3 d 2+ " d⌦ (✓)C # (7.4)

4 I (✓)+I (✓) 2 2 = HIP VIP - 2 - 0.125A( )(✓)+0.653A( )(✓) 3 d 0 2+ " d⌦ (✓)C #

(2) IVOOP(✓)-IVIP(✓) A2+(✓)= d , 1.74 d⌦ (✓)C

(2) where the two combinations for A1+(✓) from the experimental data were averaged. As shown in Ref. [199], (1 + 10) REMPI of CO allows measurement of polarization-dependent DCSs. Both V and H polarization images measured for j0 = 5 - 14 final states using (1 + 10) REMPI detection on Q-branches around 154 nm are shown in Fig.7.9. In these measurements the collision energy is 840 40 cm-1. For the j = 5 state a CO-beam only image was ± 0 subtracted from the CO+He image, for higher j0 states the parent beam signal is sufficiently far from the scattering ring that subtraction was not necessary. For j0 <5no Q-branch probe transitions free of overlap with other states could be found. Raw H and V images (without density-flux correction) are shown in the left panel of Fig. 7.9. HIP, HOOP, VIP, and VOOP angular distributions for each final state are shown in the middle panel of Fig. 7.9. HIP and VIP curves are obtained by integration around the annulus shown schematically as circles superimposed on the j0 = 9 V image, whereas the HOOP and VOOP curves are obtained from integration along the vertical stripes shown schematically on the j0 = 9 H image. As a test of the HIP-HOOP scaling procedure described above, the result- ing HIP, HOOP, VIP, and VOOP curves (middle column) are compared with predictions from theory shown in the right column of Fig.7.9. In all cases scaling factors did not vary by more than a factor of two from the ampli- tudes of the raw data curves. Although the noise is considerable, the shapes and areas of each experimental curve are in reasonable agreement with the- ory for all measured final states. Differential cross sections obtained by ap- plication of Eq. 7.3 are compared with theoretical predictions in Fig.7.10. Po- larization moments extracted from the experimental I(✓) curves are shown in Fig .7.11.

Preliminary version – January 29, 2016 7.4 results and discussion 113

Figure 7.9: Raw two-dimensional velocity map images of the j0 = 5 - 14 final states using (1 + 1 ) REMPI detection at collision energy 840 40 cm-1. The 0 ± direction of the initial relative velocity vector is shown as a white arrow in the image for the j0 = 5 state. The polarization of the (resonant) VUV photon is indicated by V for polarization perpendicular to the collision plane and H parallel to the collision plane. The sensitivity scale for each pair of images is adjusted to make the main features visible. The middle panel shows HIP, HOOP, VIP, and VOOP distributions extracted from these images and the right panel shows theoretical predictions corrected for the laser geometry. Each experimental curve was scaled by a constant factor to match the corresponding theory curve.

Preliminary version – January 29, 2016 114 he-co scattering for pure rotational transitions

Figure 7.10: DCSs for j0 = 5 - 14 computed with the use of Eq. 7.3 from the HIP, HOOP, VIP, and VOOP angular distributions extracted from the mea- sured images in comparison with calculated DCSs at collision energy 840 40 cm-1. ±

7.4.4 Discussion

CO is a closed shell molecule with a relatively small dipole moment (0.122 Debye), and He is a small atom with a low polarizability. Therefore, the interaction of CO with He is dominated by short-range interactions that requires small impact parameters. For this repulsive interaction system L- type rainbows, which have been shown to have remarkable polarization effects for NO-Kr collisions [213, 214] are not expected or seen in the theo- retical analysis. Instead, the kinetimatic apse model [178], which describes the system classically as a hard sphere colliding with a rigid ellipsoid, should yield an accurate prediction of collision-induced alignment effects, as shown in previous studies of NO-He, -Ne, and -Ar collisions. Indeed, the polarization moments predicted by theory here for the CO-He system are almost quantitatively similar to those predicted for the NO-He, Ne, and Ar systems [29, 180, 190, 191, 213, 214]. The magnitude of the energy-transfer cross sections between CO-He and CO-Ar, however, is quite different. For example, the integral cross section (ICS) for the j = 7 j = 0 excitation for 0 CO-Ar is 8 times larger than that for CO-He at 513 cm-1 collision energy, and similar differences are found for the NO-He versus NO-Ar systems. The smaller cross sections make collision-induced alignment studies with He as the collision partner much more difficult and up to the present only Meyer [178] has reported CIA studies of NO-He at the (high) collision en- ergy of 1185 cm-1. CO-He is thus not an easy system for experimental study, and as expected, this results in lower quality DCS and alignment data than CO-Ar. Regardless of the small cross section, the DCSs recovered when (2 + 1) REMPI detection is used (Fig. 7.8) and agree quite well with theoretical

Preliminary version – January 29, 2016 7.4 results and discussion 115

Figure 7.11: Renormalized alignment moments for j0 = 7 - 14 computed with the use of Eq. 7.4 from the HIP, HOOP, VIP, and VOOP angular distribu- tions extracted from the measured images (red) in comparison with calculated moments (black) at collision energy 840 40 cm-1. ±

Preliminary version – January 29, 2016 116 he-co scattering for pure rotational transitions

predictions. Note again that theory takes into account the mixture of ini- tial states in the experiment. With the relatively high-intensity laser pulses needed for the (2 + 1) REMPI detection, it is clearly possible to detect a signif- icant fraction of the scattered molecules without distortion due to complete saturation or space charge. In this aspect, the laser confocal diameter is roughly 30 micron for the 20 cm focal length lens used, but the 1 cm ⇡ confocal length of the laser beam exceeds the length of the 4 4 4 mm ⇡ ⇥ ⇥ the collision volume. Differential cross sections recovered from the full polarization study us- -1 ing 1 + 10 REMPI at 840 cm collision energy are also in reasonable agree- ment with theory, although for the lower j0 final states the second rainbow peak in the side to backscattering region is broader than predicted. The shape of the DCS, especially in this region, is sensitive to the value of the polar angle ⇥ and to the amount of j = 0 versus j = 1 in the initial beam. Considering that the experimental I(✓) curves match the shape of the corre- sponding theory curves (Fig. 7.9), when only HIP and HOOP are scaled us- ing theory, again confirms the general agreement of experiment and theory for the DCSs. Rescaling of the experimental curves was necessary because the total sensitivity for each image could not be held sufficiently constant when the laser polarization was rotated. The direct extraction method yields polarization parameters (Fig. 7.11) that are also qualitatively in agreement with theory predictions. As is typ- (2) (2) ical, the A0 (✓) parameter is most reliable whereas the A2+(✓) parameter (obtained only from the VIP-VOOP curves) is noisy and quite sensitive to (2) the scaling method. The A1+(✓) parameters show the expected shapes in general, although for the weak and quite and small j0 >10images substan- tial deviations are observed. All three experimental polarization parameters tend to reach their mini- mum values, in contrast to previous work on NO-rare gas studies, where (2) typically A0 (✓) reached only -0.4 to +0.5 instead of -1.0 at the backscat- tering angle. This difference could be due to the lack of nuclear spin in CO compared to NO, where any stray magnetic field could depolarize the nascent NO on the time scale of the scattering measurement. Another ef- fect, as suggested by Brouard and coworkers, [29] is depolarization due to secondary collisions. In our measurement, both the DCSs and polarization parameters show significant deviation from theory in the most forward scat- tering direction. We believe this is due mainly to elastic scattering following the inelastic event, as discussed next. For any study of inelastic (or reactive) scattering the effects of elastic scattering must be considered. Two features of elastic scattering are a much larger total cross section and strongly forward scattering. Figure 7.12 shows the elastic scattering DCS for the j0 = 7 state, a typical final product mea- sured in this study. To visualize both the magnitude of the cross section and the angular distribution, the DCS is weighted by sin ✓ in Fig. 7.12 where the area under the curve is then the integral cross section. The j = 7 - 7 ICS (similar to all of the other elastic ICSs) is 37 times larger than the j = 0 - 7

Preliminary version – January 29, 2016 7.4 results and discussion 117

Figure 7.12: Comparison of the DCS for elastic j = 7 - 7 and inelastic j = 0 - 7 CO-He collisions at collision energy 800 cm-1.

(2) Figure 7.13: Comparison of the renormalized aligment moments A0 (✓) and (2) A2+ (✓) for elastic j = 7 - 7 and inelastic j = 0 - 7 CO-He collisions at collision energy 800 cm-1.

Preliminary version – January 29, 2016 118 he-co scattering for pure rotational transitions

ICS, which is typical for all of the lower initial states, as seen in Fig. 7.6. The probability for an elastic collision after the j0 = 7 state is formed but before laser ionization takes place under the present condition of 5% inelastic ⇡ scattering is close to 100%. Also apparent in Fig. 7.12 is that although the elastic scattering is mostly forward, even at 20 degrees there can be a sig- nificant effect on the DCS. The probability in the side scattering region for j = 7 - 7 elastic scattering is greater than that of the 0 - 7 inelastic scatter- ing event. When a final state with a substantial fraction is detected in the parent beam (j = 1, 2, 3...) the measured image will thus also reflect elastic scattering and a simple beam-off background subtraction is not valid. An elastic scattering event shows collision-induced alignment effects that tend to be opposite to those of inelastic scattering polarization. This is il- (2) (2) lustrated for the A0 (✓) and A2+(✓) polarization parameters in Fig. 7.13 for j = 7 - 7 and j = 0 - 7 collisions. In the forward scattering region, elastic scattering can thus cause significant depolarization, and forward scattering is where the largest deviation in the polarization parameters between ex- periment and theory is seen in Fig. 7.11. Elastic scattering will not cause significant deflection of the nascent molecules. Once inelastically scattered, however, these molecules are no longer moving with their parent beam ve- locity and further collisional depolarization and elastic scattering with the carrier gas of the parent beam with much higher cross sections becomes possible. For the present CO-He system, collisional depolarization appears to mainly affect the forward scattering region whereas the backscattering region shows the maximum possible collision-induced alignment, within the large error margin of the experiment. For the CO-Ar system where the inelastic scatter- ing cross section is nearly an order of magnitude larger, collisional depolar- ization should play a more important role and this is indeed observed in the preliminary results of Ref. [199], where in the backscattering region the po- larization parameters do not reach their maximum expected values. Finally, for very dilute systems such as those with hexapole or Stark deceleration state selection the probability of an inelastic collision can be sufficiently low that depolarization by elastic scattering with the rare gas partner should not be important.

7.5 conclusion

The rotational excitation of CO molecules due to scattering with He atoms at collision energies of 513 and 840 cm-1 was studied in a crossed molec- ular beam setup and VMI with state-selective and polarization-sensitive de- tection through (1 + 10) and (2 + 1) REMPI. The incoming CO beam had a rotational temperature of T 3 K with j = 0, 1, and 2 states populated in rot ⇡ a ratio of 65 : 33 : 2 %. Differential cross sections and alignment parameters for a range of final j0 values of CO were extracted from the meaured images by a direct analysis procedure described by Suits et al. [199], but general- ized here to account for a different laser beam direction. The same DCSs

Preliminary version – January 29, 2016 7.5 conclusion 119 and alignment parameters, averaged over the initial-state distribution of CO, were obtained from quantum mechanical close-coupling calculations based on two accurate ab initio potentials for CO-He. Good agreement be- tween the experimental and theoretical results was found, in general, which illustrates the reliability of the experiment, as well as of the procedure used to extract state-to-state differential and polarization-dependent differential cross sections from the measured images.

Preliminary version – January 29, 2016 Preliminary version – January 29, 2016 H 2 -CO SCATTERING CALCULATIONS 8

Collisions of H2 and CO are important in astrochemistry. This chapter fo- cuses on the study of DCSs for pure rotational excitation of CO in collision with H2 . Experiments on this system have been done by our collaborator David H. Parker and his group. However, at the time of writing of this chapter the analysis of the experimental results was not finished and only theoretical results are presented. Cross sections are calculated with the fully quantum-mechanical close-coupling method using a four-dimensional po- tential energy surface for H2 -CO [J. Chem. Phys. 138, 084307,(2013)]. The scattering calculations are performed for a collision energy of 600 cm- 1 , corresponding to the experiment. Since CO has a small rotational constant, not only the j = 0 ground rotational state of CO is populated in the ex- periment, but also the j = 1 and j = 2 states. We examine the effect of the initial rotational state on the DCSs. We also investigate the difference between ortho-H2 and para-H2 collisions with CO. We conclude that the effect of initial excitation of CO is large and that the difference between ortho-H2 and para-H2 is small. Results for H2 -CO are similar to those of He-CO, which justifies the use of scaled rotationally inelastic He-CO rates to replace H2 -CO rates in astrochemical modeling.

Preliminary version – January 29, 2016 122 h2 -co scattering calculations

8.1 introduction

The H2 -CO complex is an important system for astrophysics. The hydro- gen molecule, the most abundant molecule in the universe, is hard to detect directly, while CO, the second most abundant molecule in the universe, is a good tracer of physical conditions, as illustrated in Sec. 1.1 of Chap. 1. Most information about hydrogen molecule in space is obtained indirectly from the measurement of the spectra of CO.[215, 216, 217] In contrast to the atom-diatom systems discussed in other chapters, both molecules can be excited or de-excited in H2 -CO collisions. The description of the complex requires six Jacobi coordinates, compared to only three for an atom-diatom system. A further complication is that H2 has two spin isomers, one called ortho hydrogen (ortho-H2 ) with the two proton spins parallel, and the other called para hydrogen (para-H2 ) with two proton spins antiparallel. [218] The parallel spins in ortho-H2 form a triple state, while antiparallel spins in para-H2 yield the singlet state. Para-H2 has an even rotational quantum

number j H 2 and ortho-H2 has odd j H 2 . The two spin isomers lead to dif- ferent ro-vibrational spectra of the H2 -CO complex [125] and also different dynamic behavior in H2 -CO collisions [26]. Jankowski and Szalewicz devoted considerable effort to constructing H2 - CO PESs.[125, 219, 220] In 1998, they built a four-dimensional ab initio PES for the H2 -CO complex with fixed intramolecular distances of H2 and CO us- ing interaction energies computed by symmetry-adapted perturbation the- ory. [219] With this potential para-H2 -CO spectral lines were predicted with an accuracy of 1 cm- 1 . In 2005 they reported another ab initio PES.[220] Here, they used the coupled-cluster super-molecular method with single, double, and noniterative triple excitations [CCSD(T)] to calculate interac- tion energies on a five-dimensional grid with only the CO distance fixed. The PES was constructed by averaging over the intermolecular vibration of H2 . The para-H2 -CO spectrum computed for this potential is accurate to 0 . 1 cm- 1 . These two potentials, however, are not sufficiently accurate for the calculation of the ortho-H2 -CO spectrum. Recently, Jankowski et al. built a new full six-dimensional H2 -CO surface using ab initio points from the single, double, complete triple, and noniterative quadruple excitations [CCSDT(Q)] method. [125] This potential is extremely accurate, it yields an ortho-H2 -CO spectrum for which most ro-vibrational lines agree with experiment within 0 . 01 cm- 1 . Integral cross sections for H2 -CO rotationally inelastic collisions have already been computed with these potentials[27, 221] In this chapter, we present the calculations of DCSs. We expect experimental results to be avail- able for comparison soon.

8.2 computational details

The cross sections are calculated with the fully quantum-mechanical CC method described in detail in Sec. 1.2 of Chap. 1. The expressions for the

Preliminary version – January 29, 2016 8.3 theoretical results 123

Figure 8.1: Jacobi coordinates for the H2-CO complex, as described in section 8.2. Figure adapted from Ref. [222].

calculation of the DCSs are given in the supplementary material of Ref [223]. Here, we present further computational details. The geometry of the H2-CO complex is described by the Jacobi vectors r1, r2, and R, where r1 connects the atoms O and C in CO, r2 connects the H atoms in H2, and R connects the enters of mass of CO and H2, as show in Fig. 8.1. The potential is expressed in the six Jacobi coordi- nates (R, r1, r2, ⇥1, ⇥2, ), with R = |R|, r1 = |r1|, r2 = |r2|, ⇥1 = \(R, r1), ⇥2 = \(R, r2), and is the dihedral angle. We treat H2 and CO as rigid rotators and we use the four-dimensional ab initio potential produced by Jankowski et al. [126], which was obtained by taking an average of a six- dimensional potential [125] over the ground-state vibrational wave func- tions of H2 and CO. [126] We employ the renormalized Numerov propa- gator to solve the coupled-channel equations. The R grid ranges from 4.5 to 40 a0 and the step size is 0.13 a0. The potential is expanded in spherical harmonics, where the highest order is 10 for CO and 4 for H2. The potential expansion coefficients are obtained by numerical integration, using 11 point Gauss-Legendre quadratures for ✓1 and ✓2, and 11 equally spaced points for . The channel basis set includes CO rotational quantum number j = 0 - 25 and H2 rotational quantum number jH2 = 0 - 4. The rotational constants of -1 -1 CO and H2 are 1.9225 cm and 59.3398 cm , respectively. Partial waves with total angular momentum J up to 90 are included to converge the ICSs within 1%. In the experiment the initial populations of the j = 0, 1, and 2 rotational states are estimated to be 65%, 33%, and 2%. Therefore, we report the DCSs for these three initial states, as well weighted average results.

8.3 theoretical results

In Fig. 8.2 we compare calculated DCSs for ortho- and para-H2 colliding with CO. Results are shown for CO rotational transitions j = 0 1, 4, 8, ! and 11 at a collision energy of 600 cm-1. At this collision energy, the ortho- H2-CO and para-H2-CO DCSs are very similar. Even though the differences between ortho- and para-H2 are small, we report 3 : 1 weighted averages in Figs. 8.3-8.5 in order to best approach the experimental conditions.

Preliminary version – January 29, 2016 124 h2 -co scattering calculations

8 0.5 ortho-H ortho-H 2 2 para-H 0.4 para-H 6 2 2

/sr) (a) j = 1 ← 0 /sr) (b) j = 4 ← 0 2 2 0.3

(A 4 (A

Ω Ω 0.2 /d /d

σ 2 σ d d 0.1

0 0 0 30 60 90 120 150 180 0 30 60 90 120 150 180 Angle θ (degrees) Angle θ (degrees)

ortho-H ortho-H 0.12 2 0.06 2 para-H para-H 2 2

/sr) 0.09 (c) j = 8 ← 0 /sr) (d) j = 11 ← 0 2 2 0.04 (A (A

Ω 0.06 Ω /d /d

σ σ 0.02

d 0.03 d

0 0 0 30 60 90 120 150 180 0 30 60 90 120 150 180 Angle θ (degrees) Angle θ (degrees)

Figure 8.2: Comparison of differential cross sections of ortho-H2-CO and para-H2- CO for transitions from j = 0 to j0=1, 4, 8, and 11 at collision energy E = 600 cm-1.

0.5 0.8

0.4 0.6 / sr) / sr) ← 2 ← 2 0.3 j = 4 0 j = 4 1 0.4 (A (A

Ω 0.2 Ω /d /d

σ σ 0.2 d 0.1 d

0 0 0 30 60 90 120 150 180 0 30 60 90 120 150 180 Angle θ (degrees) Angle θ (degrees)

8 0.5

6 0.4 65% j = 4 ← 0

/ sr) / sr) 33% j = 4 ← 1 ← 2 2 j = 4 2 0.3 4 2% j = 4 ← 2 (A (A

Ω Ω 0.2 /d /d

σ 2 σ d d 0.1

0 0 0 30 60 90 120 150 180 0 30 60 90 120 150 180 Angle θ (degrees) Angle θ (degrees)

Figure 8.3: Differential cross sections of H2-CO from initial j=0, 1, and 2 to final -1 j0 = 4 at collision energy E = 600 cm . The last panel shows the weighted average of the differential cross sections (DCSs)

Preliminary version – January 29, 2016 8.3 theoretical results 125

Figure 8.4: Integral cross sections for H2-CO for transitions from initial j = 0, 1, -1 and 2 to final j0 = 0 - 11 at collision energy of 600 cm . The ratio of ortho-H2 and para-H2 is 3 : 1.

4 0.6 0.4 ′ ′ ′

/sr) j = 3 /sr) j = 4 /sr) j = 5

2 2 0.4 2 2 0.2 0.2 DCS (A DCS (A DCS (A 0 0 0 0 45 90 135 180 0 45 90 135 180 0 45 90 135 180 θ (degrees) θ (degrees) θ (degrees) 0.15 0.2 ′ 0.2 ′ ′

/sr) j = 6 /sr) j = 7 /sr) j = 8

2 2 2 0.1

0.1 0.1 0.05 DCS (A DCS (A DCS (A 0 0 0 0 45 90 135 180 0 45 90 135 180 0 45 90 135 180 θ (degrees) θ (degrees) θ (degrees) 0.15 0.1 ′ ′ 0.06 ′

/sr) j = 9 /sr) j = 10 /sr) j = 11

2 0.1 2 2 0.04 0.05 0.05 0.02 DCS (A DCS (A DCS (A 0 0 0 0 45 90 135 180 0 45 90 135 180 0 45 90 135 180 θ (degrees) θ (degrees) θ (degrees)

Figure 8.5: The 65:33:2 weighted average of the H2-CO(j = 0, 1, 2) DCSs for exci- -1 tation to final j0 = 3 - 11 at a collision energy of 600 cm . The H2 ortho-para ratio is 3:1.

Preliminary version – January 29, 2016 126 h2 -co scattering calculations

In Fig. 8.3 we report the DCSs for excitation to the CO j = 4 rotational state from initial states with j = 0, 1, and 2. The angular dependence as well as the magnitude of the DCSs depends strongly on the initial state. The intensity of the peaks can differ by an order of magnitude. In particular, the backward scattering peak (✓ near 180 ) for the j = 0 4 transition ! is much stronger than the one for j = 1 4, while forward scattering is ! much stronger for j = 1 4. The weighted average DCS in Fig. 8.3 shows ! that even though the majority of the CO molecules is initially in the ground rotational state, the contribution from the j = 1 and j = 2 initial states is not negligible. The main theoretical results for the H2-CO system are presented in Figs. 8.4 and 8.5. Figure 8.4 shows the ICSs from initial j = 0, 1, and 2 to final j0 = 0 - 11. It reveals that elastic cross sections are an order of magnitude larger than inelastic cross sections. The ICSs decrease with increasing j con- sistent with the energy gap law. Furthermore, the transitions with j = 2 have the largest inelastic ICSs, just as in He-CO collisions. In Fig. 8.5, DCSs are shown for transitions to final states with j0 = 3 - 11. A weighted average over j = 0, 1, and 2 is taken as before and the contributions from ortho-H2-CO and para-H2 are included. The DCSs of H2- CO have more backward scattering for increasing final j0. This trend was also found for He-CO collision in Chap. 7.

8.4 conclusion

We presented calculated ICSs and DCSs for the H2-CO system at a collision energy of 600 cm-1. We consider initial CO rotational states with j = 0, 1, and 2, and final states with j0 = 0 - 11. We find that results depend strongly on the initial rotational state of CO, but that the differences between ortho- H2 and para-H2 are small. We also present results averaged over j = 0, 1, and 2 CO rotational states with populations 65 : 33 : 2 and contributions from ortho- and para-H2 in 3 : 1 ratio, to closely mimic experimental condi- tions. It is found that the results for H2-CO are similar to those for the He-CO. This validates the common approach in astrochemical applications to use scaled rotationally inelastic He-CO rates in modeling, rather than H2-CO rates which are much costly to compute. Our theoretical results are ready for comparison with experimental data. Experimental images have already been obtained by our collaborator Dave H. Parker and his group. The analyzed experimental data are expected to be available for comparison soon.

Preliminary version – January 29, 2016 BIBLIOGRAPHY

[1] “Universe–Wikipedia”. http://en.wikipedia.org/wiki/Universe.

[2] “Dark matter–Wikipedia”. http://en.wikipedia.org/wiki/Dark_ matter.

[3] Fraser, H.J., McCoustra, M.R.S., and Williams, D.A. “The molecular universe”. Astron. Geophys., vol. 43, pp. 2.10, 2002.

[4] Tielens, A.G.G.M. “The molecular universe”. Rev. Mod. Phys., vol. 85, pp. 1021, 2013.

[5] Gerin, M. The Molecular Universe, chap. 2, p. 35. Springer-Verlag Berlin Heidelberg, 2013.

[6] “Astrochemistry–Wikipedia”. http://en.wikipedia.org/wiki/ Astrochemistry.

[7] Dalgarno, A. “A Serendipitous Journey”. Annu. Rev. Astron. Astr., vol. 46, pp. 1, 2008.

[8] van Dishoeck, E.F. “Astrochemistry of dust, ice and gas: introduction and overview”. Faraday Discuss., vol. 168, pp. 9, 2014.

[9] Thi, W.F., Kamp, I., Woitke, P., van der Plas, G., Bertelsen, R., and Wiesenfeld, L. “Radiation thermo-chemical models of protoplane- tary discs - IV. Modelling CO ro-vibrational emission from Herbig Ae discs”. Astron. Astrophys., vol. 551, p. A49, 2013.

[10] “Spitzer Space Telescope–Wikipedia”. http://en.wikipedia.org/ wiki/Spitzer_Space_Telescope.

[11] Costen, M.L., Marinakis, S., and McKendrick, K.G. “Do vectors point the way to understanding energy transfer in molecular collisions?” Chem. Soc. Rev., vol. 37, pp. 732, 2008.

[12] Johnson, B.R. “The renormalized Numerov method applied to calcu- lating bound states of the coupled-channel Schrödinger equation”. J. Chem. Phys., vol. 69, p. 4678, 1978.

[13] Johnson, B.R. “The log-derivative and renormalized Numerov algo- rithms”. NRCC Proceedings, vol. 5, p. 86, 1979.

[14] Abramowitz, M. and Stegun, I.A., editors. Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables. No. 55 in National Bureau of Standards Applied Mathematics Series. U.S. Gov- ernment Printing Office, Washington, D.C., 1964.

127

Preliminary version – January 29, 2016 128 Bibliography

[15] Yang, B., Stancil, P.C., Balakrishnan, N., Forrey, R.C., and Bowman, J.M. “Quantum calculation of inelastic CO collisions with H. I. Rota- tional quenching of low-lying rotational levels”. Astrophys. J., vol. 771, p. 49, 2013.

[16] Balakrishnan, N., Yan, M., and Dalgarno, A. “Quantum-mechanical study of rotational and vibrational transitions in CO induced by H atoms”. Astrophys. J., vol. 568, pp. 443, 2002.

[17] Werner, H.J., Bauer, C., Rosmus, P., Keller, H.M., Stumpf, M., and Schinke, R. “The unimolecular dissociation of HCO: I. Oscillations of pure CO stretching resonance widths”. J. Chem. Phys., vol. 102, p. 3593, 1995.

[18] Keller, H.M., Floethmann, H., Dobbyn, A.J., Schinke, R., Werner, H.J., Bauer, C., and Rosmus, P. “The unimolecular dissociation of HCO: II. Comparison of calculated resonance energies and widths with high- resolution spectroscopic data”. J. Chem. Phys., vol. 105, p. 4983, 1996.

[19] Shepler, B.C., Yang, B.H., Kumar, T.J.D., Stancil, P.C., Bowman, J.M., Balakrishnan, N., Zhang, P., Bodo, E., and Dalgarno, A. “Low en- ergy H+CO scattering revisited - CO rotational excitation with new potential surfaces”. Astron. Astrophys., vol. 475, p. L15, 2007.

[20] Cecchi-Pestellini, C., Bodo, E., Balakrishnan, N., and Dalgarno, A. “Rotational and vibrational excitation of CO molecules by collisions with 4He atoms”. Astrophys. J., vol. 571, pp. 1015, 2002.

[21] Reid, J.P., Simpson, C.J.S.M., and Quiney, H.M. “A new He-CO inter- action energy surface with vibrational coordinate dependence. II. The vibrational deactivation of CO (v = 1) by inelastic collisions with 3He and 4He”. J. Chem. Phys., vol. 107, pp. 9929, 1997.

[22] Balakrishnan, N., Dalgarno, A., and Forrey, R.C. “Vibrational relax- ation of CO by collisions with 4He at ultracold temperatures”. J. Chem. Phys., vol. 113, pp. 621, 2000.

[23] Bodo, E., Gianturco, F., and Dalgarno, A. “Quenching of vibra- tionally excited CO (v = 2) molecules by ultra-cold collisions with 4He atoms”. Chem. Phys. Lett., vol. 353, p. 127, 2002.

[24] Yang, B., Stancil, P.C., Balakrishnan, N., and Forrey, R.C. “Close- coupling study of rotational energy transfer of CO (v = 2) by col- lisions with He atoms”. J. Chem. Phys., vol. 123, 134326, 2005.

[25] Krems, R.V. “Vibrational relaxation in CO+He collisions: Sensitivity to interaction potential and details of quantum calculations”. J. Chem. Phys., vol. 116, pp. 4525, 2002.

[26] Flower, D.R. “Rate coefficients for the rovibrational excitation of CO by H2 and He”. Mon. Not. R. Astron. Soc., vol. 425, pp. 1350, 2012.

Preliminary version – January 29, 2016 Bibliography 129

[27] Yang, B., Stancil, P.C., Balakrishnan, N., and Forrey, R.C. “Rotational quenching of CO due to H2 collisions”. Astrophys. J., vol. 718, p. 1062, 2010.

[28] Reid, J.P., Simpson, C.J.S.M., and Quiney, H.M. “The vibrational de- activation of CO (v = 1) by inelastic collisions with H2 and D2”. J. Chem. Phys., vol. 106, pp. 4931, 1997.

[29] Brouard, M., Chadwick, H., Eyles, C.J., Hornung, B., Nichols, B., Aoiz, F.J., Jambrina, P.G., Stolte, S., and de Miranda, M.P. “Rotational align- ment effects in NO(X) + Ar inelastic collisions: A theoretical study”. J. Chem. Phys., vol. 138, 104309, 2013.

[30] Schöier, F.L., van der Tak, F.F.S., van Dishoeck, E.F., and Black, J.H. “An atomic and molecular database for analysis of submillimetre line observations”. Astron. Astrophys., vol. 432, pp. 369, 2005.

[31] van der Tak, F.F.S., Black, J.H., Schoier, F.L., Jansen, D.J., and van Dishoeck, E.F. “A computer program for fast non-LTE analysis of interstellar line spectra - With diagnostic plots to interpret observed line intensity ratios”. Astron. Astrophys., vol. 468, p. 627, 2007.

[32] González-Alfonso, E., Wright, C.M., Cernicharo, J., Rosenthal, D., Boonman, A.M.S., and van Dishoeck, E.F. “CO and H2O vibrational emission toward orion peak 1 and peak 2”. Astron. Astrophys., vol. 386, p. 1074, 2002.

[33] Liszt, H.S. “Rotational excitation of CO in a cool, mixed atomic and molecular hydrogen gas”. Astron. Astrophys., vol. 458, p. 507, 2006.

[34] Bowman, J.M., Bittman, J.S., and Harding, L.B. “Ab initio calculations of electronic and vibrational energies of HCO and HOC”. J. Chem. Phys., vol. 85, p. 911, 1986.

[35] Bowman, J.M. and Gazdy, B. “A simple method to adjust potential energy surfaces: Application to HCO”. J. Chem. Phys., vol. 94, p. 816, 1991.

[36] Wang, D. and Bowman, J.M. “Complex L2 calculations of bound states and resonances of HCO and DCO”. Chem. Phys. Lett., vol. 235, p. 277, 1995.

[37] Tobiason, J.D., Dunlop, J.R., and Rohlfing, E.A. “The unimolecular dissociation of HCO: A spectroscopic study of resonance energies and widths”. J. Chem. Phys., vol. 103, p. 1448, 1995.

[38] Gray, S.K. “Wave packet dynamics of resonance decay: An itera- tive equation approach with application to HCO H+CO”. J. Chem. ! Phys., vol. 96, p. 6543, 1992.

Preliminary version – January 29, 2016 130 Bibliography

[39] Dixon, R.N. “A three-dimensional time-dependent wavepacket calcu- lation for bound and quasi-bound levels of the ground state of HCO: resonance energies, level widths and CO product state distributions”. J. Chem. Soc. Faraday Trans., vol. 88, p. 2575, 1992.

[40] Wang, D. and Bowman, J.M. “L2 calculations of resonances and final rotational distributions for HCO H+CO”. J. Chem. Phys., vol. 100, ! p. 1021, 1994.

[41] Murray, K.K., Miller, T.M., Leopold, D.G., and Lineberger, W.C. “Laser photoelectron spectroscopy of the formyl anion”. J. Chem. Phys., vol. 84, p. 2520, 1986.

[42] Chuang, M.C., Foltz, M.F., and Moore, C.B. “T1 barrier height, S1 - T1 intersystem crossing rate, and S0 radical dissociation threshold for H2CO, D2CO, and HDCO”. J. Chem. Phys., vol. 87, p. 3855, 1987.

[43] Sappey, A.D. and Crosley, D.R. “Laser-induced fluorescence in the B - X system of the HCO radical”. J. Chem. Phys., vol. 93, p. 7601, 1990. e e [44] Rumbles, G., Lee, E.K.C., and Valentini, J.J. “Laser-induced fluores- cence from the predissociating formyl radical. Part 2. Analysis of dis- persed emission from the A - X transition”. J. Chem. Soc. Faraday Trans., vol. 86, p. 3837, 1990. e e [45] Adamson, G.W., Zhao, X.S., and Field, R.W. “The HCO B2A X2A 0 $ 0 system: Fluorescence excitation and stimulated emission pumping spectra”. J. Mol. Spectrosc., vol. 160, p. 11, 1993. e e

[46] Neyer, D.W., Luo, X., Burak, I., and Houston, P.L. “Photodissocia- 2 tion dynamics of state-selected resonances of HCO X A0 prepared by stimulated emission pumping”. J. Chem. Phys., vol. 102, p. 1645, 1995. e [47] McBane, G.C., Kable, S.H., Houston, P.L., and Schatz, G.C. “Colli- sional excitation of CO by 2.3 eV H atoms”. J. Chem. Phys., vol. 94, p. 1141, 1991.

[48] Green, S., Pan, B., and Bowman, J.M. “Quantum scattering calcula- tions for vibrational and rotational excitation of CO by hot hydrogen atoms”. J. Chem. Phys., vol. 102, p. 8800, 1995.

[49] Green, S., Keller, H.M., Schinke, R., and Werner, H.J. “Vibration- rotation excitation of CO by hot hydrogen atoms: Comparison of two potential energy surfaces”. J. Chem. Phys., vol. 105, p. 5416, 1996.

[50] Werner, H.J. and Knowles, P.J. “A second order multiconfiguration SCF procedure with optimum convergence”. J. Chem. Phys., vol. 82, p. 5053, 1985.

Preliminary version – January 29, 2016 Bibliography 131

[51] Knowles, P.J. and Werner, H.J. “An efficient second-order MC SCF method for long configuration expansions”. Chem. Phys. Lett., vol. 115, p. 259, 1985.

[52] Dunning, T.H. “Gaussian basis sets for use in correlated molecu- lar calculations. I. The atoms boron through neon and hydrogen”. J. Chem. Phys., vol. 90, p. 1007, 1989.

[53] Helgaker, T., Jørgensen, P., and Olsen, J. Molecular Electronic Structure Theory. Wiley, Chichester, 2000.

[54] Knowles, P.J., Hampel, C., and Werner, H.J. “Coupled cluster theory for high spin, open shell reference wave functions”. J. Chem. Phys., vol. 99, p. 5219, 1993.

[55] Knowles, P.J., Hampel, C., and Werner, H.J. “Erratum: "Coupled cluster theory for high spin, open shell reference wave functions” [J. Chem. Phys. 99, 5219 (1993)]”. J. Chem. Phys., vol. 112, p. 3106, 2000.

[56] Watts, J.D., Gauss, J., and Bartlett, R.J. “Coupled-cluster methods with noniterative triple excitations for restricted open-shell Hartree–Fock and other general single determinant reference functions. Energies and analytical gradients”. J. Chem. Phys., vol. 98, p. 8718, 1993.

[57] Werner, H.J., Knowles, P.J., Knizia, G., Manby, F.R., Schütz, M., et al. “Molpro, version 2010.1, a package of ab initio programs”. See http://www.molpro.net, 2010.

[58] van Mourik, T. and Dunning, T.H. “A new ab initio potential energy curve for the helium dimer”. J. Chem. Phys., vol. 111, p. 9248, 1999.

[59] van Mourik, T., Wilson, A.K., and Dunning, T.H. “Benchmark calcula- tions with correlated molecular wavefunctions. XIII. Potential energy curves for He2, Ne2 and Ar2 using correlation consistent basis sets through augmented sextuple zeta”. Mol. Phys., vol. 96, p. 529, 1999.

[60] Boys, S.F. and Bernardi, F. “The calculation of small molecular inter- actions by the differences of separate total energies. Some procedures with reduced errors”. Mol. Phys., vol. 19, p. 553, 1970.

[61] Lee, T.J. and Taylor, P.R. “A diagnostic for determing the quality of single-reference electron correlation methods”. Int. J. Quantum. Chem. Symp., vol. 23, p. 199, 1989.

[62] Jayatilaka, D. and Lee, T.J. “Open-shell coupled-cluster theory”. J. Chem. Phys., vol. 98, p. 9734, 1993.

[63] Lee, T.J. “Comparison of the T1 and D1 diagnostics for electronic structure theory: a new definition for the open-shell D1 diagnostic”. Chem. Phys. Lett., vol. 372, p. 362, 2003.

Preliminary version – January 29, 2016 132 Bibliography

[64] Le Floch, A. “Revised molecular constants for the ground state of CO”. Mol. Phys., vol. 72, p. 133, 1991.

[65] Roy, R.J.L. “A computer program implementing the first-order RKR method for determining diatomic molecule potential energy curves”. Tech. rep., University of Waterloo Chemical Physics Research Report CP-657R, 2004.

[66] Tikhonov, A.N. and Arsenin, V.Y. Solutions of Ill-posed problems. Hal- sted, New York, 1977.

[67] van der Avoird, A., Wormer, P., Mulder, F., and Berns, R. “Ab initio studies of the interactions in Van der Waals molecules”. In “Van der Waals Systems”, vol. 93 of Topics in Current Chemistry, pp. 1–51. Springer Berlin Heidelberg, 1980.

[68] Tang, K.T. and Toennies, J.P. “An improved simple model for the van der Waals potential based on universal damping functions for the dispersion coefficients”. J. Chem. Phys., vol. 80, p. 3726, 1984.

[69] Ho, T.S. and Rabitz, H. “A general method for constructing multidi- mensional molecular potential energy surfaces from ab initio calcula- tions”. J. Chem. Phys., vol. 104, p. 2584, 1996.

[70] Ho, T.S. and Rabitz, H. “Proper construction of ab initio global po- tential surfaces with accurate long-range interactions”. J. Chem. Phys., vol. 113, p. 3960, 2000.

[71] Colbert, D.T. and Miller, W.H. “A novel discrete variable represen- tation for quantum mechanical reactive scattering via the S-matrix Kohn method”. J. Chem. Phys., vol. 96, p. 1982, 1992.

[72] Groenenboom, G.C. and Colbert, D.T. “Combining the discrete vari- able representation with the S-matrix Kohn method for quantum re- active scattering”. J. Chem. Phys., vol. 99, p. 9681, 1993.

[73] Krupenie, P.H. and Weissman, S. “Potential-Energy Curves for CO and CO+”. J. Chem. Phys., vol. 43, p. 1529, 1965.

[74] Yardley, J.T. “Laser action in highly-excited vibrational levels of CO”. J. Mol. Spectrosc., vol. 35, p. 314, 1970.

[75] Mantz, A.W., Maillard, J.P., Roh, W.B., and Rao, K.N. “Ground state molecular constants of 12C16O”. J. Mol. Spectrosc., vol. 57, p. 155, 1975.

[76] Austin, J.A., Levy, D.H., Gottlieb, C.A., and Radford, H.E. “Mi- crowave spectrum of the HCO radical”. J. Chem. Phys., vol. 60, p. 207, 1974.

[77] Johns, J.W.C., Priddle, S.H., and Ramsay, D.A. “Electronic absorption spectra of HCO and DCO radicals”. Discuss. Faraday Soc., vol. 35, p. 90, 1963.

Preliminary version – January 29, 2016 Bibliography 133

[78] Wang, H.Y., Eyre, J.A., and Dorfman, L.M. “Activation energy for the gas phase reaction of hydrogen atoms with carbon monoxyde”. J. Chem. Phys., vol. 59, p. 5199, 1973.

[79] Lee, K.T. and Bowman, J.M. “Rotational distributions from resonant and direct scattering in H+CO and tests of statistical theories”. J. Chem. Phys., vol. 86, pp. 215, 1987.

[80] Sears, T.J. “The calculation of the energy levels of an asymmetric top free radical in a magnetic field”. Comp. Phys. Repts., vol. 2, p. 1, 1984.

[81] Brown, J.M., Radford, H.E., and Sears, T.J. “Avoided crossings in the far-infrared laser magnetic resonance spectrum of HCO”. J. Mol. Spectrosc., vol. 148, p. 20, 1991.

[82] Tennyson, J., Kostin, M.A., Barletta, P., Harris, G.J., Polyansky, O.L., Ramanlal, J., and Zobov, N.F. “DVR3D: a program suite for the calcu- lation of rotation-vibration spectra of triatomic molecules”. Comput. Phys. Commun., vol. 163, p. 85, 2004.

[83] Clegg, P.E., Ade, P.A.R., Armand, C., et al. “The ISO long-wavelength spectrometer”. Astron. Astrophys., vol. 315, pp. L38, 1996.

[84] Kessler, M.F., Steinz, J.A., Anderegg, M.E., et al. “The infrared space observatory (ISO) mission”. Astron. Astrophys., vol. 315, pp. L27, 1996.

[85] Cernicharo, J., Barlow, M.J., Gonzalez-Alfonso, E., et al. “The ISO/LWS far infrared spectrum of IRC+10216.” Astron. Astrophys., vol. 315, pp. L201, 1996.

[86] Polehampton, E. T., Menten, K. M., van der Tak, F. F. S., and White, G. J. “The ISO long wavelength spectrometer line spectrum of VY Canis Majoris and other oxygen-rich evolved stars”. Astron. Astrophys., vol. 510, p. A80, 2010.

[87] Liu, X.W., Barlow, M.J., Nguyen-Q-Rieu, et al. “The ISO LWS grating spectrum of NGC 7027”. Astron. Astrophys., vol. 315, pp. L257, 1996.

[88] Nisini, B., Lorenzetti, D., Cohen, M., et al. “LWS-spectroscopy of Her- big Haro objects and molecular outflows in the Cha II dark cloud”. Astron. Astrophys., vol. 315, pp. L321, 1996.

[89] Schneider, N., Stutzki, J., Winnewisser, G., and Block, D. “The Rosette Molecular Complex. I. CO observations”. Astron. Astrophys., vol. 335, pp. 1049, 1998.

[90] Nordh, H.L., von Schéele, F., Frisk, U., et al. “The Odin orbital obser- vatory”. Astron. Astrophys., vol. 402, pp. L21, 2003.

Preliminary version – January 29, 2016 134 Bibliography

[91] Persson, C.M., Olofsson, A.O.H., Koning, N., et al. “A spectral line survey of Orion KL in the bands 486-492 and 541-577 GHz with the Odin satellite. II. Data analysis”. Astron. Astrophys., vol. 476, pp. 807, 2007.

[92] Pilbratt, G.L., Riedinger, J.R., Passvogel, T., et al. “Herschel Space Observatory. An ESA facility for far-infrared and submillimetre as- tronomy”. Astron. Astrophys., vol. 518,L1, 2010.

[93] Pilleri, P., Fuente, A., Cernicharo, J., et al. “Herschel/HIFI observa- tions of CO, H2O and NH3 in Monoceros R2”. Astron. Astrophys., vol. 544,A110, 2012.

[94] Yıldız, U.A., van Dishoeck, E.F., Kristensen, L.E., et al. “Her- schel/HIFI observations of high-J CO lines in the NGC 1333 low-mass star-forming region”. Astron. Astrophys., vol. 521,L40, 2010.

[95] de Graauw, T., Helmich, F.P., Phillips, T.G., et al. “The Herschel- Heterodyne Instrument for the Far-Infrared (HIFI)”. Astron. Astro- phys., vol. 518,L6, 2010.

[96] Griffin, M.J., Abergel, A., Abreu, A., et al. “The Herschel-SPIRE in- strument and its in-flight performance”. Astron. Astrophys., vol. 518, L3, 2010.

[97] Panuzzo, P., Rangwala, N., Rykala, A., et al. “Probing the molecu- lar interstellar medium of M82 with Herschel-SPIRE spectroscopy”. Astron. Astrophys., vol. 518,L37, 2010.

[98] Goldsmith, P.F. “Diffuse Molecular Cloud Densities from UV Mea- surements of CO Absorption”. Astrophys. J., vol. 774, 134, 2013.

[99] Chu, S.I. and Dalgarno, A. “The rotational excitation of carbon monoxide by hydrogen atom impact”. R. Soc. London Proc. Ser. A, vol. 342, pp. 191, 1975.

[100] Green, S. and Thaddeus, P. “Rotational excitation of CO by collisions with He, H, and H2 under conditions in interstellar clouds”. Astro- phys. J., vol. 205, pp. 766, 1976.

[101] Dunning, T.H. and Hay, P. Methods of Electronic Structure Theory. New York: Plenum, 1977.

[102] Huzinaga, S. “Approximate Atomic Wave Functions. I, Chemistry Report”. Tech. rep., Edmonton: University of Alberta, 1971.

[103] Purvis, G.D. and Bartlett, R.J. “A full coupled-cluster singles and doubles model: The inclusion of disconnected triples”. J. Chem. Phys., vol. 76, pp. 1910, 1982.

Preliminary version – January 29, 2016 Bibliography 135

[104] Werner, H.J. and Knowles, P.J. “An efficient internally contracted multiconfiguration-reference configuration interaction method”. J. Chem. Phys., vol. 89, pp. 5803, 1988.

[105] Woon, D.E. and Dunning, T.H. “Gaussian basis sets for use in cor- related molecular calculations. IV. Calculation of static electrical re- sponse properties”. J. Chem. Phys., vol. 100, pp. 2975, 1994.

[106] Song, L., van der Avoird, A., and Groenenboom, G.C. “Three- dimensional ab initio potential energy surface for H-CO(X2A0)”. J. Phys. Chem. A, vol. 117, pp. 7571, 2013. e [107] Arthurs, A.M. and Dalgarno, A. “The theory of scattering by a rigid rotator”. P. Roy. Soc. A Math. Phy., vol. 256, pp. 540, 1960.

[108] McGuire, P. and Kouri, D.J. “Quantum mechanical close coupling ap- proach to molecular collisions. jz-conserving coupled states approxi- mation”. J. Chem. Phys., vol. 60, p. 2488, 1974.

[109] Hutson, J.M. and Green, S. Molscat computer code, version 14 (1994), distributed by Collaborative Computational Project No. 6 of the Engineering and Physical Sciences Research Council (UK).

[110] Valiron, P. and McBane, G.C. Mixed MPI and OpenMP extension of the MOLSCAT v14 package, 2008.

[111] Walker, K. VRRMM: Vibrational/Rotational Rich Man’s MOLSCAT v3.1, 2013.

[112] Alexander, M.H. and Manolopoulos, D.E. “A stable linear reference potential algorithm for solution of the quantum close-coupled equa- tions in molecular scattering theory”. J. Chem. Phys., vol. 86, pp. 2044, 1987.

[113] Yang, B., Perera, H., Balakrishnan, N., Forrey, R.C., and Stancil, P.C. “Quenching of rotationally excited CO in cold and ultracold collisions with H, He and H2”. J. Phys. B At. Mol. Opt., vol. 39, p. S1229, 2006.

[114] Osterbrock, D.E. and Ferland, G.J. Astrophysics of Gaseous Nebulae and Active Galactic Nuclei. University Science Books, 2006.

[115] Pontoppidan, K.M., Blake, G.A., van Dishoeck, E.F., Smette, A., Ire- land, M.J., and Brown, J. “Spectroastrometric imaging of molecular gas within protoplanetary disk gaps”. Astrophys. J., vol. 684, p. 1323, 2008.

[116] van der Plas, G., van den Ancker, M.E., Acke, B., Carmona, A., Do- minik, C., Fedele, D., and Waters, L.B.F.M. “Evidence for CO deple- tion in the inner regions of gas-rich protoplanetary disks”. Astron. Astrophys., vol. 500, p. 1137, 2009.

Preliminary version – January 29, 2016 136 Bibliography

[117] Pack, R.T. “Space-fixed vs body-fixed axes in atom-diatomic molecule scattering. Sudden approximations”. J. Chem. Phys., vol. 60, p. 633, 1974.

[118] Lester, W.A. The N coupled-channel problem, pp. 1–32. Plenum, New York, 1976.

[119] von Rosenberg, C.W., Taylor, R.L., and Teare, J.D. “Vibrational relax- ation of CO in nonequilibrium nozzle flow, and the effect of hydrogen atoms on CO relaxation”. J. Chem. Phys., vol. 54, p. 1974, 1971.

[120] Glass, G.P. and Klronde, S. “Vibrational relaxation of carbon monox- ide in collisions with atomic hydrogen”. J. Phys. Chem., vol. 86, p. 908, 1982.

[121] Kozlov, P.V., Makarov, V.N., Pavlov, V.A., and Shatalov, O.P. “Exper- imental investigation of CO vibrational deactivation in a supersonic cooling gas flow”. Shock Waves, vol. 10, p. 191, 2000.

[122] Yang, B., Stancil, P.C., and Balakrishnan, N. “A close-coupling study of vibrational-rotational quenching of CO by collision with hydrogen atoms”. J. Chem. Phys., vol. 123, p. 094308, 2005.

[123] Heijmen, T.G.A., Moszynski, R., Wormer, P.E.S., and van der Avoird, A. “A new He-CO interaction energy surface with vibrational coor- dinate dependence. I. Ab initio potential and infrared spectrum”. J. Chem. Phys., vol. 107, pp. 9921, 1997.

[124] Peterson, K.A. and McBane, G.C. “Erratum: "A hierarchical family of three-dimensional potential energy surfaces for He-CO" [J. Chem. Phys.123, 084314 (2005)]”. J. Chem. Phys., vol. 124, p. 229901, 2006.

[125] Jankowski, P., McKellar, A.R.W., and Szalewicz, K. “Theory untangles the high-resolution infrared spectrum of the ortho-H2-CO van der Waals complex”. Science, vol. 336, pp. 1147, 2012.

[126] Jankowski, P., Surin, L.A., Potapov, A., Schlemmer, S., McKellar, A.R.W., and Szalewicz, K. “A comprehensive experimental and the- oretical study of H2-CO spectra”. J. Chem. Phys., vol. 138, p. 084307, 2013.

[127] Scilab Enterprises. Scilab: Free and Open Source software for numerical computation. Scilab Enterprises, Orsay, France, 2012.

[128] Martinazzo, R., Bodo, E., and Gianturco, F.A. “A modified Variable- Phase algorithm for multichannel scattering with long-range poten- tials”. Comput. Phys. Commun., vol. 151, p. 187, 2003.

[129] Krems, R.V. “Vibrational relaxation of vibrationally and rotationally excited CO molecules by He atoms”. J. Chem. Phys., vol. 116, p. 4517, 2002.

Preliminary version – January 29, 2016 Bibliography 137

[130] Neufeld, D.A. and Hollenbach, D.J. “Dense molecular shocks and accretion onto protostellar disks”. Astrophys. J., vol. 428, p. 170, 1994.

[131] Hollenbach, D. and McKee, C.F. “Molecule formation and infrared emission in fast interstellar shocks. I. Physical processes”. Astrophys. J. Suppl. Ser., vol. 41, p. 555, 1979.

[132] Song, L., Balakrishnan, N., van der Avoird, A., Karman, T., and Groe- nenboom, G.C. “Quantum scattering calculations for ro-vibrational de-excitation of CO by hydrogen atoms”. J. Chem. Phys., vol. 142, p. 204303, 2015.

[133] Thompson, R.I., Schnopper, H.W., Mitchell, R.I., and Johnson, H.L. “1-4-MICRON Spectra of Four M Stars and Alpha Tauri”. Astrophys. J. Lett., vol. 158, p. L117, 1969.

[134] Johnson, H.L. and Méndez, M.E. “Infrared spectra for 32 stars.” As- tron. J., vol. 75, pp. 785, 1970.

[135] Johnson, H.L., Thompson, R.I., Forbes, F.F., and Steinmetz, D.L. “The infrared spectrum of Alpha Herculis from 4000 to 4800 cm-1”. Publ. Astron. Soc. Pac., vol. 84, p. 775, 1972.

[136] Scoville, N.Z., Krotkov, R., and Wang, D. “Collisional and infrared ra- diative pumping of molecular vibrational states - The carbon monox- ide infrared bands”. Astrophys. J., vol. 240, pp. 929, 1980.

[137] Ayres, T. “The dichotomy between CO absorption and Ca II emis- sions in the sun and stars - an indirect diagnostic for gas disturbed by magnetic fields?” High. Astron., vol. 7, pp. 425, 1986.

[138] Mitchell, G.F., Maillard, J.P., Allen, M., Beer, R., and Belcourt, K. “Hot and cold gas toward young stellar objects”. Astrophys. J., vol. 363, pp. 554, 1990.

[139] Najita, J., Carr, J.S., and Mathieu, R.D. “Gas in the Terrestrial Planet region of disks: CO fundamental emission from T Tauri stars”. Astro- phys. J., vol. 589, pp. 931, 2003.

[140] Rettig, T.W., Haywood, J., Simon, T., Brittain, S.D., and Gibb, E. “Dis- covery of CO gas in the lnner disk of TW Hydrae”. Astrophys. J. Lett., vol. 616, pp. L163, 2004.

[141] Carmona, A., van den Ancker, M.E., Thi, W.F., Goto, M., and Henning, T. “Upper limits on CO 4.7 µm emission from disks around five Herbig Ae/Be stars”. Astron. Astrophys., vol. 436, pp. 977, 2005.

[142] Salyk, C., Blake, G.A., Boogert, A.C.A., and Brown, J.M. “Molecular gas in the inner 1 AU of the TW Hya and GM Aur transitional disks”. Astrophys. J. Lett., vol. 655, pp. L105, 2007.

Preliminary version – January 29, 2016 138 Bibliography

[143] Brittain, S.D., Najita, J.R., and Carr, J.S. “Tracing the inner edge of the disk around HD 100546 with rovibrational CO emission lines”. Astrophys. J., vol. 702, pp. 85, 2009.

[144] Goto, M., Regály, Z., Dullemond, C.P., et al. “Fundamental vibrational transition of CO during the outburst of EX Lupi in 2008”. Astrophys. J., vol. 728, p. 5, 2011.

[145] Brown, J.M., Pontoppidan, K.M., van Dishoeck, E.F., Herczeg, G.J., Blake, G.A., and Smette, A. “VLT-CRIRES survey of rovibrational CO emission from protoplanetary disks”. Astrophys. J., vol. 770, p. 94, 2013.

[146] Hein Bertelsen, R. P., Kamp, I., Goto, M., van der Plas, G., Thi, W.-F., Waters, L. B. F. M., van den Ancker, M. E., and Woitke, P. “CO ro- vibrational lines in HD 100546 - A search for disc asymmetries and the role of florescence”. Astron. Astrophys., vol. 561, p. A102, 2014.

[147] van der Tak, F. “Radiative transfer and molecular data for astrochem- istry”. In J. Cernicharo and R. Bachiller, editors, “The Molecular Uni- verse”, vol. 7 of Proceedings of the International Astronomical Union, pp. 449–460, 2011.

[148] Walker, K.M., Yang, B.H., Stancil, P.C., Balakrishnan, N., and Forrey, R.C. “On the validity of collider-mass scaling for molecular rotational excitation”. Astrophys. J., vol. 790, p. 96, 2014.

[149] Walker, K.M., Song, L., Yang, B.H., Groenenboom, G.C., van der Avoird, A., Balakrishnan, N., Forrey, R.C., and Stancil, P.C. “Quan- tum calculation of inelastic CO collisions with H. II. Pure rotational quenching of high rotational levels”. Astrophys. J., vol. 811, p. 27, 2015.

[150] Faure, A. and Josselin, E. “Collisional excitation of water in warm astrophysical media. I. Rate coefficients for rovibrationally excited states”. Astron. Astrophys., vol. 492, pp. 257, 2008.

[151] Bruderer, S., Harsono, D., and van Dishoeck, E.F. “Ro-vibrational excitation of an organic molecule (HCN) in protoplanetary disks”. Astron. Astrophys., vol. 575, p. A94, 2015.

[152] Woitke, P., Kamp, I., and Thi, W.-F. “Radiation thermo-chemical mod- els of protoplanetary disks - I. Hydrostatic disk structure and inner rim”. Astron. Astrophys., vol. 501, pp. 383, 2009.

[153] Meijerink, R., Spaans, M., Kamp, I., Aresu, G., Thi, W.F., and Woitke, P. “Tracing the Physical Conditions in Active Galactic Nuclei with Time-Dependent Chemistry”. J. Phys. Chem. A, vol. 117, pp. 9593, 2013.

Preliminary version – January 29, 2016 Bibliography 139

[154] Hollenbach, D. and McKee, C.F. “Molecule formation and infrared emission in fast interstellar shocks. III - Results for J shocks in molec- ular clouds”. Astrophys. J., vol. 342, pp. 306, 1989.

[155] van der Plas, G., van den Ancker, M. E., Waters, L. B. F. M., and Dominik, C. “The structure of disks around Herbig Ae/Be stars as traced by CO ro-vibrational emission”. Astron. Astrophys., vol. 574, p. A75, 2015.

[156] Neufeld, D.A. and Dalgarno, A. “Fast molecular shocks. I - Reforma- tion of molecules behind a dissociative shock”. Astrophys. J., vol. 340, pp. 869, 1989.

[157] Suits, A.G., Bontuyan, L.S., Houston, P.L., and Whitaker, B.J. “Differ- ential cross sections for state-selected products by direct imaging: Ar+ NO”. J. Chem. Phys., vol. 96, pp. 8618, 1992.

[158] Antonova, S., Lin, A., Tsakotellis, A.P., and McBane, G.C. “State to state He-CO rotationally inelastic scattering”. J. Chem. Phys., vol. 110, pp. 2384, 1999.

[159] Bontuyan, L.S., Suits, A.G., Houston, P.L., and Whitaker, B.J. “State- resolved differential cross sections for crossed-beam argon-nitric ox- ide inelastic scattering by direct ion imaging”. J. Phys. Chem., vol. 97, pp. 6342, 1993.

[160] von Zastrow, A., Onvlee, J., Vogels, S.N., Groenenboom, G.C., van der Avoird, A., and van de Meerakker, S.Y.T. “State-resolved diffraction oscillations imaged for inelastic collisions of NO radicals with He, Ne and Ar”. Nat. Chem., vol. 6, p. 216, 2014.

[161] Gilijamse, J.J., Hoekstra, S., van de Meerakker, S.Y.T., Groenenboom, G.C., and Meijer, G. “Near-threshold inelastic collisions using molec- ular beams with a tunable velocity”. Science, vol. 313, pp. 1617, 2006.

[162] Kay, J.J., Steill, J.D., Kłos, J., Paterson, G., Costen, M.L., Strecker, K.E., McKendrick, K.G., Alexander, M.H., and Chandler, D.W. “Collisions of electronically excited molecules: differential cross-sections for ro- tationally inelastic scattering of NO (A2⌃+) with Ar and He”. Mol. Phys., vol. 110, pp. 1693, 2012.

[163] Lorenz, K.T., Chandler, D.W., Barr, J.W., Chen, W., Barnes, G.L., and Cline, J.I. “Direct measurement of the preferred sense of NO rotation after collision with argon”. Science, vol. 293, pp. 2063, 2001.

[164] Meyer, H. “State-Resolved Cross Sections and Collision-Induced Alignment from Counterpropagating Beam Scattering of NH3+ He”. J. Phys. Chem., vol. 99, pp. 1101, 1995.

Preliminary version – January 29, 2016 140 Bibliography

[165] Wade, E.A., Thomas Lorenz, K., Chandler, D.W., Barr, J.W., Barnes, G.L., and Cline, J.I. “Ion imaging studies of product rotational align- 2 ment in collisions of NO (X ⇧1/2, j = 0.5) with Ar”. Chem. Phys., vol. 301, pp. 261, 2004.

[166] Brouard, M., Chadwick, H., Eyles, C.J., Aoiz, F.J., and Kłos, J. “The k-j-j’ vector correlation in inelastic and reactive scattering”. J. Chem. Phys., vol. 135, p. 084305, 2011.

[167] Brouard, M., Hornung, B., and Aoiz, F.J. “Origin of Collision-Induced Molecular Orientation”. Phys. Rev. Lett., vol. 111, p. 183202, 2013.

[168] Cline, J.I., Lorenz, K.T., Wade, E.A., Barr, J.W., and Chandler, D.W. “Ion imaging measurement of collision-induced rotational alignment in Ar-NO scattering”. J. Chem. Phys., vol. 115, pp. 6277, 2001.

[169] Eyles, C.J., Brouard, M., Yang, C.H., Kłos, J., Aoiz, F.J., Gijsbertsen, A., Wiskerke, A.E., and Stolte, S. “Interference structures in the differen- tial cross-sections for inelastic scattering of NO by Ar”. Nat. Chem., vol. 3, pp. 597, 2011.

[170] Alexander, M.H. “Polarization and steric effects in inelastic collisions of NO (X2⇧) with Ar and He”. Faraday Discuss., vol. 113, pp. 437, 1999.

[171] Chandler, D.W. and Houston, P.L. “Two-Dimensional Imaging of State-Selected Photodissociation Process Detected by Multiphoton Ionization”. J. Chem. Phys., vol. 87, p. 1445, 1987.

[172] Eppink, A.T.J.B. and Parker, D.H. “Velocity map imaging of ions and electrons using electrostatic lenses: Application in photoelectron and photofragment ion imaging of molecular oxygen”. Rev. Sci. Instrum., vol. 68, pp. 3477, 1997.

[173] Hall, G.E., Sivakumar, N., Houston, P.L., and Burak, I. “Measure- ment of the Angular Correlation between Recoil Velocity and Angu- lar Momentum Vectors in Molecular Photodissociation”. Phys. Rev. Lett., vol. 56, pp. 1671, 1986.

[174] Hall, G.E. and Houston, P.L. “Vector correlations in photodissociation dynamics”. Annu. Rev. Phys. Chem., vol. 40, pp. 375, 1989.

[175] Bernstein, R.B., Herschbach, D.R., and Levine, R.D. “Dynamical as- pects of stereochemistry”. J. Phys. Chem., vol. 91, pp. 5365, 1987.

[176] Mestdagh, J.M., Visticot, J.P., Meynadier, P., Sublemontier, O., and Suits, A.G. “Four-vector study of the spin-changing collision Ba [(6s)1 (6p)1, 1P (6s)1 (6p)1, 3P ]/H ”. J. Chem. Soc., Faraday Trans., vol. 89, 1 ! 2 2 pp. 1413, 1993.

Preliminary version – January 29, 2016 Bibliography 141

[177] Balint-Kurti, G.G. and Vasyutinskii, O.S. “Vector Correlation Analy- sis for Inelastic and Reactive Collisions between Partners Possessing Spin and Orbital Angular Momentum”. J. Phys. Chem. A, vol. 113, pp. 14281, 2009.

[178] Kim, Y. and Meyer, H. “REMPI detection of alignment in NO colli- sions”. Chem. Phys., vol. 301, pp. 273, 2004.

[179] Steill, J.D., Kay, J.J., Paterson, G., Sharples, T.R., Kłos, J., Costen, M.L., Strecker, K.E., McKendrick, K.G., Alexander, M.H., and Chandler, D.W. “Rotational Alignment of NO (A2⌃+) from Collisions with Ne”. J. Phys. Chem. A, vol. 117, pp. 8163, 2013.

[180] Brouard, M., Chadwick, H., Eyles, C.J., Hornung, B., Nichols, B., Aoiz, F.J., Jambrina, P.G., and Stolte, S. “Rotational alignment effects in NO(X) + Ar inelastic collisions: An experimental study”. J. Chem. Phys., vol. 138, 104310, 2013.

[181] Hertel, I.V. and Stoll, W. “Collision experiments with laser excited atoms in crossed beams”. Adv. Atom. Mol. Phys., vol. 13, pp. 113, 1978.

[182] Orr-Ewing, A.J. and Zare, R.N. “Orientation and Alignment of Reac- tion Products”. Annu. Rev. Phys. Chem., vol. 45, pp. 315, 1994.

[183] Yan, B., Claus, P.F.H., van Oorschot, B.G.M., Gerritsen, L., Eppink, A.T.J.B., van de Meerakker, S.Y.T., and Parker, D.H. “A new high intensity and short-pulse molecular beam valve”. Rev. Sci. Instrum., vol. 84, 023102, 2013.

[184] Tunnermann, A., Momma, C., Mossavi, K., Windolph, C., and Welleghausen, B. “Generation of tunable short pulse VUV radiation by four-wave mixing in xenon with femtosecond KrF-excimer laser pulses”. IEEE J. Quantum. Elect., vol. 29, pp. 1233, 1993.

[185] Hilbig, R. and Wallenstein, R. “Tunable VUV radiation generated by two-photon resonant frequency mixing in xenon”. IEEE J. Quantum. Elect., vol. 19, pp. 194, 1983.

[186] Wu, S.M., Yang, X., and Parker, D.H. “Velocity map imaging study of OCS photodissociation followed by S (1S) autoionization at 157 nm”. Mol. Phys., vol. 103, pp. 1797, 2005.

[187] Bondo Pedersen, T., López Cacheiro, J., Fernández, B., and Koch, H. “Rovibrational structure of the Ar-CO complex based on a novel three- dimensional ab initio potential”. J. Chem. Phys., vol. 117, pp. 6562, 2002.

[188] Alexander, M.H. “Close-coupling studies of the orientation depen- dence of rotationally inelastic collisions”. J. Chem. Phys., vol. 67, p. 2703, 1977.

Preliminary version – January 29, 2016 142 Bibliography

[189] Fano, U. and Macek, J.H. “Impact Excitation and Polarization of the Emitted Light”. Rev. Mod. Phys., vol. 45, p. 553, 1973.

[190] Brouard, M., Chadwick, H., Eyles, C.J., Hornung, B., Nichols, B., Scott, J.M., Aoiz, F.J., Kłos, J., Stolte, S., and Zhang, X. “The fully quantum state-resolved inelastic scattering of NO(X) + Ne: experiment and the- ory”. Mol. Phys., vol. 111, pp. 1759, 2013.

[191] Nichols, B., Chadwick, H., Gordon, S.D.S., Eyles, C.J., Hornung, B., Brouard, M., Alexander, M.H., Aoiz, F.J., Gijsbertsen, A., and Stolte, S. “Steric effects and quantum interference in the inelastic scattering of NO(X) + Ar”. Chem. Sci., vol. 6, pp. 2202, 2015.

[192] Kirste, M., Wang, X., Schewe, H.C., Meijer, G., Liu, K., van der Avoird, A., Janssen, L.M.C., Gubbels, K.B., Groenenboom, G.C., and van de Meerakker, S.Y.T. “Quantum-state resolved bimolecular collisions of velocity-controlled OH with NO radicals”. Science, vol. 338, pp. 1060, 2012.

[193] Yang, C.H., Sarma, G., Parker, D.H., ter Meulen, J.J., and Wiesenfeld, L. “State-to-state differential and relative integral cross sections for rotationally inelastic scattering of H2O by hydrogen”. J. Chem. Phys., vol. 134, 204308, 2011.

[194] Tkáˇc,O., Saha, A.K., Loreau, J., Ma, Q., Dagdigian, P.J., Parker, D.H., van der Avoird, A., and Orr-Ewing, A.J. “Rotationally inelastic scatter- ing of ND3 with H2 as a probe of the intermolecular potential energy surface”. Mol. Phys., vol. 113, p. 3925, 2015.

[195] Tkáˇc,O., Ma, Q., Stei, M., Orr-Ewing, A.J., and Dagdigian, P.J. “Ro- tationally inelastic scattering of methyl radicals with Ar and N2”. J. Chem. Phys., vol. 142, 014306, 2015.

[196] Peterson, K.A. and McBane, G.C. “A hierarchical family of three- dimensional potential energy surfaces for He-CO”. J. Chem. Phys., vol. 123, 084314, 2005.

[197] Bergeat, A., Onvlee, J., Naulin, C., van der AvoirdAd, and Costes, M. “Quantum dynamical resonances in low-energy CO (j = 0) +He inelastic collisions”. Nat. Chem., vol. 7, pp. 349, 2015.

[198] Hines, M.A., Michelsen, H.A., and Zare, R.N. “2+1 resonantly en- hanced multiphoton ionization of CO via the E1⇧-X1⌃+ transition: From measured ion signals to quantitative population distributions”. J. Chem. Phys., vol. 93, pp. 8557, 1990.

[199] Suits, A.G., Bishwakarma, C.K., Song, L., Groenenboom, G.C., van der Avoird, A., and Parker, D.H. “Direct extraction of alignment moments from inelastic scattering images”. J. Phys. Chem. A, vol. 119, pp. 5925, 2015.

Preliminary version – January 29, 2016 Bibliography 143

[200] Butz, H., Feltgen, R., Pauly, H., and Vehmeyer, H. “Totale effektive Streuquerschnitte für die Streuung von He, HD und D2 an verschiede- nen Molekülen”. Z. Phys. A Hadron. Nucl., vol. 247, pp. 70, 1971.

[201] Nerf, R.B. and Sonnenberg, M.A. “Pressure broadening of the J = 1 0 transition of carbon monoxide”. J. Mol. Spectrosc., vol. 58, pp.

474 , 1975.

[202] LeRoy, R.J., Bissonnette, C., Wu, T.H., Dham, A.K., and Meath, W.J. “Improved modelling of atom-molecule potential-energy surfaces: il- lustrative application to He-CO”. Faraday Discuss., vol. 97, pp. 81, 1994.

[203] Lorenz, K.T., Chandler, D.W., and McBane, G.C. “State-to-state differ- ential cross sections by velocity mapping for rotational excitation of CO by Ne”. J. Phys. Chem. A, vol. 106, pp. 1144, 2002.

[204] McBane, G.C. and Cybulski, S.M. “An ab initio potential energy sur- face for Ne-CO”. J. Chem. Phys., vol. 110, pp. 11734, 1999.

[205] Moszynski, R., Korona, T., Wormer, P.E.S., and van der Avoird, A. “Ab initio potential energy surface and infrared spectrum of the Ne-CO complex”. J. Phys. Chem. A, vol. 101, pp. 4690, 1997.

[206] Smith, T.C., Hostutler, D.A., Hager, G.D., Heaven, M.C., and McBane, G.C. “State-to-state rotational rate constants for CO+He: Infrared dou- ble resonance measurements and simulation of the data using the SAPT theoretical potential energy surface”. J. Chem. Phys., vol. 120, pp. 2285, 2004.

[207] Levine, R.D. and Bernstein, R.B. Molecular reaction dynamics. Oxford University Press, 1974.

[208] Levine, R.D. Molecular reaction dynamics. Cambridge University Press, 2005.

[209] Tkáˇc,O., Saha, A.K., Onvlee, J., Yang, C.H., Sarma, G., Bishwakarma, C.K., van de Meerakker, S.Y.T., van der Avoird, A., Parker, D.H., and Orr-Ewing, A.J. “State-to-state resolved differential cross sections for rotationally inelastic scattering of ND3 with He”. Phys. Chem. Chem. Phys., vol. 16, pp. 477, 2014.

[210] Baker, J., Lemaire, J.L., and Couris, S. “A 2+1 REMPI study of the E-X transition in CO”. AIP Conference Proceedings, vol. 312, pp. 355, 1994.

[211] Western, C.M. “Pgopher, a Program for Simulating Rotational Struc- ture, University of Bristol”. http://pgopher.chm.bris.ac.uk.

[212] McBane, G.C. Simulation and analysis of image data from crossed beam ex- periments, vol. 770, chap. 14, pp. 215–229. American Chemical Society, 2000.

Preliminary version – January 29, 2016 144 Bibliography

[213] Brouard, M., Chadwick, H., Gordon, S.D.S., Hornung, B., Nichols, B., Kłos, J., Aoiz, F.J., and Stolte, S. “Fully quantum state-resolved inelas- tic scattering of NO(X) + Kr: Differential cross sections and product rotational alignment”. J. Chem. Phys., vol. 141, 164306, 2014.

[214] Chadwick, H., Nichols, B., Gordon, S.D.S., Hornung, B., Squires, E., Brouard, M., Kłos, J., Alexander, M.H., Aoiz, F.J., and Stolte, S. “Inelas- tic Scattering of NO by Kr: Rotational Polarization over a Rainbow”. J. Phys. Chem. Lett., vol. 5, pp. 3296, 2014.

[215] Paglione, T.A.D., Wall, W.F., Young, J.S., Heyer, M.H., Richard, M., Goldstein, M., Kaufman, Z., Nantais, J., and Perry, G. “A Mapping Survey of the 13CO and 12CO Emission in Galaxies”. Astrophys. J. Suppl. Ser., vol. 135, pp. 183, 2001.

[216] Wada, K. and Tomisaka, K. “Molecular Gas Structure around an Active Galactic Nucleus with Nuclear Starburst: Three-dimensional Non-LTE Calculations of CO Lines”. Astrophys. J., vol. 619, pp. 93, 2005.

[217] Narayanan, D., Krumholz, M.R., Ostriker, E.C., and Hernquist, L. “A general model for the CO-H2 conversion factor in galaxies with appli- cations to the star formation law”. Mon. Not. R. Astron. Soc., vol. 421, pp. 3127, 2012.

[218] “Spin isomers of hydrogen–Wikipedia”. https://en.wikipedia.org/ wiki/Spin_isomers_of_hydrogen.

[219] Jankowski, P. and Szalewicz, K. “Ab initio potential energy surface and infrared spectra of H2-CO and D2-CO van der Waals complexes”. J. Chem. Phys., vol. 108, pp. 3554, 1998.

[220] Jankowski, P. and Szalewicz, K. “A new ab initio interaction energy surface and high-resolution spectra of the H2-CO van der Waals com- plex”. J. Chem. Phys., vol. 123, 104301, 2005.

[221] Antonova, S., Tsakotellis, A.P., Lin, A., and McBane, G.C. “State-to- -1 state rotational excitation of CO by H2 near 1000 cm collision en- ergy”. J. Chem. Phys., vol. 112, pp. 554, 2000.

[222] Yang, B., Zhang, P., Wang, X., Stancil, P.C., Bowman, J.M., Balakrish- nan, N., and Forrey, R.C. “Quantum dynamics of CO-H2 in full di- mensionality”. Nat Commun, vol. 6, p. 6629, 2015.

[223] Ma, Q., van der Avoird, A., Loreau, J., Alexander, M.H., van de Meer- akker, S.Y.T., and Dagdigian, P.J. “Resonances in rotationally inelastic scattering of NH3 and ND3 with H2”. J. Chem. Phys., vol. 143, p. 044312, 2015.

Preliminary version – January 29, 2016 SUMMARY

This thesis is devoted to a quantum mechanical study of inelastic collision processes for astronomical applications. We were interested, in particular, in the scattering of CO with its collision partners in space, such as H, He and H2. The main goal of the work was to provide state-to-state rate coef- ficients for pure rotational and ro-vibrational transitions of CO induced by inelastic collisions with H atoms. The calculated results were imported into the Leiden Atomic and Molecular Database (LAMDA) in use for current and future astrophysical modeling. Furthermore, we investigated He-CO, H2-CO and Ar-CO collisions. For those systems, differential cross sections and/or polarization dependent differential cross sections were calculated at the collision energies corresponding to the measurements of our experi- mental collaborators in this work. Our results, in comparison with the data from molecular beam experiments, allowed us to check the accuracy of the ab initio intermolecular interaction potentials used in our studies and to obtain detailed information on the dynamics of the collision processes.

Chapter 1 is the introduction, which describes the astrochemical back- ground of our work, the theoretical methods used, and the systems studied. It contains two subsections on theoretical methods, one with an introduc- tion to the methods applied in electronic structure computations, the other one on the principles of quantum-mechanical scattering methods.

Chapter 2 presents the construction of a three-dimensional ab initio po- 2 tential surface for the H-CO(X A0) complex. The ab initio calculations and the fit procedure are described in detail. The potential was used to calcu- late the vibrational frequenciese and rotational constants of H-CO. Compar- ison with experimental data demonstrates that the new three-dimensional potential energy surface reproduces multiple spectroscopic parameters pre- cisely. Furthermore, some preliminary two-dimensional scattering calcula- tions were performed and compared with previous calculations to confirm the accuracy of the long range part of the new potential.

Chapter 3 reports quantum scattering calculations for pure rotational de-excitation transitions of CO induced by collisions with H atoms, based on the three-dimensional H-CO potential in Chapter 2. State-to-state cross sections were calculated for collision energies ranging from 10-5 to 15000 cm-1, and the corresponding rate coefficients were produced for a temper- ature range of 1 to 3000 K. Both the full quantum close-coupling method and the coupled states approximation were used in calculations for CO de- excitation from rotational states with j = 1 - 5, 10, 15, 20, 25, 30, 35, 40 and 45 in the vibrational v = 0 ground state, while a scaling approach was intro- duced for the intermediate rotational states. Comparison with previous re-

145

Preliminary version – January 29, 2016 146 Bibliography

sults from scattering calculations by Balakrishnan et al. [Astrophys. J., 568, 443 (2002)] revealed that those results are not completely reliable, due to the use of a potential that is inaccurate in its long range part. Moreover, our calculations extend the temperature range and the number of transitions of the previously available H-CO pure rotational rate coefficients used for astrophysical modeling.

Chapter 4 describes full quantum close-coupling calculations for ro-vibr- ational relaxation of CO in collision with H atoms, based on our three- dimensional H-CO potential energy surface. First, we tested the conver- gence with respect to the channel basis set and the accuracy of the coupled- state and infinite-order-sudden approximations. Then, we computed a large number of state-to-state cross sections for ro-vibrational transitions from v = 1, j = 0 - 30 to v0 = 0, j0 at collision energies between 0.1 and 15000 cm-1. Pure vibrational v = 1 v = 0 transition cross sections and rate co- ! 0 efficients were obtained both from the full close-coupling method and from the infinite-order-sudden approximation. Comparison of these data shows that vibrational rate coefficients from the infinite-order-sudden approxima- tion agree well with those from full close-coupling calculations. The com- puted v = 1 v = 0 rate coefficients also agree well with data measured ! 0 by Glass et al. [J. Phys. Chem., 86, 908 (1982)].

Chapter 5 presents ro-vibrational transition rate coefficients for H-CO collisions for gas temperatures from 10 to 3000 K, obtained with the use of the cross sections from Chapter 4. Moreover, pure vibrational rate co- efficients for transitions v = 1 - 5 v

Chapter 6 describes experimental and theoretical studies on CO in colli- sions with Ar. Differential cross sections (DCSs) and polarization depen- dent differential cross sections (PDDCSs) were calculated using the full quantum close-coupling method. These physically relevant parameters were also extracted from experiments using a crossed-beam setup, combined with the velocity map imaging (VMI) technique with state-selective and polarization-sensitive detection by (1 + 10) resonance enhanced multipho- ton ionization (REMPI). A novel way of analyzing velocity-map images

Preliminary version – January 29, 2016 Bibliography 147 was developed that directly yields all the parameters that characterize the angular momentum polarization of the scattered, rotationally excited CO molecules. The agreement between experiment and theory manifests the reliability of this new analysis method.

Chapter 7 reports a joint experimental and theoretical study of state-to- state rotationally inelastic DCSs and PDDCSs for CO (v = 0, j = 0, 1, 2) molecules colliding with He. The experiments probing the rotational excita- tion of CO by these collisions were similar to those described in Chapter 6 for CO-Ar collisions. Data were recorded with (2 + 1) REMPI and (1 + 10) REMPI for collision energies of 513 40 cm-1 and 840 40 cm-1, respec- ± ± tively. Also the theoretical work was similar; we used the full quantum close-coupling method in well-converged calculations of the rotationally inelastic DCSs and PDDCSs. Since for CO-He the laser beam is not perpen- dicular to the relative velocity vector, as in the case of CO-Ar, the situation is more complex and an additional parameter is required to characterize the angular momentum polarization of the CO molecules. Still, the com- parison of experimental and calculated results confirmed that also here the experiment and the new analysis method are reliable and provided detailed information on the dynamics of the He-CO collision process.

Chapter 8 shows calculated DCSs for CO colliding with H2 obtained from full close-coupling calculations. Also CO-H2 collisions are astrophysi- cally relevant, just as He-CO and H-CO collisions. We present state-to-state DCSs for rotational transitions of CO from the v = 0, j = 0, 1, 2 initial states -1 to final states with j0 = 3 - 11 at a collision energy at 600 cm . We found that the DCSs for collisions of CO with ortho-H2 and para-H2 are very sim- ilar at this energy. The similarity of the H2-CO DCSs with those of He-CO shows that, contrary to the case of H-CO, it might be a good approxima- tion for astrochemical applications to obtain rate coefficients for H2-CO by scaling the results of He-CO, at least at higher energies.

Preliminary version – January 29, 2016 Preliminary version – January 29, 2016 SAMENVATTING

Dit proefschrift is gewijd aan het kwantummechanische onderzoek van in- elastische botsingsprocessen voor astronomische toepassingen. Bijzondere belangstelling hadden wij voor botsingen van CO moleculen met zijn bots- ingspartners in de ruimte, zoals H, He en H2. Het hoofddoel van het werk was om snelheidsconstanten te bepalen voor rotatie- en ro-vibratieovergangen van CO geïnduceerd door botsingen met waterstofatomen. De resultaten van deze berekeningen zijn te vinden in de Leiden Atomic and Molecu- lar Database (LAMDA) en kunnen worden gebruikt in astrofysische mod- ellen. Verder hebben we gekeken naar CO-He, CO-H2 en CO-Ar botsin- gen. Voor deze systemen hebben we differentiële botsingsdoorsneden en/of polarisatieafhankelijke differentiële botsingsdoorsneden berekend. Door te vergelijken met experimenten met moleculaire bundels konden wij de nau- wkeurigheid van de gebruikte ab initio interactie potentialen toetsen en verkregen wij gedetailleerde informatie over de dynamica van het bots- ingsproces.

Hoofdstuk 1 geeft een introductie in de astrochemische achtergrond van ons werk, de gebruikte theoretische methoden en de moleculaire systemen die bestudeerd worden. Dit hoofdstuk bevat twee subsecties over de theo- retische methoden, de een bevat een inleiding tot methoden gebruikt in elec- tronenstructuur berekeningen, de ander een inleiding tot kwantummecha- nische botsingsberekeningen.

Hoofdstuk 2 beschrijft de constructie van een driedimensionaal ab initio 2 potentiële energie oppervlak voor het H-CO(X A0) complex. De ab initio berekeningen en fit procedure worden in detail beschreven. De potentiaal wordt gebruikt voor de berekening van vibratiefrequentiese en rotatiecon- stanten van H-CO. Vergelijking met experimentele gegevens laat zien dat de nieuwe driedimensionale potentiaal meerdere spectroscopische param- eters nauwkeurig reproduceert. Verder voeren we tweedimensionale bots- ingsberekeningen uit en vergelijken de resultaten met eerdere berekeningen om de nauwkeurigheid van de potentiaal op grote afstand te verifiëren.

Hoofdstuk 3 beschrijft kwantummechanische botsingsberekeningen voor puur rotationele de-excitatie van CO in botsing met waterstofatomen, gebas- eerd op de driedimensionale potentiaal uit Hoofdstuk. 2. Toestandsopgeloste botsingsdoorsneden zijn berekend voor energieën tussen 10-5 en 15 000 cm-1 en de corresponderende snelheidsconstanten zijn bepaald voor tempera- turen tussen 1 en 3000 K. Zowel volledige kwantummechanische gekop- pelde kanalen (close coupling) berekeningen alsook de gekoppelde toes- tanden benadering (coupled-states approximation) zijn gebruikt voor de berekening van CO de-excitatie uit rotatie toestanden met j = 1 - 5, 10, 15, 20,

149

Preliminary version – January 29, 2016 150 Bibliography

25, 30, 35, 40 en 45 in de vibratie grondtoestand, en een schalingsmethode wordt geïntroduceerd voor de tussenliggende rotatie toestanden. Vergelijk- ing met eerdere resultaten van botsingsberekeningen, uitgevoerd door Bal- akrishnan et al. [Astrophys. J., 568, 443 (2002)], laat zien dat de eerdere berekeningen niet volledig betrouwbaar zijn door onnauwkeurigheid van de potentiaal op grote afstand. Bovendien bestrijken onze berekeningen een groter temperatuurgebied en een groter aantal overgangen dan voorheen beschikbaar voor puur rotationele de-excitatie van CO-H, die gebruikt wor- den in astrofysische modellen.

Hoofdstuk 4 behandelt volledig kwantummechanische gekoppelde kan- alen berekeningen voor ro-vibrationele relaxatie van CO in botsing met waterstofatomen, gebaseerd op de driedimensionale potentiaal uit Hoofd- stuk. 2. Eerst testen we de convergentie van de berekening met de kanalen basis en testen we de nauwkeurigheid van de gekoppelde toestanden (cou- pled states) en infinite-order-sudden benadering. Daarna berekenen we een groot aantal botsingsdoorsneden voor ro-vibratieovergangen van v = 1, j = -1 0 - 30 naar v0 = 0, j0 bij botsingsenergieën tussen 0.1 and 15 000 cm . Bots- ingsdoorsneden voor pure vibratieovergangen v = 1 v = 0 zijn berekend ! 0 met volledige gekoppelde kanalen berekeningen alsook in de infinite-order- sudden benadering. Vergelijking van deze gegevens laat zien dat de infinite- order-sudden benadering puur vibrationele overgangen nauwkeurig beschri- jft. De berekende snelheidsconstanten voor v = 1 v = 0 zijn ook in goede ! 0 overeenstemming met de metingen van Glass et al. [J. Phys. Chem., 86, 908 (1982)].

Hoofdstuk 5 bevat snelheidsconstanten voor ro-vibratieovergangen door CO-H botsingen bij temperaturen tussen 10 en 3000 K, berekend uit de botsingsdoorsneden uit Hoofdstuk. 4. Verder zijn snelheidsconstanten voor puur vibrationele overgangen met v = 1 - 5 v

Hoofdstuk 6 beschrijft de experimentele en theoretische studie van CO- Ar botsingen. Differentiële botsingsdoorsneden (DCS voor differential cross section) en polarisatieafhankelijke differentiële botsingsdoorsneden (PDDCS

Preliminary version – January 29, 2016 Bibliography 151 voor polarisation-dependent DCS) worden berekend met de volledig kwan- tummechanische gekoppelde kanalen methode. Deze fysisch relevante pa- rameters worden ook bepaald door experimenten met gekruiste molec- ulaire bundels, gecombineerd met de velocity map imaging (VMI) tech- niek met toestandsselectieve en polarisatiegevoelige detectie door (1 + 10) resonance enhanced multiphoton ionization (REMPI). We passen een nieuwe analyse van de velocity-map images toe waarmee direct alle parameters worden bepaald die de polarisatie van het impulsmoment van de gebot- ste en rotationeel geëxciteerde CO moleculen beschrijven. De overeenstem- ming tussen experiment en theorie bevestigd de betrouwbaarheid van deze nieuwe analysemethode.

Hoofdstuk 7 beschrijft de theoretische en experimentele studie van toe- standsopgeloste DCSs en PDDCSs voor CO (v = 0, j = 0, 1, 2) moleculen in botsing met heliumatomen. De experimenten waarmee CO moleculen gemeten worden die door de botsing rotationeel zijn geëxciteerd is vergeli- jkbaar met wat in Hoofdstuk. 6 is beschreven voor CO-Ar botsingen. De detectie is gebaseerd op (2 + 1) REMPI en (1 + 10) REMPI voor botsingsen- ergieën van respectievelijk 513 40 en 840 40 cm-1. Het theoretische werk ± ± was eveneens vergelijkbaar; wij gebruikte goed geconvergeerde en volledig kwantummechanische gekoppelde kanalen berekeningen van rotationeel inelastische DCSs en PDDCSs. Echter is de situatie in dit hoofdstuk in- gewikkelder omdat, in tegenstelling tot de experimenten in Hoofdstuk 6, de laser in dit geval niet loodrecht staat op de relatieve snelheidsvector. Hi- erdoor is een extra parameter vereist om de polarisatie van het impulsmo- ment van de gebotste CO moleculen te kunnen beschrijven. Ook hier laat de vergelijking van experiment en theorie de betrouwbaarheid van de analyse methode zien, en levert dit gedetailleerde informatie op over de dynamica van CO-He botsingen.

Hoofdstuk 8 laat de berekende DCSs voor CO in botsing met H2 zien, verkregen uit gekoppelde kanalen berekeningen. Evenals de eerder beschre- ven botsingscomplexen zijn CO-H2 botsingen astrofysisch relevant. Wij geven toestandsopgeloste DCSs voor rotatieovergangen van CO uit begintoestanden met v = 0, j = 0, 1, 2 naar eindtoestanden met j0 = 3 - 11, bij een botsingsen- ergie van 600 cm-1. Bij deze energie wordt er weinig verschil gevonden tussen de DCSs voor botsingen van CO met ortho-H2 en para-H2. De geli- jkenis met de DCSs voor CO-H2 en CO-He suggereert dat men bij - voor as- trochemische toepassingen - nauwkeurige benadering de snelheidsconstan- ten voor CO-H2 kan verkrijgen door de resultaten voor CO-He te schalen, in ieder geval voor deze hogere energieën.

Preliminary version – January 29, 2016 Preliminary version – January 29, 2016 X Å

,∫áÑ;ÅÂ\/⇢«œPõf°óπ’v)áS˚-Ñ^9'∞û« ↵⇥⌘Ït£Ñ/'≥CO ⌃+zÙ-Ñ"üPH &He å"⌃PH2 KÙ—Ñ^9'∞û«↵⇥v-@Õ°óÜ"üP '≥H-CO ∞ûS˚-Øl®-/l-Ñ^9'Ô⌃c⌅*b åá8p⇥⌘ÏÑ°ó”ú´±üP⌃PpnìLAMDA 6Uv ⌃î(0)ái⌃Ñ!fl-⇥⌘Ïÿ€evÜ&'≥He-CO "⌃P'≥H2-CO )'≥Ar-CO Ñ^9'∞û«↵ °óÜ(yö∞û˝œ↵ŸõS˚ÑÆ⌃*båw IO/ ÈÑÆ⌃* b⇥yöÑ∞û˝œ/⌘Ï\⇧Ñûåaˆ¯;Ñ⇥(⌘ÏÑ°ó” úûåKœ”ú€L˘gÔ¿å˙1Œ4ó’ó˙ѯí\(ø˝bÑ æn¶Â ∞û«↵-ÑÊ∆Ñ®õf·o⇥

,‡⇢ÄÀœÜ⌘ÏÂ\Ñ)áfÃo–(Ñ⌃∫π’å@ vS˚Ñ∞∂⇥v-⌃∫π’⌃$*Ë⌃,Ë⌃ÀÕ5P”Ñ°ó ,åË⌃ÀÕœPc⌅⌃∫⇥

2 ,å‡⇢–(Œ4ó’Ñ˙Ü"üPå'≥H-CO X A0Ñ Ù ø˝b⇥ ⌘ÏÊ∆œÜŒ4°óåø˝bÑfl«↵⇥ –(∞˙Ñø˝ b⌘Ï°óÜ"üPå'≥H-CO Ñ/®ëáåle®8pvä Ÿõpnûå∑óÑpn€L‘É—∞∞˙ø˝bÔÂæn0Õ∞˙' œÑûåI1¬p⇥Ê⌘ÏgLÜõeÑåÙc⌅°óv⌃°ó ”úM∫Ñ”ú€L‘ÉŒó˙∞˙ø˝bw ænÑ↵Ë⌃⇥

, ‡⇢˙é,å‡Ñ˙Ñ ÙH-COø˝b⌃∫vÜ'≥ CO ´"üPH ∞û¸ÙÑØl®¿—«↵⇥(10-5015000 cm-1∞ û˝œ↵⌘Ï°óÜ-^9'∞û*bvó0¯˘î)¶1 - 3000 K↵ Ñ^9'∞ûá8p⇥⌘Ï«(ÜœPõfclose-couplingπ’åcoupled states—<π’vÜ⌅é/®˙v = 0 Ñ'≥CO Œl®j = 1 - 5, 10, 15, 20, 25, 30, 35, 40, 45Ñ¿—«↵˘éj<45Ñv÷l®Ñ ¿—«↵,⌘Ï«(ÜÕ ‘ã“<Ñπ’⇥M∫BalakrishinanI[As- trophys. J., 568, 443 (2002)]Ñc⌅°ó”ú‘É⌘Ï—∞÷Ï«(Ü↵ Ë⌃æn'Ñø˝b¸ÙÜv°ó”úåhÔ`⇥Ê:Ü)ái ⌃!fl⌘Ï°óÜÙ')¶⇤Ù↵åÙ⇢√¡*pÑH-COØl®^9' ∞ûá8p

,€‡⇢˙é,å‡Ñ ÙH-COø˝b–(close-couplingπ’°óÜ ÂS˚Ñ/®l®√¡«↵⇥ ñH⌘ÏK’Ü⇢S˙ƒchannel basis set Ñ6[' coupled statesåinfinite-order-sudden —<π’Ñæn ¶⇥6⌘Ï°óÜ'y∞û˝(0.1015000 cm-1KÙÑ-^9' /®l®∞û*b⇧ÏÜŒv = 1, j = 0 - 300»v0 = 0, j0Ñ√ ¡⇥Ê⌘Ï⌃+«(close-couplingπ’åinfinite-order-sudden—<π

153

Preliminary version – January 29, 2016 154 Bibliography

’°óÜØ/®√¡v = 1 0↵Ñ∞û*bå¯î)¶↵Ñá8p⇥⇢ ! «‘ÉÂ⌦p<—∞1infinite-order-sudden—<π’@ó0Ñá8p ÔÂclose-couplingπ’ó0Ñ”ú›ÿ¶Ù⇥ ˆŸõ°óp< GlassI∫J. Phys. Chem., 86, 908 (1982) ÑûåKœ”ú¯S;⇥

,î‡⇢⇢«˘,€‡Ñ∞û*bÑÌsGó0Ü(10 - 3000 K)¶ ↵H-COS˚Ñ^9'∞û/l√¡á8p⇥ˆ⌘Ï(infinite-order- sudden—<π’°óÜØ/®√¡v = 1 - 5 v

,m‡⇢)(°óåûåvÜAr-CO^9'∞û«↵⇥ ⌘Ï–(å hÑœPclose-couplingπ’°óÜÆ⌃∞û*båIO/ ÈÑÆ⌃ ∞û*b⇥ ˆ)(”ªP¶⇣œVMI Ä/å˘IO/OÑ 1 + 10 q/†:⇢IP5ªREMPI Ä/Ñ⌃P_ûå≈nŒûå˛ œ-–÷Ü°ó”ú¯˘îÑi⌃œ⇥,‡ÇÿÀÕÜÕ∞ñÑ⌅⌃ ¶˛œÑπ’⇥)(ŸÕπ’ÔÂÙ•–÷∞ûÕ '≥CO “ ®œ¿—Åy'Ñi⌃œ⇥ûåå⌃∫pnÑÙ'U:ÜŸÕ∞ñ⌃ê π’ÑÔ`'⇥

,⇤‡⇢•Sܢ'≥CO (v = 1, j = 0, 1, 2) &He -Æ ⌃*bå˘IO/ ÈÑÆ⌃*bÑûå⌃∫v⇥,m‡Ar-COS˚¯ <ûå¢KÜ1é^9'∞û¸ÙÑ'≥CO l®¿—⇥12 + 1 å1 + 1 REMPI∞UÜ∞û˝⌃+:513 40 cm-1å840 40 cm-1Ñ 0 ± ± ûåpn⇥⌃∫Â\_,m‡¯<⌘Ï«(hœPclose - couplingπ ’°óÜl®^9'Æ⌃*bå˘IO/ ÈÑÆ⌃∞û*b⇥1éHe- COS˚Ar-COS˚ûåÑ¿I˝S˚ѯ˘¶‚œÇÙ⇥ ‡d˘éHe-COS˚eÙ≈µÙ†BÅ*ùѬpeœ' ≥CO “®œÅyÅ⇥ûåå⌃∫°ó”ún§Üûåå⌃êπ’ /Ô`ш–õÜHe-CO∞û«↵Ñ®õf·o⇥

,k‡⇢U:Ü1åhÑclose-couplingπ’°óó0ÑH2-COS˚Ñ^ 9'Æ⌃∞û*b⇥ åHe-CO H-CO7H2-COÑûå⌃∫v˘) ái⌃ ÕÅј<⇥ ⌘ÏŸ˙ÜŸ*S˚(600 cm-1∞û˝œ↵Œ v = 0, j = 0, 1, 2√¡0»j0 = 3 - 11Ñl®√¡Æ⌃*b⇥⌘Ï—∞( Ÿ*∞û˝œ↵'≥CO c"(ortho-H2 ∞ûåÚ"para- H2 ∞û@ó0ÑÆ⌃*b^8¯<⇥H-COS˚H2-CO∞ûS˚ ÑÆ⌃*bHe-COÅ:¯<⇥1d⌘Ï®K(‘ÉÿÑ∞û˝œ↵⇢ «˘He-COÑá8p€L‘ã)>Ô—<ó0H2-COS˚Ñ”ú⇥

Preliminary version – January 29, 2016 LIST OF PUBLICATIONS

1. Lei Song, Gerrit C. Groenenboom, Ad van der Avoird, Chandan Kumar Bishwakarma, Gautam Sarma, David H. Parker, Arthur G. Suits In- elastic Scattering of CO with He: polarization dependent differential state-to- state cross sections, J. Phys. Chem. A, 119, 12526 (2015)

2. Lei Song, N. Balakrishnan, K. M. Walker, P. C. Stancil, W. F. Thi, I. Kamp, A. van der Avoird and G. C. Groenenboom, Quantum calcula- tion of inelastic CO collisions with H. III. Rate coefficients for ro-vibrational transitions, Astrophys. J., 813, 96 (2015)

3. Lei Song, N. Balakrishnan, Ad van der Avoird, Tijs Karman, and Gerrit C. Groenenboom, Quantum scattering calculations for ro-vibrational de-excitation of CO by Hydrogen atoms, J. Chem. Phys., 142, 204303 (2015)

4. Kyle M. Walker, Lei Song, B.H. Yang, G. C. Groenenboom, A. van der Avoird, N. Balakrishnan, R. C. Forrey, P. C. Stancil, Quantum calcula- tions of inelastic CO collisions with H. II. Pure Rotational Quenching of High Rotational Levels, Astrophys. J. 811, 27 (2015)

5. Arthur G. Suits, Chandan Kumar Bishwakarma, Lei Song, Gerrit C. Groenenboom, Ad van der Avoird, David H. Parker, Direct Extraction of Alignment Moments from Inelastic Scattering Images, J. Phys. Chem. A, 119, 5925 (2015)

6. Lei Song, Ad van der Avoird, and Gerrit C. Groenenboom, Three- 2 dimensional ab initio potential energy surface for H-CO(X A0), J. Phys. Chem. A, 117, 7571 (2013) e

155

Preliminary version – January 29, 2016 Preliminary version – January 29, 2016 CURRICULUM VITAE

Name Lei Song Date of birth 5 January, 1986 Place of birth Qingdao, China

2011-2015 Ph.D., Theoretical Chemistry, Radboud University Nijmegen 2008-2011 M.Sc., Physical Chemistry, University of Science and Technology of China Thesis title: "Theoretical investigation of the reaction mechanism of Hydroxide anion with chlorofluoromethane" 2004-2008 B.Sc., Applied Physics, Shandong Normal University, China

157

Preliminary version – January 29, 2016 Preliminary version – January 29, 2016 ACKNOWLEDGEMENTS

Closing, I would like to express my sincerest thanks to all the people who have helped me directly or indirectly throughout my Ph.D period. Without your help, this thesis would not exist in its current form. My deep gratitude goes first to my supervisor Gerrit. You gave me this valuable chance to study abroad for a meaningful astrochemistry project. You spent plenty of time and efforts teaching me from the basic theory to the practical algorithm. Your guidance and suggestions have been always effective and efficient to solve my encountered problems. Besides science, the improvement of my listening and oral English has benefited from the frequent discussions with you. In addition, you encouraged me to give pre- sentations in conferences and summer/winter schools, which proved to be helpful for job hunting. Ad, my second supervisor, I appreciate your scientific guidance and care- ful modifications on my publications and dissertation. You were also my office mate throughout the whole 4 years of my Ph.D period. It has been very nice with your company in the office. When I had questions, it was super-cool to get your prompt guidance just by turning my head to your direction and asking. It was funny when we met in front of the door, you, as a gentleman, always opened the door and let me come in first. That is very different from the situation in China, where the professor has the priority. Next, I want to thank the rest of the Theoretical Chemistry group. The first one is Herma. I appreciate your modifications on the astrochemistry background of this thesis and helpful guidance on my presentations. I also enjoyed our short talks when you were waiting for coffee in my office. Then, I want to thank another two office mates Eduardo and Tijs. It was very nice to share the office with you. Sometimes, you together prepared a surprise for me, which reminds me of the joyful time we spent. Respec- tively, thank Eduardo for lunchtime conversation; thank Tijs for writing Dutch summary for this thesis and scientific help. There are a lot of people who made the great atmosphere in our group. Thanks to Adrien, Boy, Den- nis, Engin, Gilles, Jolijn, Joost, Koos, Liesbeth, Leendertjan, Mark, Michal, Pegah, Peter, Roy, Sasha, Simon, Thanja, Xander and all the others. I also would like to thank my collaborators: Dave, Chandan, Gautam, Inga, Wing-Fai, Kyle, Prof. Arthur G. Suits, Prof. Balakrishnan Naduvalath and Prof. Phillip C. Stancil. With your valuable discussions and academic support, I finished my Ph.D work successfully. Besides, I would like to thank Ewine for suggesting this interesting study of the H+CO system and I thank you and François Lique for valuable discussions. It is also impor- tant to thank the people who work in the computer and communication department (C&CZ) of the Faculty of Science for computer resources and technical support.

159

Preliminary version – January 29, 2016 160 Bibliography

My friends’ company made my life more colorful. I want to thank Rudy, Lantian Chang, Ying Du, Juehan Gao, Zhi Gao, Huiyan Huang, Damei Ke, Shulai Lei, Shujuan Li, Xiaohu Li, Huaifeng Liu, Yujie Ma, Jing Shen, Zhongfa Sun, Zhipeng Wang, Chengzhi Xu, Jialiang Xu, Shengjun Yuan, Liang Zhou, and all the others. Finally, thanks to my family: my parents and other relatives. Special thanks to my husband, Mingli Niu. Thanks for commuting every work day from Nijmegen to Amsterdam. Thanks for taking care of me in daily life. Without your support and company, I would not live happily abroad.

Preliminary version – January 29, 2016