isomer-specific kinetics of the c + h o reaction at the … + h2o... · 2020. 11. 9. ·...

40
Isomer-specific kinetics of the C + + H 2 O reaction at the temperature of interstellar clouds Tiangang Yang, 1,6# Anyang Li, 2# Gary Chen, 1 Qian Yao, 3 Arthur G. Suits, 4 Hua Guo, 3 Eric R. Hudson 1,5 and Wesley C. Campbell 1,5 1 Department of Physics and Astronomy, University of California Los Angeles, Los Angeles, California 90095, USA 2 Key Laboratory of Synthetic and Natural Functional Molecule Chemistry, Ministry of Education, College of Chemistry and Materials Science, Northwest University, 710127 Xi’an, P. R. China 3 Department of Chemistry and Chemical Biology, University of New Mexico, Albuquerque, New Mexico 87131, USA 4 Department of Chemistry, University of Missouri, Columbia, Missouri 65211, USA 5 UCLA Center for Quantum Science and Engineering, University of California – Los Angeles, Los Angeles, California 90095, USA 6 Department of Chemistry, Southern University of Science and Technology, Shenzhen 518055, China # These authors contributed equally to this work. ABSTRACT: The reaction C + +H2O→HCO + /HOC + +H is one of the most important astrophysical sources of HOC + ions, considered a marker for interstellar molecular clouds exposed to intense UV or X-ray radiation. Despite much study, there is no consensus on rate constants for formation of the formyl ion isomers in this reaction. This is largely because of difficulties in laboratory study of ion-molecule reactions under relevant conditions. Here, we use a novel experimental platform combining a cryogenic buffer-gas beam with an integrated,

Upload: others

Post on 22-Mar-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

Isomer-specific kinetics of the C+ + H2O reaction at the temperature

of interstellar clouds

Tiangang Yang,1,6# Anyang Li,2# Gary Chen,1 Qian Yao,3 Arthur G. Suits,4 Hua Guo,3 Eric

R. Hudson1,5 and Wesley C. Campbell1,5

1 Department of Physics and Astronomy, University of California Los Angeles, Los Angeles,

California 90095, USA

2 Key Laboratory of Synthetic and Natural Functional Molecule Chemistry, Ministry of

Education, College of Chemistry and Materials Science, Northwest University, 710127

Xi’an, P. R. China

3 Department of Chemistry and Chemical Biology, University of New Mexico, Albuquerque,

New Mexico 87131, USA 4 Department of Chemistry, University of Missouri, Columbia, Missouri 65211, USA

5 UCLA Center for Quantum Science and Engineering, University of California – Los Angeles, Los Angeles, California 90095, USA 6 Department of Chemistry, Southern University of Science and Technology, Shenzhen 518055, China # These authors contributed equally to this work.

ABSTRACT: The reaction C++H2O→HCO+/HOC++H is one of the most important

astrophysical sources of HOC+ ions, considered a marker for interstellar molecular clouds

exposed to intense UV or X-ray radiation. Despite much study, there is no consensus on rate

constants for formation of the formyl ion isomers in this reaction. This is largely because of

difficulties in laboratory study of ion-molecule reactions under relevant conditions. Here, we

use a novel experimental platform combining a cryogenic buffer-gas beam with an integrated,

Page 2: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

laser-cooled ion trap and high-resolution time-of-flight mass spectrometer to probe this

reaction at the temperature of cold interstellar clouds. We report a reaction rate constant of

𝑘𝑘 = 7.7(6) × 10−9 cm3s-1 and a branching ratio of formation 𝜂𝜂 = HOC+/HCO+ = 2.1(4).

Theoretical calculations suggest that this branching ratio is due to the predominant formation

of HOC+ followed by isomerization of products with internal energy over the isomerization

barrier.

Page 3: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

INTRODUCTION:

Stars and their accompanying solar systems are born from dense, cold molecular clouds

that host a surprisingly rich and complex chemistry (1). This chemistry is largely driven by

intense ultraviolet (UV) radiation from nearby massive young stars in so-called Photon

Dominated Regions (PDRs) on the surface of such clouds (2), by cosmic rays that penetrate

deep into the clouds in regions opaque to UV light, or in some cases by X-ray dominated

regions (XDRs) from the circumnuclear disks (CND) of active galactic nuclei (AGN)

containing supermassive black holes (3, 4). This complex chemistry then seeds newly formed

solar systems and the process continues in successive generations of star formation and demise.

The chemistry within these clouds sensitively reflects the local conditions of radiation

exposure, temperature, density, structure and dynamics. With the advent of interferometric

radio telescopes such as NOEMA and ALMA and others that can image specific molecules at

unprecedented spatial resolution (5, 6), the broad-brush view of these diverse environments is

giving way to rich detail for probing and cataloging their structure, composition, and local

conditions, even in distant galaxies. An understanding of the associated chemistry is essential

to build accurate models and also identify species that are reliable tracers for conditions such

as UV or X-ray irradiation. Laboratory experiments that can investigate specific reactions

under conditions resembling those in low-density interstellar clouds at 10-50 K are essential to

develop such models and identify suitable reporter molecules, complementing the enormous

investment in the observational tools.

Advanced experimental techniques developed in atomic physics for preparation of

ultracold matter and quantum information science have recently been adapted to chemical

Page 4: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

studies. While these techniques allow preparation and coherent control of a reagent in single

internal and external quantum states, they have so far been limited to a narrow range of species,

e.g. the alkali atoms. As a result, despite groundbreaking work that has assembled single

molecules (7), observed a conformer-dependent reaction (8) and produced novel molecules (9),

no study to date has adapted these techniques to a reaction where all reagents are chemically

relevant. Here, we develop a novel experimental platform based on these atomic physics

techniques that allows studies of reactions of virtually any atomic ion with a wide range of

neutral molecules. Specifically, we combine a laser-cooled ion trap with integrated mass

spectrometer and cryogenic buffer gas beam (Fig. 1(A)). These tools have been independently

used in a number of pioneering studies (8-12), but they have typically been restricted to laser-

cooled atomic reagents and produce reagents at temperatures significantly colder than the

interstellar medium (ISM). In this work we adapt them to produce C+ and H2O at more

characteristic temperatures (≈20 K) to properly simulate cold interstellar clouds in the

laboratory.

Using this platform, we study the reaction C+ + H2O → HOC+/HCO+ + H with product

isomer specificity under conditions relevant to those in cold molecular clouds. This reaction is

believed to be the chief source of the metastable ion HOC+, which has been observed in a

variety of sources including of nearby PDRs (13), toward our own galactic center (14, 15), in

the early starburst galaxy M82 (16, 17), a later starburst galaxy NGC 253 (18), in the

ultraluminous infrared galaxy Mrk273 (19), and in the AGNs of NGC 1068 and other active

galaxies. The lower energy isomer HCO+, on the other hand, is largely formed via protonation

of CO by the ubiquitous ion H3+, and it is generally far more abundant. The relative abundance

Page 5: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

of HCO+ to HOC+ is seen to vary by many orders of magnitude, from well over 104 in the

weakly irradiated PDR S140 (13) to 10 in the inner CND of Mrk273 (19). One question is,

to what extent can HOC+ serve as a specific marker for XDRs (3)? A recent survey toward two

quiescent cloud complexes of the galactic center, one of which is likely exposed to strong X-

ray radiation, suggested that the CS:HOC+ ratio was a reliable marker for PDR/XDR

components in galactic nuclei exposed to large scale shocks, but there was no molecular tracer

useful to distinguish PDR and XDR regions (20). For models to be used reliably to reveal

conditions in distant molecular clouds in these very diverse environments, the yield of the

associated reactions must be accurately known. These relative abundances obviously depend

on the product branching ratios of the underlying chemical reactions that produce these

molecules (13, 21). For the title reaction, the product channels are:

C+ + H2O → HOC+ + H (1)

C+ + H2O → HCO+ + H (2)

As such, the rate and branching ratio, 𝜂𝜂 = HOC+/HCO+, of this reaction has attracted

significant interest over the past 30 years; however, there is still no consensus on the latter

value and it has never been measured at low temperature (22-26). Freeman and coworkers

investigated this reaction with a selected-ion flow tube (SIFT) at room temperature and

reported 𝜂𝜂 ≈ 5(2) by using the different proton transfer rates of the isomers with N2O to

distinguish HCO+ from HOC+ (22). Bierbaum and coworkers studied the reaction more

recently using the flowing-afterglow-SIFT technique to eliminate the contribution of

electronically excited C+ ions but did not measure branching and quite reasonably assigned all

the products to HOC+ (27). Sonnenfroh et al. carried out a crossed beam experiment to study

Page 6: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

this reaction at collision energies of 0.62 eV and 2.14 eV, in which they concluded that 𝜂𝜂 ≤

2.3 and pointed out that the high internal temperature of the products due to the exothermicity

of the reaction (Δ𝐻𝐻 = −4.34 eV ) may affect the identification of the isomers in subsequent

titration reactions (25). In theoretical calculations, Defrees, McClean and Herbst used phase-

space arguments to calculate 𝜂𝜂 = 2 (28). However, using quasi-classical trajectories (QCT)

studies, Ishikawa, Binning, and Ikegami found that HOC+ dominated the initial products of the

reaction giving 𝜂𝜂 ≈ 100 (29). The authors discussed the possibility of subsequent

isomerization due to internal energy of the products to explain the existence of HCO+ in

experimental observations, however, no calculations for the isomerization rate were performed.

Despite these disagreements, the room-temperature measurement of 𝜂𝜂 (22) has been widely

employed to understand ISM chemistry and classify PDRs for decades (13).

Here, we build on pioneering work of Gerlich, Schlemmer, Rowe, and others (30-32) and

report the first study of the C+ + H2O reaction under conditions present in cold molecular clouds

of the ISM. We find the reaction proceeds with a rate constant of 𝑘𝑘 = 7.7(6) × 10−9 cm3s-1 at

≈ 20 K, slightly lower than the value reported at 27 K (26). Titrating the reaction products with

15N2, which reacts with HOC+ but not HCO+, we measure the ratio of HOC+ to HCO+ exposed

to the titrant to be 1.4(2). Accounting for isomerization during the titration process, we report

𝜂𝜂 = 2.1(4). QCT calculations on an accurate potential energy surface (PES) found that the C+

+ H2O reaction produces 97.6% HOC+, of which 87.8% of these products are produced with

an internal energy above the isomerization barrier. We calculate that the density of HOC+

vibrational states is twice that of HCO+ near the isomerization barrier. Therefore, we expect

that roughly 2/3 of the above-barrier products decay to HOC+, which modifies the predicted

Page 7: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

branching ratio to 𝜂𝜂 = 2.3, in agreement with experiment.

Materials and Methods:

A schematic of the apparatus used in this study is shown in Fig. 1(A). Briefly, an integrated

laser-cooled ion trap and time-of-flight mass spectrometer (33), which has been described in

detail elsewhere (34, 35), is combined with a cryogenic buffer gas beam (CBGB) (18). Laser

ablation of beryllium and graphite is used to produce Be+ and C+ ions simultaneously, which

are trapped in a linear radio frequency Paul trap. Laser cooling (36) is used to cool the

translational motion of the Be+ ions, which sympathetically (37, 38) cools the co-trapped C+

ions to <1 K. This process takes several seconds; during that time, any metastable (4P) C+,

which may have plagued other measurements as explained in Ref. (27), spontaneosly relaxes

(39). Water vapor is entrained in the neon CBGB (40), which is cooled by a closed-cycle

refrigerator to 20.0(5) K, and directed towards the center of the ion trap with a forward

velocity of 150(2) m/s (see supplementary information, SI) to facilitate C+ + H2O reactions.

The reactions occur inside the ion trap, which is held in a high vacuum chamber (<5 × 10−10

mbar) (34, 35), and the trapped ionic reactants and products are analyzed by an integrated time-

of-flight (TOF) mass spectrometer (TOF-MS)(9, 41, 42).

Page 8: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

Figure 1. (A) A schematic diagram of the experimental setup. (B) TOF trace (10 sample average) of Be+ and C+

exposed to water from the CBGB and 15N2 from the leak valve (10 s) at a density of 1 × 109 cm−3. (C) The

fraction of the titrated isomers saturating as a function of 15N2 density. Fitted parameters yield a ratio of trapped

reaction products of 1.4(2) and 𝑘𝑘3 = �(6.6 ± 1.0) × 10−10� cm3/s. (D) Same as (B), but with 20 s of C+ +

H2O reaction and isomerization via CO introduced as a background gas before titration with N2. The lack of a

peak for 15N2H+ indicates that the CO has converted all of the HOC+ to HCO+ and that HOC+ does not react with

N2.

Two isomers of the formyl ion, HCO+ and HOC+, are produced in this reaction; however,

TOF-MS is unable to distinguish these ions as they have the same mass and charge. In order to

titrate the HCO+ distinctly from the HOC+, a third gas with proton affinity between the two

isomers is introduced. Specifically, N2 is chosen due to its low reactivity with other trapped

species, instead of N2O or other gases with suitable proton affinity (22) (see SI), to simplify

Page 9: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

the analysis of the titrated products. The titration reactions of interest are:

HOC+ + N2 → N2H+ + CO, (3)

HCO+ + N2 → No reaction. (4)

Unfortunately, the mass of diazenylium (14N2H+), 𝑚𝑚𝐷𝐷 = 29.01 amu, is nearly equal to that of

the formyl ion, 𝑚𝑚𝐹𝐹 =29.00 amu and these products cannot be resolved by our mass

spectrometer. Therefore, we use 15N2 as a titrant gas to produce diazenylium with a mass of

31.01 amu, which is easily resolved by our mass spectrometer. Reaction (4) has been validated

experimentally, shown in Fig.1(D).

The experimental sequence is described in the SI. Briefly, trapped C+ ions are

sympathetically cooled by the co-trapped Be+ ions. Next, the H2O beam is introduced into the

trap for 20 s and reactions occur. After an approximately one second delay to allow the

internally excited product ions to isomerize and/or relax via radiative decay (43) the 15N2 titrant

gas is introduced for approximately 10 s. Afterwards, all ions are ejected into the TOF-MS and

Fig. 1(B) shows a typical TOF signal. From these data, we find the reaction rate constant 𝑘𝑘1 +

𝑘𝑘2 = 7.7(6) × 10−9 cm3s-1 (See SI.3 for the details).

Several other species, not studied here, are also apparent in the TOF-MS signal. These are

understood as follows. Be+ ions are known to react with H2O to make BeOH+ and a small

amount of H3O+ (34, 35). The majority of H3O+ comes from the reactions:

HCO+ + H2O → H3O+ + CO, (5)

HOC+ + H2O → H3O+ + CO. (6)

The rate coefficient of reaction (5) has been measured as (2.5 ± 0.3) × 10−9 cm3/s at 300 K

(44), but not at lower temperature. Though the appearance of H3O+ presents no problem for

Page 10: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

our study, a relative rate difference between these two reactions ((5) and (6)) at low temperature

would systematically shift the trapped HOC+/HCO+ ratio and cause an error in the

determination of branching ratio 𝜂𝜂. To constrain this potential error, we introduced CO to

produce samples with varying compositions of HCO+ and HOC+ (details can be found in SI)

from the reaction:

HOC+ + CO→ HCO+ + CO. (7).

We find 𝑘𝑘5 ≈ 𝑘𝑘6 = (1.7 ± 0.2) × 10−8cm3/s and therefore the HOC+/HCO+ ratio is not

measurably affected.

Therefore, since isomerization of the isoformyl ion by H2O does not occur (45) (see SI),

the branching ratio of HOC+ and HCO+ can be inferred from the ratio of the integrated m/z =

31.01 amu/e (15N2H+) and m/z = 29.00 amu/e (HCO+) peaks in the TOF signal, which is found

to be 1.4(2). Repeating this process over various densities (Fig. 1(C)) of 15N2 allows us to

confirm this measurement as well as determine the proton transfer rate constant of reaction (3)

to be 𝑘𝑘3 = 6.6(1) × 10−10cm3/𝑠𝑠 (see SI).

In addition, to verify that 15N2 is only able to react with HOC+ and not HCO+. We repeat

the above process in the presence of a large amount of neutral CO gas for converting all of the

HOC+ to HCO+ via the isomerization reaction (7), and then introduce excess 15N2. Since there

is no evidence of m/z = 31.01 amu/e (15N2H+) in the TOF trace (Fig. 1(D)), reactions (3) and

(4) are validated.

Results:

Before the branching ratio 𝜂𝜂 can be extracted from the measured product ratios, it is

necessary to account for isomerization by the titrant gas, i.e.

Page 11: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

HOC+ + 15N2 → HCO+ + 15N2. (8)

As HCO+ is unreactive to 15N2 at these temperatures, this reaction artificially lowers the

measured branching ratio. To understand this effect, quasi-classical trajectory (QCT)

calculations (46) on an long-range corrected ab initio potential energy surface for the HOC+ +

N2 system were performed (47). These calculations yield a 14% isomerization (see SI) fraction.

The isomerization fraction can be constrained with the assumption that the total rate constant

for HOC+ + 15N2 is given by the Langevin capture rate constant, 𝑘𝑘L = 8.0 × 10−10 cm3/s.

Thus, the measured value of 𝑘𝑘3 suggests a possible isomerization fraction of up to 18%. Using

the isomerization rate found in the QCT calculations, the measured reaction branching ratio

becomes 𝜂𝜂 = 2.1(4).

Discussion:

To better understand this reaction, a six-dimensional global PES for reactions (1) and (2)

was constructed from the explicitly correlated unrestricted coupled cluster singles, doubles,

and perturbative triples [UCCSD(T)-F12a] calculations (details of ab initio calculations and

fitting are given in SI). The energetics for reactions (1) and (2) are shown in Figure 2. As C+

approaches water, it can be transiently trapped in the IM1 well, in which C is bonded to O.

Given the system energy, however, the intermediate is relatively short lived and not

significantly affected by the two transition states. Due to the large frequency mismatch among

its vibrational modes, the vibrational energy is hardly randomized in the intermediate. In most

cases, the large energy in the C-O mode is quickly transferred to a OH bond, leading to its

cleavage and resulting in the HOC+ product. In a small number of trajectories, however, the

energy transfer is not sufficient to dissociate the H, but allows the H to undergo large amplitude

Page 12: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

motion around the HOC moiety, leading to the formation of HOCH intermediates (IM3, IM4,

and IM6). Eventually, the dissociation of the O-H bond results in the formation of HCO+. Two

exemplary trajectories are illustrated in Figure S7. The large HOC+/HCO+ branching ratio is a

testament of non-statistical nature of the reaction.

The QCT calculations performed on the PES using the experimental conditions with a total

of 48656 trajectories found that the thermal rate constant and branching ratio are 𝑘𝑘1 + 𝑘𝑘2 =

(5.02 ± 0.03) × 10−9cm3/s and 𝜂𝜂 = 40.5 ± 2.0, respectively. While the rate constant is in

reasonable agreement with the measured value, the branching ratio is ≈ 20 times larger.

However, because reaction (1) and (2) are quite exothermic 87.8% (29.6%) of the HOC+

(HCO+) is produced with internal energy above the isomerization barrier. As a result, as the

product molecules radiatively cool, 𝜂𝜂 can change dramatically. Since this cooling proceeds via

vibrational decay the isomerization fraction can be estimated via the density of vibrational

states of each isomer near the barrier (see SI). Due primarily to the lower vibrational frequency

of the HOC+ bending mode, it exhibits ≈ 2 × higher density of states near the isomerization

barrier, indicating that roughly 2/3 (1/3) of the above-barrier products will decay to HOC+

(HCO+). Thus, the radiative cooling process lowers the QCT-calculated branching ratio to 𝜂𝜂 ≈

2.3, in agreement with the experimental value.

Page 13: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

Figure 2. Energetics of the ground state reaction pathways for the C+ + H2O reaction. The relative energies are

marked, along with the ZPE corrected values (italic and underlined) in kcal/mol

Thus, the agreement of the present measurement of the branching ratio with that predicted by

phase-space theory (28) is potentially misleading. The reaction is not statistical, but direct and

is followed by radiative cooling that alters the end product ratio substantially. This view is

consistent with crossed molecular beam experiments from Farrar and coworkers at high

collision energies (25). They mapped the product velocity distributions on a microsecond time

scale and noted that the angular distribution was asymmetric, suggesting a lifetime for the

transient intermediate on the order of a rotational period, ≈10-13 s. We note one of the unique

features of the present approach as compared to previous work on the C+ + H2O reaction is the

capability to permit the products to relax fully prior to detection, as would occur in low density

conditions in the ISM.

Summary

Page 14: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

We report the use of a recently developed tool, a cryogenic buffer gas beam coupled to an

integrated ion trap and mass spectrometer, to simulate conditions in the cold interstellar

medium in the laboratory. We use this tool to study the C+ + H2O → HCO+/HOC+ + H reaction,

finding a reaction rate constant ( 𝑘𝑘 = 7.7(6) × 10−9 cm3s-1) and branching ratio ( 𝜂𝜂 =

HOC+/HCO+ = 2.1(4)) that can correct inconsistencies in the models (48, 49). Quasi-classical

trajectory calculations on a highly accurate potential energy surface, reveal the reaction

proceeds by a direct mechanism to overwhelmingly (97.6%) produce HOC+. However, due to

the large exothermicity of the reaction, the internal energy of a majority (87.8%) of these

products is larger than the isomerization barrier. During radiative cooling ~1/3 of the above-

barrier HOC+ are expected to isomerize to HCO+, lowering the predicted branching ratio to

𝜂𝜂 = 2.3.

Interstellar molecular clouds ultimately yield the seedbed from which new stars and planets

are born, as well as the raw materials from which life likely developed. Thus, understanding

the chemical evolution of the ISM is essential to understanding these important processes.

Chemical reaction rates and branching ratios underlying models of this evolution are often

(necessarily) taken from experiments and theory that are not appropriate for the conditions

present in cold molecular clouds. The apparatus employed here is quite versatile and could be

applied in its present form to study most atomic ion – neutral molecule reactions under

conditions relevant to cold molecular clouds of the ISM chemistry.

Page 15: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

Supporting Information

1. Characterization of Cryogenic Buffer Gas Beam (CBGB)

Gaining a quantitative understanding of the reaction temperature in the beam + trap scenario

in this work requires characterization of the velocity of the target species entrained within the

buffer gas beam. For this, we employ an atomic tracer species that can be directly detected in

small amounts. Ytterbium metal is known to have good ablation properties and the produced

neutral isotopes have well known isotope shifts and hyperfine splittings. By ablating ytterbium

foil inside of the cryogenic cell while the neon gas is being introduced, the ytterbium is cooled

by the buffer gas and carried out of the cell. As long as the target species number density is a

trace amount in comparison to the bulk buffer gas number density (0.1%), the flow

characteristics are dominated by the buffer gas species(50). To determine the velocity of the

beam, we use a 399 nm laser to excite the Yb and fluorescence is detected. The laser is scanned

over 3 GHz, encompassing all naturally occurring Yb isotope transitions. Two scans are taken,

one with the 399 nm light perpendicular to the beam path (transverse), and the other with the

laser at an angle of 𝜃𝜃 = 57.3° relative to the beam, shown in Figure S1. Since the individual

transitions are all thermally broadened well beyond the natural linewidth, the spectrum is fitted

with summed Gaussian functions with predetermined isotopic shifts taken from literature. This

yields the center frequency as well as the thermally broadened width of the lines. Between the

transverse and angled scans, we find an offset in the spectrum caused by the doppler shift, seen

in figure S2,

Δ𝑓𝑓 = Δ𝑣𝑣𝑣𝑣𝑓𝑓 cos(𝜃𝜃), (S1)

where 𝑓𝑓 is the rest-frame center frequency, Δ𝑓𝑓 is the offset observed, 𝑣𝑣 is the forward

Page 16: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

velocity of the excited Yb, and 𝜃𝜃 is the angle of the atomic beam with respect to the excitation

laser beam. The fitted shared widths indicate a beam temperature of (20 ± 1) K, which is the

temperature of the cryogenic cell. Using equation S1, the forward velocity of the beam is ≈

(150±3)ms

, in good agreement with a mildly boosted neon buffer gas of equilibrium temperature

20 K. The neon buffer gas densities are sufficient to ensure that the water seeded into the beam

makes 100s of collisions before extraction into a ballistic beam. These 100s of collisions ensure

that the internal rotational states of the water have reached 20 K. Before extraction, the

translational degrees of freedom have also reached 20 K, however, extraction of a beam

operating within the intermediate flow regime causes collisions at the aperture, providing slight

forward boosting of its forward velocity as well as narrowing of the velocity distribution. In

the center-of-mass frame, the reaction occurs with an internal temperature for the H2O of ~10-

15 K (given by 𝐸𝐸𝑘𝑘𝑘𝑘

= 𝜇𝜇𝑣𝑣𝑟𝑟𝑟𝑟𝑟𝑟2

2𝑘𝑘𝑘𝑘, 𝜇𝜇 is reduced mass). We report this collision energy as ~20 K in

the manuscript to convey that it is an approximate collision energy. Due to the anisotropic

boosting, the velocity doesn’t follow a Maxwell Boltzmann distribution, and we cannot make

many concrete statements about it, except that it is narrower than a Maxwell Boltzmann.

Figure S1. Diagram of entrained neutral Yb calibration of the CBGB. H2O and Yb are co-entrained in a buffer gas

Page 17: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

of neon. The 399 nm laser is scanned over 3 GHz, recording the Yb LIF spectrum both transverse, and at a 57.3°

angle, to the beam.

Figure S2. Angled longitudinal scan of Yb fluorescence collected by a PMT ≈ 24 cm from the cell aperture.

The frequency indication is relative to the measured 174Yb frequency from the scan with the excitation laser

perpendicular to the atomic beam. A fit of Gaussians on the observed Yb isotopic transitions is shown with fixed

relative line centers, while a shared width and the individual heights are fitting parameters. Fits yield a forward

velocity of ≈ (150 ± 3) m/s and broadened width of (20 ± 1) K. We find that the Yb is entrained within

the neon and sympathetically cooled to the cryogenic cell’s temperature. The water that is also introduced in the

beam will be at a similar temperature and forward velocity since the beam dynamics is dominated by the properties

of the buffer gas species – that is, since the neon density is much larger than that of the water(51).

To control the beam flux, we insert a vacuum compatible Optical Shutter (Uniblitz VS35) in

the molecular beam line. The shutter does not create a seal within the chamber, and background

gas molecules can potentially find their way around and influence the beam.

Page 18: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

To measure the H2O in the beam directly and ascertain the effectiveness of the shutters, we

use two methods. First, we use a residual gas analyzer (RGA) in the beam path. By opening

and closing shutters in the beam path we observe a clear, prompt extinction of the water signal.

Second, we use the ions in the trap themselves to observe the beam and constrain any

background water. Trapped Be+ ions react with H2O to predominately produce BeOH+, which

we see as a drop in the fluorescence in Figure S3 and Figure S4. These figures compare the

fluorescence decay as a beam from the CBGB is blocked by in vacuo shutter in the beam line

to the unblocked case. Examining the fitted reaction rates, we find that they agree with the

background reaction rate shown in figure S4. This indicates that a beam of cryogenic water is

impinging upon the trapped ions and reactions of background H2O are insignificant relative to

the other reaction rates.

Figure S3. Fluorescence decay of trapped Be+ ions exposed to a cryogenic water beam with an inline shutter in

vacuo either opened, in green (𝜏𝜏 = 7.2 × 10−3 s) or closed, in blue (𝜏𝜏 = 63.7 × 10−3 s)

Page 19: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

Figure S4. Fluorescence decay of trapped Be+ ions exposed to a cold-water beam with the inline shutter closed at

𝑡𝑡 ≈ 25 s. Fits to the decay with the shutter open vs. closed yield Be+ lifetime of 𝜏𝜏 = 7.6 × 10−3 s and 𝜏𝜏 =

53.7 × 10−3 s, respectively.

2. Proton transfer reactions

To distinguish the two isomers of the formyl ion produced in the title reaction, we use proton

transfer reactions with 15N2. This choice is based on a number of factors explained in this

section.

A general form of the desired titration reaction for analyzing the isomeric branching ratio

with species X is:

HOC+ + X → XH+ + CO

HCO+ + X → No Reaction

In selecting a titration gas, the most important property, other than possessing a proton affinity

Page 20: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

between the two isomers, is a low reactivity with the other trapped species. If the titration gas

reacts with Be+, C+, or any other ionized species in the trap, it can be difficult to disambiguate

the products and the branching ratio measurement may be affected. Table 1, shows candidate

titration gases with proton affinity between 4.43 eV and 6.16 eV that only react with HOC+.

The previous measurement of the branching ratio used in a selected ion flow tube (SIFT)

instrument used N2O at 305 K. Further, rate coefficients for reactions with O2, Kr, Xe, CO,

CO2, and CH4 have been explored (22, 52). In our work, we explored titration with Xe, CO2,

CH4, N2O, and 15N2.

The reaction of N2O and Be+ has not been previously studied. We observed multiple

subsequent reactions occurring that lead to loss of Be+ without replacement by another ion –

i.e. overall trap loss. This is typically only possible when a much heavier charged product is

produced that is not stable in the ion trap. Without the coolant ion, the mass resolution of the

TOF declines significantly making discrimination of the target masses difficult. Further, N2O

is known to react with C+ to make NO+ and CN with a total rate coefficient of 9.1 ×

10−10 cm3/s at 300 K (53), which adds more complexity to the mass spectra. Similarly, CH4

reacted with the trapped C+, producing new m/z peaks in the mass spectra due to both direct

and higher-order reactions. With Xe, the charged product XeH+ has a mass to charge ratio of

m/z = 132, which was not stable in our trap.

For these reasons, we chose not to use N2O, CH4, or Xe as titration gases and instead used

N2. N2 has several benefits including ease of use and a lack of strong reaction with C+ or Be+.

N2 had not previously been used in the literature, most likely because the reaction product N2H+

has nearly the same m/z as the formyl ion isomer, making mass spectrometry difficult to apply.

Page 21: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

We circumvented this problem by using the rarer isotope, 15N2.

Table S1: Proton affinities of gasses between formyl isomers where (*) indicates H bonding

location (54).

3. Experimental Sequence and Chemical Rate Equations

3.1 Experimental sequence

A typical experimental sequence and ion “cloud” image are shown in Fig. S5. C+ ions are

first loaded and sympathetically cooled by Be+ ions in the ion trap. Typically, hundreds of ions

are trapped, but not crystallized into a Coulomb crystal, shown on the bottom of Fig. S5. In the

un-crystallized phase, dark C+ ions sit at the trap center and are surrounded by laser cooled,

fluorescing, Be+ ions. The 20 K, H2O beam is then introduced into the trap for 20 s and

reactions occur. The beam is then blocked by the in vacuo molecular beam shutter (100 ms

response time) and (on a 1 second timescale) a gate valve. An approximately one second delay

is provided to allow the internally excited product ions to isomerize and/or relax to their

Page 22: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

vibrational ground state via radiative decay (43). Next, the 15N2 titrant gas is introduced via a

room-temperature leak valve for approximately 10 s. Afterwards, all ions are ejected into the

TOF-MS.

Figure S5. A schematic diagram of the experimental setup, time sequence and ion cloud.

3.2 The production of the formyl isomers and subsequent reactions with water:

mathematical model

In this section, we consider in experimental detail the following reactions in the main text:

C+ + H2O → HOC+ + H (1)

C+ + H2O → HCO+ + H (2)

[HCO]+ + H2O → H3O+ + CO. (5-6)

Page 23: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

We treat the case in which the water density in the beam is constant, and assume that pseudo

first order kinetics applies. The differential forms of the densities subject to the above reactions

are

�̇�𝐶(𝑡𝑡) = −𝑘𝑘1+2𝜌𝜌𝐻𝐻2𝑂𝑂𝐶𝐶(𝑡𝑡)

[𝐻𝐻𝐶𝐶𝐶𝐶]̇ (𝑡𝑡) = 𝜌𝜌𝐻𝐻2𝑂𝑂(𝑘𝑘1+2𝐶𝐶(𝑇𝑇) + 𝑘𝑘5[𝐻𝐻𝐶𝐶𝐶𝐶](𝑇𝑇) )

𝐻𝐻3�̇�𝐶(𝑡𝑡) = 𝑘𝑘5𝜌𝜌𝐻𝐻2𝑂𝑂[𝐻𝐻𝐶𝐶𝐶𝐶](𝑇𝑇)

where 𝐶𝐶(𝑡𝑡) is the carbon ion density, [𝐻𝐻𝐶𝐶𝐶𝐶](𝑡𝑡) is the density of both isomers of the formyl

ion, 𝜌𝜌𝐻𝐻2𝑂𝑂 is the density of water, and 𝐻𝐻3𝐶𝐶(𝑡𝑡) is the density of H3O+. These differential

equations are solved by,

𝐶𝐶(𝑡𝑡) = 𝐶𝐶0𝑒𝑒−𝑘𝑘1+2 𝜌𝜌𝐻𝐻2𝑂𝑂𝑡𝑡 (S2)

𝐻𝐻𝐶𝐶𝐶𝐶(𝑡𝑡) = 𝑒𝑒−(𝑘𝑘1+2+𝑘𝑘5)𝜌𝜌𝐻𝐻2𝑂𝑂𝑡𝑡

𝑘𝑘1+2−𝑘𝑘5(𝑒𝑒𝑘𝑘1+2𝜌𝜌𝐻𝐻2𝑂𝑂𝑡𝑡((𝐶𝐶0 + [𝐻𝐻𝐶𝐶𝐶𝐶]0)𝑘𝑘1+2 − [𝐻𝐻𝐶𝐶𝐶𝐶]0𝑘𝑘5 − 𝐶𝐶0𝑒𝑒−(𝑘𝑘1+2+𝑘𝑘5)𝜌𝜌𝐻𝐻2𝑂𝑂𝑡𝑡) (S3)

𝐻𝐻3𝐶𝐶(𝑡𝑡) = 𝐻𝐻3𝐶𝐶0 + [𝐻𝐻𝐶𝐶𝐶𝐶]0�1 − 𝑒𝑒−𝑘𝑘5𝜌𝜌𝐻𝐻2𝑂𝑂𝑡𝑡� +𝐶𝐶0(𝑘𝑘1+2�1−𝑒𝑒

−𝑘𝑘5𝜌𝜌𝐻𝐻2𝑂𝑂𝑡𝑡�+𝑘𝑘5�𝑒𝑒−𝑘𝑘1+2𝜌𝜌𝐻𝐻2𝑂𝑂𝑡𝑡−1�)

𝑘𝑘1+2−𝑘𝑘5. (S4)

3.3 The reactions of the isomers [HCO]+ with titration gas 15N2

For the reactions

HOC+ + 15N2 → 15N2H+ + CO (3)

and

HCO+ + 15N2 → No reaction, (4)

the differential form of the densities of these species is given by

𝐻𝐻𝐶𝐶𝐶𝐶+(𝑡𝑡)̇ = −𝑘𝑘3𝜌𝜌𝐻𝐻𝐶𝐶𝐶𝐶+(𝑡𝑡)

Page 24: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

and

𝑁𝑁 15 2𝐻𝐻+(𝑡𝑡)̇ = 𝑘𝑘3𝜌𝜌𝐻𝐻𝐶𝐶𝐶𝐶+(𝑡𝑡),

with 𝐻𝐻𝐶𝐶𝐶𝐶+(𝑡𝑡) the density of HOC+, 𝜌𝜌 the density of 15N2, and 𝑁𝑁 15 2𝐻𝐻+(𝑡𝑡) the density of

15N2H+.

Since we do not distinguish between HOC+ and HCO+ in the mass spectra, the mass spectrum

reveals only the value of

[𝐻𝐻𝐶𝐶𝐶𝐶]+(𝑡𝑡) = 𝐻𝐻𝐶𝐶𝐶𝐶+(𝑡𝑡) + 𝐻𝐻𝐶𝐶𝐶𝐶+(𝑡𝑡) .

Due to the unreactivity of HCO+ with N2 (reaction 4), 𝐻𝐻𝐶𝐶𝐶𝐶(𝑡𝑡) = 𝐻𝐻𝐶𝐶𝐶𝐶0, the initial density of

HCO+ formed by the title reaction. The solutions to the observable densities, then, are

[𝐻𝐻𝐶𝐶𝐶𝐶]+(𝑡𝑡) = 𝐻𝐻𝐶𝐶𝐶𝐶0+𝑒𝑒−𝑘𝑘3𝜌𝜌𝑡𝑡 + 𝐻𝐻𝐶𝐶𝐶𝐶0+ (S5)

𝑁𝑁 15 2𝐻𝐻+(𝑡𝑡) = 𝐻𝐻𝐶𝐶𝐶𝐶0+(1 − 𝑒𝑒−𝑘𝑘3𝜌𝜌𝑡𝑡) + 𝑁𝑁 15 2𝐻𝐻0+ (S6)

Manipulation of solutions S5 and S6, setting 𝑁𝑁 15 2𝐻𝐻0+ = 0, yields the ratio of N2H+ and

[HCO]+,

𝑁𝑁 15

2𝐻𝐻+(𝑡𝑡)𝑁𝑁 15 2𝐻𝐻+(𝑡𝑡)+[𝐻𝐻𝐶𝐶𝑂𝑂]+(𝑡𝑡)

= 𝜁𝜁(1 − 𝑒𝑒𝑘𝑘3𝜌𝜌𝑡𝑡) (S7)

where 𝜁𝜁 ≡ 𝐻𝐻𝑂𝑂𝐶𝐶0𝐻𝐻𝑂𝑂𝐶𝐶0+𝐻𝐻𝐶𝐶𝑂𝑂0

. As shown in Fig. 1(C) and described above, to distinguish the isomers

[HCO]+, the 15N2 titrant gas is introduced via a room-temperature leak valve for approximately

10 s. Repeating this process over various densities of 15N2 allows us to determine the isomer

branching ratio as well as the proton transfer rate constant. A fit performed on the data over

various densities (refer to Fig.1(C) in the main text) yields a rate constant of 𝑘𝑘3 =

(6.6 ± 1.0) × 10−10 cm3/s, and a final branching ratio of HOC+ to HCO+ is 1.4 (2). The

Page 25: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

error bars here are the statistical errors. and we estimate another 30% systematic errors for the

rate constants due to the measurement of absolute density. The details about the density

measurement can be found in our previous work(34). The branching ratio is independent of the

absolute H2O density, because these are measured in the same H2O gas.

3.4 Verification of N2 proton affinity

As described in the main text, (5) and (6) are the secondary reactions in our system. Though

the appearance of H3O+ presents no problem for our study, a relative rate difference between

these two reactions ((5) and (6)) at low temperature would systematically shift the trapped

HOC+/HCO+ ratio and cause an error in the determination of branching ratio 𝜂𝜂. To constrain

this potential error, we repeat the sequence described above in the presence of a large amount

of neutral CO gas to produce samples with varying compositions of HCO+ and HOC+ via

reaction (7).

HCO+ + H2O → H3O+ + CO, (5)

HOC+ + H2O → H3O+ + CO. (6)

HOC+ + CO→ HCO+ + CO. (7)

CO is introduced continuously via the leak valve in the differential cross region to produce

a pressure of ≈ 4 × 10−9 Torr as measured in the trap chamber. The constant introduction of

CO ensures full conversion of HOC+ to HCO+ at a rate ~10 times faster than that of reactions

(1) and (2) ensuring we observe the time evolution of reaction (5) as seen in Figure S6b. A

Page 26: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

similar procedure of continuously exposing the trap to the CBGB without CO yields a

combination of the rates of reactions (5) and (6), also shows as reaction (5-6) is seen in Figure

S6a).

The rates of reactions (1), (5), and reaction (5-6) are found with least-squares fitting of the

solutions to differential equations found above. Beam densities are determined for each run

individually by considering the Be+ + H2O reaction.

Figure S6. Time evolution of C+ and H2O introduced via CBGB as well as subsequent reaction products. a) TOF

traces without flooding of CO where fitted rate constants are found to be 𝑘𝑘1 + 𝑘𝑘2 = (7.7 ± 0.6) × 10−9 cm3/s,

and 𝑘𝑘5 = (1.7 ± 0.2) × 10−8 cm3/s. b) TOF traces with flooding of CO where fitted rate constants are found

to be 𝑘𝑘1 + 𝑘𝑘2 = (7.9 ± 0.6) × 10−9 cm3/s, and 𝑘𝑘5 = (1.7 ± 0.2) × 10−8 cm3/s.

By producing only HCO+, we directly observe the reaction (5) at cold temperature and

measure the rate coefficient to be 𝑘𝑘5 = (1.7 ± 0.2) × 10−8 cm3/s . Although we cannot

directly measure the rate coefficient of reaction (6), we repeat this measurement with or without

the presence of CO and, therefore, to make the pure HCO+ or the combination of HCO+ and

HOC+ We find there is no deviation between the pure HCO+ + H2O and [HCO]+ + H2O (recall

that [HCO]+ represents the combination of the two isomers). Thus, we use the measured 𝑘𝑘5

for both reactions (5) and (6). Theoretically, the similar rate constant may not be surprising

Page 27: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

given that the only difference at the capture theory level between reactions (5) and (6) comes

from the short-ranged, dipole-dipole portion of the interaction, which is much weaker than the

ion-dipole and ion-quadrupole contributions.

Typical TOF-MS data for this procedure is shown in Fig. 1(D), and show no production of

any 15N2H+. Therefore, we conclude that both a pure HCO+ sample can be prepared with the

presence of CO in the ion trap (52) and reaction (4) is indeed energetically forbidden.

3.5 Isomerization by H2O

Additionally, in principle, H2O could isomerize the isoformyl ion in a second order

reaction according to HOC+ + H2O → HCO+ + H2O. However, H2O readily captures a

proton from HOC+, and because its proton affinity is greater than both HCO+ and HOC+ (see

table S1), it does not return the proton to the carbon end of CO to give HCO+. Instead, it is

energetically profitable for the water to retain the proton and form CO + H3O+ rather, i.e.,

reactions (6), and isomerization by H2O does not occur (45).

4. Quasi-classical trajectory calculations on N2 + HOC+ → N2H+ + CO and N2 +

HCO+

The product branching of this process is of great importance to determine the product

branching ratio of the C+ + H2O reaction because of the titration scheme used in the experiment.

To this end, some of the authors have recently published a global potential energy surface

(PES) and performed dynamical calculations using quasi-classical trajectory (QCT) (55). Here,

we have improved the original PES by adding long-range terms needed to understand low-

temperature dynamics. Specifically, the following long-range interaction between N2 and

HOC+ is used:

Page 28: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

𝑉𝑉𝑖𝑖𝑖𝑖𝑡𝑡(𝑅𝑅,𝜃𝜃) = 𝑉𝑉𝑞𝑞𝛩𝛩(𝑅𝑅, 𝜃𝜃) + 𝑉𝑉𝑞𝑞𝑞𝑞(𝑅𝑅,𝜃𝜃) = 𝑞𝑞𝛩𝛩(3𝑐𝑐𝑐𝑐𝑐𝑐2𝜃𝜃−1)2𝑅𝑅3

−𝑞𝑞2[𝑞𝑞+

𝛼𝛼∥−𝛼𝛼⊥3 (3𝑐𝑐𝑐𝑐𝑐𝑐2𝜃𝜃−1)]

2𝑅𝑅4, (S8)

where R denotes the distance between the center of mass of N2 and O and θ the angle between

the R vector and the N-N bond. Here the HOC+ is approximated as a point charge located on

O. The quadrupole of N2 is -1.47 Debye∙Å, and its polarizabilities are given as 𝛼𝛼 =

1.767 Å3,𝛼𝛼∥ − 𝛼𝛼⊥ = 0.69 Å3 (56). This long-range potential consists of the interaction

potential and the potentials for the molecules

𝑉𝑉𝐿𝐿𝑅𝑅 = 𝑉𝑉𝑖𝑖𝑖𝑖𝑡𝑡 + 𝑉𝑉𝑁𝑁2 + 𝑉𝑉𝐻𝐻𝑂𝑂𝐶𝐶+, (S9)

which is connected with the short-range one (55)

𝑉𝑉𝑃𝑃𝐸𝐸𝑃𝑃 = 𝑆𝑆𝑉𝑉𝑃𝑃𝑅𝑅 + (1 − 𝑆𝑆)𝑉𝑉𝐿𝐿𝑅𝑅, (S10)

using the switching function:

𝑆𝑆 = 1−tanh [2(𝑅𝑅−6.5Å)]2

. (S11)

With this long-range corrected PES, we have recalculated the branching between the N2H+

+ CO and N2 + HCO+ channels at experimental conditions using 5000 trajectories, and found

that the isomerization channel (N2 + HCO+) has a 18% probability. The details of the

calculation can be found in Ref. (55).

5. Quasi-classical trajectory calculations on the C+ + H2O reaction

The global PES for the C+(2P) + H2O(X1A1) HCO+/HOC+(X1Σ+) + H(2S) reaction consists

of a long-range (LR) term in the reactant asymptote and a short-range (SR) term for the strongly

interaction region. The former, which is important for an accurate description of the reaction

kinetics, accounts for the leading electrostatic interactions between the two reactants when they

are far apart. The latter, on the other hand, is obtained by fitting ab initio data. The two segments

of the PES are connected smoothly via a switching function, 𝑆𝑆𝑔𝑔𝑔𝑔𝑐𝑐𝑘𝑘𝑔𝑔𝑔𝑔:

Page 29: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

𝑉𝑉 = 𝑆𝑆𝑔𝑔𝑔𝑔𝑐𝑐𝑘𝑘𝑔𝑔𝑔𝑔𝑉𝑉SR + �1 − 𝑆𝑆𝑔𝑔𝑔𝑔𝑐𝑐𝑘𝑘𝑔𝑔𝑔𝑔�𝑉𝑉LR, (S12)

𝑆𝑆𝑔𝑔𝑔𝑔𝑐𝑐𝑘𝑘𝑔𝑔𝑔𝑔 = 1−tanh[3(𝜉𝜉−9.0)]2

, (S13)

where 𝜉𝜉 is the distance (in Å) between C and O. the PES is dominated by VLR for 𝜉𝜉 > 9.5 Å,

while by VSR for 𝜉𝜉 < 8.5 Å.

The LR PES can be written as:

𝑉𝑉LR = 𝑉𝑉H2O + 𝑉𝑉ES, (S14)

where 𝑉𝑉H2O is the PES for the isolated H2O molecule. This local PES is developed by a three-

dimensional cubic spline interpolation of 1274 ab initio points at the explicitly correlated

unrestricted coupled cluster singles, doubles, and perturbative triples (UCCSD(T)) method (57)

with a specially optimized triples zeta correlation consistent F12 basis set (VTZ-F12) level (58),

the same as the SR PES calculations (vide infra). All the potential energies of ab initio points

are related to the potential minimum of H2O at its equilibrium geometry. The electrostatic

interactions VES between the C+ ion and H2O molecule consist of several long-range terms. In

describing these interactions, we assume that the water molecule is in its equilibrium geometry

with the O atom in the z axis and the molecule is in the yz plane. R is the distance between C+

ion and the center of mass of H2O. We define θx, θy and θz as the angles between the 𝑅𝑅�⃗ vector

and x, y and z axes, respectively. The leading electrostatic terms include the charge-dipole

(proportional to R-2), charge-quadruple (proportional to R-3), and charge-induced dipole

(proportional to R-4) interactions (59, 60):

𝑉𝑉ES = 𝑉𝑉c−d + 𝑉𝑉c−q + 𝑉𝑉c−id, (S15)

where the charge-dipole interaction potential is:

𝑉𝑉c−d = 𝑞𝑞𝜇𝜇H2O𝑅𝑅−2 cos 𝜃𝜃𝑧𝑧, (S16)

Page 30: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

the charge-quadruple interaction potential is:

𝑉𝑉c−q = 12𝑞𝑞𝑅𝑅−3�Θ𝑧𝑧𝑧𝑧(3 cos2 𝜃𝜃𝑧𝑧 − 1) + Θ𝑥𝑥𝑥𝑥(3cos2 𝜃𝜃𝑥𝑥 − 1) + Θ𝑦𝑦𝑦𝑦�3cos2 𝜃𝜃𝑦𝑦 − 1�� (S17)

and the charge-induced dipole interaction potential is:

𝑉𝑉c−id = −12𝑞𝑞𝛼𝛼𝑅𝑅−4 �1 + �𝑞𝑞𝑧𝑧𝑧𝑧−1 2�𝑞𝑞𝑥𝑥𝑥𝑥+𝑞𝑞𝑦𝑦𝑦𝑦�⁄ �

3𝑞𝑞(3 cos2 𝜃𝜃𝑧𝑧 − 1)�. (S18)

For the H2O molecule, the dipole moment 𝜇𝜇H2O is 1.855 Debye, the three non-zero quadruple

moments Θ𝑥𝑥𝑥𝑥, Θ𝑦𝑦𝑦𝑦 and Θ𝑧𝑧𝑧𝑧 are -2.5, 2.63 and -0.13 Debye Å, and the polarizability α is

9.92 a.u. (αxx, αyy and αzz are 10.31, 9.55 and 9.91 a.u.).

For the SR PES, we first carried out ab initio calculations using the UCCSD(T)-F12a method

with a basis set VTZ-F12. The stationary points along the reaction path are shown in Figure 2

and their geometries and energies are listed in Table S2. These ab initio points are then fitted

using the high-fidelity permutation invariant polynomial-neural network (PIP-NN) method

(61). In particular, 17 PIPs up to second order have included as the input layer of the NN in

order to take advantage of the permutation symmetry of this ABC2 system. The details of the

NN training can be found in ref. (61). The fitting is carried out in two different regions:

(a) The reactant asymptotic region V1: There are 5113 points with the distance between C

and O from 4 Å to 50 Å. The NN consists of 2 layers, with 20 and 60 interconnected

neurons. The total number of parameters is thus 1681. The RMSE of the fitting is less

than 0.05 kcal/mol.

(b) The main region V2: There are 23749 points with the distance between C and O less than

6 Å. The NN consists of 2 layers, with 30 and 60 interconnected neurons. The total

number of parameters is thus 2461. The RMSE of the fitting is less than 0.5 kcal/mol.

Page 31: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

The final SR PES VSR is obtained by connecting these two fits with another switching

function 𝑆𝑆𝑓𝑓𝑖𝑖𝑡𝑡:

𝑉𝑉𝑓𝑓𝑖𝑖𝑡𝑡 = 𝑆𝑆𝑓𝑓𝑖𝑖𝑡𝑡𝑉𝑉1 + �1 − 𝑆𝑆𝑓𝑓𝑖𝑖𝑡𝑡�𝑉𝑉2, (S19)

𝑆𝑆𝑓𝑓𝑖𝑖𝑡𝑡 = 1−tanh[3(𝜉𝜉−5.0)]2

,

(S20)

As shown in Table S1, the PIP-NN PES faithfully reproduces energies and geometries of the

stationary points. The stationary points along the reaction path are also optimized at the

CASSCF/6-31G* level, and corrected for dynamical correlations at the multi-reference

configuration interaction (MRCI) level with the AVTZ basis set.

Table S2 Energies (kcal/mol) and harmonic frequencies (cm−1) of the stationary points for the C + +

H2O → HCO+ /HOC+ + H reaction.

Species Method E (kcal/mol)

Frequency (cm-1)

1 2 3 4 5 6 IM1

CASSCF* UCCSD(T)

PES

-87.2 -87.7 -87.5

3491.2 3568.3 3570.5

3393.1 3465.7 3419.2

1697.0 1630.1 1649.3

970.9 936.6 904.3

863.6 899.9 876.6

531.8 460.0 418.7

IM2

CASSCF UCCSD(T)

PES

-84.9 -85.2 -85.4

2508.0 2553.8 2630.5

1963.6 1934.7 1908.7

825.5 696.0 759.2

809.1 661.9 580.8

728.4 661.4 580.8

76. 65.8

111.6 IM3

CASSCF UCCSD(T)

PES

-140.7 -141.6 -141.9

3486.8 3499.0 3576.4

3083.2 3064.2 3052.4

1660.9 1691.5 1671.7

1263.8 1253.5 1269.7

1021.8 999.0

1148.0

935.4 951.6 993.2

IM4

CASSCF UCCSD(T)

PES

-136.9 -137.9 -137.9

3372.5 3430.3 3481.2

2872.2 3044.1 3086.4

1781.1 1716.4 1653.3

1374.9 1169.4 1303.8

1146.3 985.9

1058.2

972.5 924.4 935.1

IM5

CASSCF UCCSD(T)

PES

-145.9 -146.3 -146.7

2940.8 2905.4 2895.1

2818.6 2796.2 2880.2

1654.9 1666.7 1686.1

1299.4 1260.5 1241.5

1076.8 1061.1 1032.0

917.5 846.6 798.0

IM6

CASSCF UCCSD(T)

PES

-119.5 -120.7 -121.0

3207.4 3015.1 3145.8

2181.8 2171.2 2105.1

820.9 900.7 936.4

807.2 898.9 936.4

215.5 363.1 553.0

101.7 152.3 176.7

TS12

CASSCF UCCSD(T)

PES

-56.8 -55.6 -56.5

3394.9 3439.8 3622.1

1496.3 1549.8 1611.3

1247.6 1127.4 1229.6

819.9 772.2 791.3

570.0 463.2 507.3

2477.4i 2380.6i 2214.7i

Page 32: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

TS13

CASSCF UCCSD(T)

PES

-53.9 -53.7 -53.9

3279.5 3328.0 3231.6

2303.3 2202.1 2278.9

1251.4 1285.9 1325.9

997.3 967.1

1045.6

297.5 155.1i 324.3

2133.8i 2057.9i 2121.9i

TS34

CASSCF UCCSD(T)

PES

-121.8 -123.2 -122.5

3525.1 3351.8 3359.9

3118.5 3243.5 3057.6

1656.1 1692.8 1692.6

1244.7 1164.5 1174.8

860.4 837.9 376.2i

1234.2i 1242.1i 1136.1i

TS35

CASSCF UCCSD(T)

PES

-97.6 -101.2 -102.1

2089.7 2999.5 2814.1

1801.5 2273.6 2205.3

1180.8 1712.5 1785.6

1080.0 1051.1 1110.5

989.0 678.9 874.6

1238.2i 2133.9i 2130.6i

TS4

CASSCF UCCSD(T)

PES

-97.5 -101.9 -102.2

2530.0 3161.8 3339.4

1818.5 1954.2 1980.8

1084.5 868.7 999.9

959.0 829.5 859.6

653.4 652.3 788.3

1689.3i 1964.0i 2328.8i

TS56

CASSCF UCCSD(T)

PES

-118.9 -118.0 -117.3

3977.2 3196.4 3248.8

2055.8 2131.1 2029.0

1144.7 939.4 865.4

158.9 839.3 765.5

987.1i 517.8 550.0

2584.5i 673.6i 909.1i

*The geometry optimizations and frequencies analysis by CASSCF/6-31G*, and the energies are corrected at the MRCI/AVTZ level.

Because of the large density of states in this system, it is still impractical to investigate the

dynamics with a quantum mechanical method. For such an exoergic and barrierless reaction,

on the other hand, the QCT approach is reasonable because quantum effects such as tunneling

are not expected to be important. The QCT calculations are performed using VENUS (62). The

trajectories are initiated with a 41.0 Å separation between reactants, and terminated when

products reach a separation of 10.0 Å. The largest impact parameter chosen is bmax=39 Å to

ensure the inclusion of long-range interactions. The propagation time step is 0.05 fs. The total

energy conservation is found to be better than 0.04 kcal/mol for all the trajectories. A few

trajectories are discarded if the propagation time reached 300 ps. The violation of the product

zero-point energy is expected to be minimal due to the large exothermicity of the reaction.

The thermal rate constant can be computed by the following expression:

𝑘𝑘(𝑇𝑇) = �8𝑘𝑘B𝑇𝑇𝜋𝜋𝜇𝜇

�1/2

𝜋𝜋𝑏𝑏𝑚𝑚𝑔𝑔𝑥𝑥2 𝑁𝑁𝑟𝑟𝑁𝑁𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑟𝑟

, (S21)

Page 33: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

where μ is the reduced mass between the two reactants, kB the Boltzmann constant, and T

temperature in Kelvin. Nr and Ntotal are the number of reactive trajectories and total number of

trajectories, respectively. At the experimental temperatures of 20 K, the initial rotational

energies of the water reactant and relative translational energies between C+ and H2O are

sampled according to the Boltzmann distribution. A total of 48465 trajectories are run at this

temperature. The branching ratio is computed from the rate coefficients of the two product

channels.

Despite the higher energy of the HOC+ isomer, the QCT results indicate that it accounts for

97% of the products. The dominance of the HOC+ product is consistent with the prediction by

an earlier room temperature kinetic study (22) and direct dynamics studies on the same process

using a lower-level ab initio theory (29, 63). As discussed in the main text, the reaction

typically forms a short-lived intermediate (IM1) with the configuration C-OH2. The transfer of

the energy from the C-O mode to O-H mode leads the dissociation to major product H + HOC+

channel. On the other hand, some trajectories fail to break the O-H bond but the energy

imparted into the H atom leads to large amplitude motion and the formation of HCOH species,

which eventually dissociates into the minor HCO+ + H channel by cleaving the O-H bond. This

is illustrated in Figure S7 where the OH, OH’, CO, CH and CH’ distances are plotted for two

exemplary trajectories leading to these two channels.

Page 34: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

Figure S7. Internuclear distances in two exemplary trajectories leading to HOC+ + H and HCO+

+ H channels.

6. Isomerization in radiative decay of hot HOC+/HCO+ products

Our QCT results on the ground state PES indicate significant internal excitations of both

HCO+ and HOC+ products. Specifically, 87.8% of HOC+ and 29.6% of HCO+ have internal

energies above the isomerization barrier. Since these ions are stored in the trap for more than

1 s before the introduction of N2, any high-lying vibrational states with energy above the barrier

would have relaxed via spontaneous emission to lower-lying, below-the-barrier vibrational

states (64, 65). This relaxation is likely accompanied with isomerization and alters the product

branching ratio from the value found in the QCT calculation. This is facilitated by gateway

states immediately above the isomerization barrier, which are vibrational eigenstates that

possess amplitudes in both the HOC+ and HCO+ wells. The formation of such states is

facilitated by the accidental degeneracy of zeroth order states associated with the two isomers

and anharmonic couplings between them, as illustrated in recent work on the vinylidene-

acetylene system (66). The radiative decay of such a gateway state will thus lead to population

Page 35: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

transfer from one well to another. More importantly, such isomerization is irreversible if the

gateway state is just above the isomerization barrier, as the vibrational relaxation can only take

place to local vibrational states in the two wells. Unfortunately, a spectroscopically accurate

determination of these gateway states and the rest of the highly excited vibrational energy

levels is still beyond our capability, despite the availability of the HCO+/HOC+ PES (67).

The amount of isomerization can be estimated in terms of the density of states (DOS) of

the two isomers. The vibrational DOS at the barrier has been calculated for both isomers, based

on the harmonic frequencies listed in Table S3. These frequencies are in reasonably good

agreement with earlier theory (67). Despite the higher energy of HOC+, it has a higher

vibrational DOS (40 /cm-1) at the isomerization barrier than that of HCO+ (20 /cm-1) due to its

extremely low bending frequency. It can thus be estimated based on the DOSs that 33% of the

HOC+ above the barrier would isomerize during its radiative decay, leading to a final

HOC+:HCO+ ratio of 2.3, in good agreement with the experimental result. As 97% of the initial

product population is in the HOC+ well, on the other hand, it is conceivable that the

isomerization is unidirectional, namely the isomerization of above-the-barrier HCO+, which

has an initial population of merely 3%, can be neglected.

Table S3. Calculated energies, geometries, and harmonic frequencies of the HCO+ and HOC+

Species Energy

(kcal/mol)

rCO

(Å)

rCH

(Å)

rOH

(Å)

C-O stretching

(cm-1)

C-H/O-H stretching

(cm-1)

Bending (cm-1)

HCO+ 0 1.108 1.094 2.202 3219 2208 846

HOC+ 44.7 1.159 2.151 0.992 3460 1937 84

Page 36: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

References:

1. Y. Aikawa, V. Wakelam, R. T. Garrod, E. Herbst, Molecular evolution and star formation: From prestellar cores to protostellar cores. Astrophys J 674, 984-996 (2008).

2. D. J. Hollenbach, A. G. G. M. Tielens, Photodissociation regions in the interstellar medium of galaxies. Rev Mod Phys 71, 173-230 (1999).

3. A. Usero, S. García-Burillo, A. Fuente, J. Martín-Pintado, N. J. Rodríguez-Fernández, Molecular gas chemistry in AGN. A&A 419, 897-912 (2004).

4. R. Meijerink, M. Spaans, F. P. Israel, Diagnostics of irradiated dense gas in galaxy nuclei. A&A 461, 793-811 (2007).

5. S. García-Burillo et al., Molecular line emission in NGC 1068 imaged with ALMA*. A&A 567, A125 (2014).

6. A. Labiano et al., Fueling the central engine of radio galaxies. A&A 564, A128 (2014). 7. L. R. Liu et al., Building one molecule from a reservoir of two atoms. Science 360, 900-

903 (2018). 8. Y.-P. Chang et al., Specific chemical reactivities of spatially separated 3-Aminophenol

conformers with cold Ca+ ions. Science 342, 98-101 (2013). 9. P. Puri et al., Synthesis of mixed hypermetallic oxide BaOCa+ from laser-cooled

reagents in an atom-ion hybrid trap. Science 357, 1370-1375 (2017). 10. S. Willitsch, M. T. Bell, A. D. Gingell, S. R. Procter, T. P. Softley, Cold reactive

collisions between laser-cooled ions and velocity-selected neutral molecules. Phys Rev Lett 100, 043203 (2008).

11. J. Deiglmayr, A. Göritz, T. Best, M. Weidemüller, R. Wester, Reactive collisions of trapped anions with ultracold atoms. Phys Rev A 86, 043438 (2012).

12. F. H. J. Hall, S. Willitsch, Millikelvin reactive collisions between sympathetically cooled molecular ions and laser-cooled atoms in an ion-atom hybrid trap. Phys Rev Lett 109, 233202 (2012).

13. C. Savage, L. M. Ziurys, Ion chemistry in photon-dominated regions: Examining the [HCO+]/[HOC+]/[CO+] chemical network. Astrophys J 616, 966-975 (2004).

14. L. M. Ziurys, A. J. Apponi, Confirmation of interstellar HOC+: Reevaluating the [HCO]+/[HOC]+ abundance ratio. Astrophys J 455, L73-L76 (1995).

15. J. Armijos-Abendaño et al., On the effects of UV photons/X-Rays on the chemistry of the Sgr B2 cloud. Astrophys J 895, 57 (2020).

16. A. Fuente et al., On the chemistry and distribution of HOC+ in M 82. A&A 492, 675-684 (2008).

17. A. Fuente et al., Photon-dominated chemistry in the nucleus of M82: Widespread HOC+ emission in the inner 650 parsec disk. Astrophys J 619, L155-L158 (2005).

18. S. Martín, J. Martín-Pintado, S. Viti, Photodissociation chemistry footprints in the starburst galaxy NGC 253. Astrophys J 706, 1323-1330 (2009).

19. R. Aladro et al., Molecular gas in the northern nucleus of Mrk 273: Physical and chemical properties of the disc and its outflow. A&A 617, A20 (2018).

20. J. Armijos-Abendaño, J. Martín-Pintado, M. A. Requena-Torres, S. Martín, A. Rodríguez-Franco, 3-mm spectral line survey of two lines of sight towards two typical cloud complexes in the Galactic Centre. Mon Not R Astron Soc 446, 3842-3862 (2014).

Page 37: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

21. R. C. Woods et al., The [HCO+]/[HOC+] abundance ratio in molecular clouds. 270, 583-588 (1983).

22. C. G. Freeman, M. J. McEwan, A selected-ion flow tube study of the C+ + H2O reaction. Int J Mass Spectrom Ion Process 75, 127-131 (1987).

23. J. R. Flores, A. B. González, The role of the excited electronic states in the C+ + H2O reaction. J Chem Phys 128, 144310 (2008).

24. Å. Larson, A. E. Orel, Electronic resonant states of HCO and HOC. Phys Rev A 80, 062504 (2009).

25. D. M. Sonnenfroh, R. A. Curtis, J. M. Farrar, Collision complex formation in the reaction of C+ with H2O. J Chem Phys 83, 3958-3964 (1985).

26. J. B. Marquette, B. R. Rowe, G. Dupeyrat, G. Poissant, C. Rebrion, Ion-polar-molecule reactions: A CRESU study of He+, C+, N+ + H2O, NH3 at 27, 68 and 163 K. Chem Phys Lett 122, 431-435 (1985).

27. J. O. Martinez et al., Gas Phase Study of C+ Reactions of Interstellar Relevance. Astrophys J 686, 1486-1492 (2008).

28. D. J. Defrees, A. D. McLean, E. Herbst, Calculations concerning the HCO+/HOC+ abundance ratio in dense interstellar clouds. Astrophys J 279, 322-334 (1984).

29. Y. Ishikawa, R. C. Binning, T. Ikegami, Direct ab initio molecular dynamics study of C++H2O. Chem Phys Lett 343, 413-419 (2001).

30. M. A. Smith, S. Schlemmer, J. von Richthofen, D. Gerlich, HOC+ + H2 isomerization

rate at 25 K: Implications for the observed [HCO+]/[HOC+] ratios in the interstellar medium. Astrophys J 578, L87-L90 (2002).

31. B. R. Rowe, J. B. Marquette, CRESU studies of ion/molecule reactions. Int J Mass Spectrom Ion Process 80, 239-254 (1987).

32. D. Gerlich, M. Smith, Laboratory astrochemistry: studying molecules under inter- and circumstellar conditions. Phys Scripta 73, C25-C31 (2005).

33. W. Paul, Electromagnetic traps for charged and neutral particles. Rev Mod Phys 62, 531-540 (1990).

34. T. Yang et al., Optical control of reactions between water and laser-cooled Be+ ions. J Phys Chem Lett 9, 3555-3560 (2018).

35. G. K. Chen et al., Isotope-selective chemistry in the Be+(2S1/2) + HOD → BeOD+/BeOH+ + H/D reaction. Phys Chem Chem Phys 21, 14005-14011 (2019).

36. D. J. Wineland, W. M. Itano, Laser cooling of atoms. Phys Rev A 20, 1521-1540 (1979). 37. E. R. Hudson, Sympathetic cooling of molecular ions with ultracold atoms. EPJ Tech

Instrum 3, 8 (2016). 38. E. R. Hudson, Method for producing ultracold molecular ions. Phys Rev A 79, 032716

(2009). 39. Z. Fang, V. H. S. Kwong, J. Wang, W. H. Parkinson, Measurements of radiative-decay

rates of the 2s22p(2P0)-2s2p2(4P) intersystem transition of C+. Phys Rev A 48, 1114-1122 (1993).

40. D. Patterson, J. Rasmussen, J. M. Doyle, Intense atomic and molecular beams via neon buffer-gas cooling. New Journal of Physics 11, 055018 (2009).

Page 38: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

41. S. J. Schowalter, K. Chen, W. G. Rellergert, S. T. Sullivan, E. R. Hudson, An integrated ion trap and time-of-flight mass spectrometer for chemical and photo-reaction dynamics studies. Rev Sci Instrum 83, 043103 (2012).

42. C. Schneider, S. J. Schowalter, K. Chen, S. T. Sullivan, E. R. Hudson, Laser-Cooling-Assisted mass spectrometry. Phys Rev Appl 2, 034013 (2014).

43. G. Mauclaire et al., Radiative lifetimes for an ion of astrophysical interest: HCO+. Int J Mass Spectrom Ion Process 149-150, 487-497 (1995).

44. N. G. Adams, D. Smith, D. Grief, Reactions of HnCO+ ions with molecules at 300 K. Int J Mass Spectrom Ion Phys 26, 405-415 (1978).

45. A. J. Chalk, L. Radom, Proton-Transport catalysis:  A systematic study of the rearrangement of the isoformyl cation to the formyl cation. J Am Chem Soc 119, 7573-7578 (1997).

46. X. Hu, W. L. Hase, T. Pirraglia, Vectorization of the general Monte Carlo classical trajectory program VENUS. J Comput Chem 12, 1014-1024 (1991).

47. Q. Yao, C. Xie, H. Guo, Competition between proton transfer and proton isomerization in the N2 + HOC+ reaction on an Ab Initio-based global potential energy surface. J Phys Chem A 123, 5347-5355 (2019).

48. D. McElroy et al., The UMIST database for astrochemistry 2012*. A&A 550, A36 (2013).

49. M. Röllig et al., A photon dominated region code comparison study*. A&A 467, 187-206 (2007).

50. N. R. Hutzler, H.-I. Lu, J. M. Doyle, The Buffer Gas Beam: An Intense, Cold, and Slow Source for Atoms and Molecules. Chem Rev 112, 4803-4827 (2012).

51. P. Zhao, Z. Xiong, L. Jie, L. He, B. Lu, Magneto-Optical Trapping of Ytterbium Atoms with a 398.9 nm Laser. Chin Phys Lett 25, 3631-3634 (2008).

52. C. G. Freeman, J. S. Knight, J. G. Love, M. J. McEwan, The reactivity of HOC+ and the proton affinity of CO at O. Int J Mass Spectrom Ion Process 80, 255-271 (1987).

53. V. G. Anicich, Evaluated Bimolecular Ion‐Molecule Gas Phase Kinetics of Positive Ions for Use in Modeling Planetary Atmospheres, Cometary Comae, and Interstellar Clouds. J Phys Chem Ref Data 22, 1469-1569 (1993).

54. D. J. Russell, NIST Computational Chemistry Comparison and Benchmark Database list of experimental proton affinities. (2019).

55. Q. Yao, C. Xie, H. Guo, Competition between proton transfer and proton isomerization in the N2 + HOC+ reaction on an ab initio-based global potential energy surface. J. Phys. Chem. A 123, 5347-5355 (2019).

56. NIST, Computational Chemistry Comparison and Benchmark Database, NIST Standard Reference Database Number 101, Release 17b, September 2015, Editor: Russell D. Johnson III, http://cccbdb.nist.gov/.

57. G. Knizia, T. B. Adler, H.-J. Werner, Simplified CCSD(T)-F12 methods: Theory and benchmarks. J Chem Phys 130, 054104 (2009).

58. K. A. Peterson, T. B. Adler, H.-J. Werner, Systematically convergent basis sets for explicitly correlated wavefunctions: The atoms H, He, B–Ne, and Al–Ar. J Chem Phys 128, 084102 (2008).

Page 39: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

59. A. J. Stone, The Theory of Intermolecular Forces. (Oxford University press, ed. Second Edition, 2013).

60. A. D. Buckingham, Permanent and induced molecular moments and long-range intermolecular forces. Adv. Chem. Phys. 12, 107-142 (1967).

61. B. Jiang, J. Li, H. Guo, Potential energy surfaces from high fidelity fitting of ab initio points: The permutation invariant polynomial-neural network approach. Int. Rev. Phys. Chem. 35, 479-506 (2016).

62. X. Hu, W. L. Hase, T. Pirraglia, Vectorization of the general Monte Carlo classical trajectory program VENUS. J. Comput. Chem. 12, 1014-1024 (1991).

63. J.-i. Yamamoto, Ab initio molecular dynamics simulation on H2O+C+ reaction. J. Mol. Struct. THEOCHEM 957, 55-60 (2010).

64. G. Mauclaire et al., Radiative lifetimes for an ion of astrophysical interest: HCO+. Int. J. Mass Spectro. Ion Proc. 149-150, 487-497 (1995).

65. R. Wester et al., Relaxation dynamics of deuterated formyl and isoformyl cations. J. Chem. Phys. 116, 7000-7011 (2002).

66. J. A. DeVine et al., Encoding of vinylidene isomerization in its anion photoelectron spectrum. Science 358, 336-339 (2017).

67. M. Mladenović, S. Schmatz, Theoretical study of the rovibrational energy spectrum and the numbers and densities of bound vibrational states for the system HCO+/HOC+. J. Chem. Phys. 109, 4456-4470 (1998).

Page 40: Isomer-specific kinetics of the C + H O reaction at the … + H2O... · 2020. 11. 9. · Isomer-specific kinetics of the C+ + H 2O reaction at the temperature of interstellar clouds

Acknowledgements

All authors declare that they have no competing interests. All data needed to evaluate the

conclusions in the paper are present in the paper and Supplementary Materials. T.Y. and G.C.

took the experimental data, A.L., Q.Y, and H.G. performed the theory calculations, T.Y., G.C.,

A.S. E.R.H. and W.C.C. designed the experiment and analyzed the experimental data, and all

authors participated in the preparation of the manuscript. The authors are indebted to

Christian Schneider and David Patterson for helpful technical advice. This work was supported

by the Air Force Office of Scientific Research grant nos. FA9550-16-1-0018 and FA9550-18-

1-0413. A. L. acknowledges the support of the Key Science and Technology Innovation Team

of Shanxi Province (2017KCT-37). H. G. thanks the Alexander von Humboldt Foundation for

a Humboldt Research Award.