photoelectron spectroscopy of hc4n− · and dense molecular clouds; because nitrogen is a major...

10
Photoelectron spectroscopy of HC4N− Kristen M. Vogelhuber, Scott W. Wren, Christopher J. Shaffer, Robert J. McMahon, Anne B. McCoy et al. Citation: J. Chem. Phys. 135, 204307 (2011); doi: 10.1063/1.3663617 View online: http://dx.doi.org/10.1063/1.3663617 View Table of Contents: http://jcp.aip.org/resource/1/JCPSA6/v135/i20 Published by the American Institute of Physics. Related Articles New view of the ICN A continuum using photoelectron spectroscopy of ICN− J. Chem. Phys. 136, 044313 (2012) Nuclear motion captured by the slow electron velocity imaging technique in the tunnelling predissociation of the S1 methylamine J. Chem. Phys. 136, 024306 (2012) Hydrogen bonds in the nucleobase-gold complexes: Photoelectron spectroscopy and density functional calculations J. Chem. Phys. 136, 014305 (2012) Zero kinetic energy photoelectron spectroscopy of jet cooled benzo[a]pyrene from resonantly enhanced multiphoton ionization J. Chem. Phys. 135, 244306 (2011) High-resolution threshold photoelectron study of the propargyl radical by the vacuum ultraviolet laser velocity- map imaging method J. Chem. Phys. 135, 224304 (2011) Additional information on J. Chem. Phys. Journal Homepage: http://jcp.aip.org/ Journal Information: http://jcp.aip.org/about/about_the_journal Top downloads: http://jcp.aip.org/features/most_downloaded Information for Authors: http://jcp.aip.org/authors Downloaded 31 Jan 2012 to 128.138.107.159. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

Upload: others

Post on 19-Oct-2020

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Photoelectron spectroscopy of HC4N− · and dense molecular clouds; because nitrogen is a major component of interstellar dust, cyanopolyynes are also im-portant constituents of

Photoelectron spectroscopy of HC4N−Kristen M. Vogelhuber, Scott W. Wren, Christopher J. Shaffer, Robert J. McMahon, Anne B. McCoy et al. Citation: J. Chem. Phys. 135, 204307 (2011); doi: 10.1063/1.3663617 View online: http://dx.doi.org/10.1063/1.3663617 View Table of Contents: http://jcp.aip.org/resource/1/JCPSA6/v135/i20 Published by the American Institute of Physics. Related ArticlesNew view of the ICN A continuum using photoelectron spectroscopy of ICN− J. Chem. Phys. 136, 044313 (2012) Nuclear motion captured by the slow electron velocity imaging technique in the tunnelling predissociation of theS1 methylamine J. Chem. Phys. 136, 024306 (2012) Hydrogen bonds in the nucleobase-gold complexes: Photoelectron spectroscopy and density functionalcalculations J. Chem. Phys. 136, 014305 (2012) Zero kinetic energy photoelectron spectroscopy of jet cooled benzo[a]pyrene from resonantly enhancedmultiphoton ionization J. Chem. Phys. 135, 244306 (2011) High-resolution threshold photoelectron study of the propargyl radical by the vacuum ultraviolet laser velocity-map imaging method J. Chem. Phys. 135, 224304 (2011) Additional information on J. Chem. Phys.Journal Homepage: http://jcp.aip.org/ Journal Information: http://jcp.aip.org/about/about_the_journal Top downloads: http://jcp.aip.org/features/most_downloaded Information for Authors: http://jcp.aip.org/authors

Downloaded 31 Jan 2012 to 128.138.107.159. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

Page 2: Photoelectron spectroscopy of HC4N− · and dense molecular clouds; because nitrogen is a major component of interstellar dust, cyanopolyynes are also im-portant constituents of

THE JOURNAL OF CHEMICAL PHYSICS 135, 204307 (2011)

Photoelectron spectroscopy of HC4N−

Kristen M. Vogelhuber,1 Scott W. Wren,1 Christopher J. Shaffer,2 Robert J. McMahon,2

Anne B. McCoy,3 and W. Carl Lineberger1,a)

1JILA and Department of Chemistry and Biochemistry, University of Colorado, 440 UCB, Boulder,Colorado 80309, USA2Department of Chemistry, University of Wisconsin, Madison, Wisconsin 53706, USA3Department of Chemistry, The Ohio State University, Columbus, Ohio 43210, USA

(Received 6 October 2011; accepted 2 November 2011; published online 30 November 2011)

We report the 364-nm photoelectron spectrum of HC4N−. We observe electron photodetachmentfrom the bent X̃2A′′ state of HC4N− to both the near-linear X̃3A′′ and the bent ã 1A′ states ofneutral HC4N. We observe an extended, unresolved vibrational progression corresponding to X̃3A′′

← X̃2A′′ photodetachment, and we measure the electron affinity (EA) of the X̃3A′′ state of HC4N tobe 2.05(8) eV. Photodetachment to the bent ã 1A′ state results in a single intense origin peak at a bind-ing energy of 2.809(4) eV, from which we determine the singlet-triplet splitting (�EST) of HC4N:0.76(8) eV. For comparison and to aid in the interpretation of the HC4N− spectrum, we also reportthe 364-nm photoelectron spectra of HCCN− and DCCN−. Improved signal-to-noise over the previ-ous HCCN− and DCCN− photoelectron spectra allows for a more precise determination of the EAsand �ESTs of HCCN and DCCN. The EAs of HCCN and DCCN are measured to be 2.001(15) eVand 1.998(15) eV, respectively; �EST(HCCN) is 0.510(15) eV and �EST(DCCN) is 0.508(15) eV.These results are discussed in the context of other organic carbene chains. © 2011 American Instituteof Physics. [doi:10.1063/1.3663617]

I. INTRODUCTION

The cyanopolyyne (HCnN) HC4N is a carbene (:CXY),as it contains a neutral divalent carbon atom with two elec-trons occupying two orbitals that lie close in energy. Car-benes are highly reactive and serve as important intermedi-ates in chemical reactions, and their reactivity is affected bywhether the carbene exists as a singlet or a triplet. The car-bene’s substituents determine the electronic ground state ofthe molecule, as well as the difference in energy between thesinglet and triplet states, known as the singlet-triplet splitting(�EST).1 The simplest carbene, methylene (CH2), possesses atriplet ground state because the energy separation between theσ and π orbitals is less than the electron correlation energy.2

The donation of π electrons, e.g., by halogen atoms, tends toincrease the separation between the σ and π orbitals, lead-ing to a singlet ground state.3, 4 Although CN is considereda “pseudohalogen,” it does not donate π electrons. Previousexperiments have shown that HCCN has a triplet ground statewith a relatively large �EST.5

In addition to their fundamental importance as car-benes, cyanopolyynes have attracted interest because oftheir significance in astrophysics. These species are abun-dant in circumstellar environments of carbon-rich starsand dense molecular clouds; because nitrogen is a majorcomponent of interstellar dust, cyanopolyynes are also im-portant constituents of molecular clouds.6 Odd cyanopolyynechains (HCnN where n = odd) up to HC11N have beendetected in space,7 and microwave rotational spectra of

a)Author to whom correspondence should be addressed. Electronic mail:[email protected].

odd cyanopolyynes up to n = 17 have been recorded.8

These studies have established that, like the isoelectronicpolyacetylenes HC2nH, odd cyanopolyynes possess a linearsinglet ground state. The chain has alternating single andtriple bonds, giving rise to alternating bond lengths.9

There have been many reports of the observation of oddcyanopolyynes in circumstellar and laboratory environments;however, the same do not exist for the even cyanopolyynes,implying that odd cyanopolyynes are more stable than evenspecies.6 In fact, McCarthy and Thaddeus report a factor of10–100 lower abundance of the even cyanopolyynes HC4N,HC6N, and HC8N as compared to odd cyanopolyynes of sim-ilar length.10 They attribute this discrepancy to the greater sta-bility of the closed-shell singlet odd cyanopolyynes as com-pared to the reactive triplet even cyanopolyynes, and to thefact that the mechanisms of carbon chain formation11 favorthe formation of odd cyanopolyynes. Furthermore, while thegeometry of odd cyanopolyynes has been characterized, thegeometry of the lowest energy isomers of even cyanopolyynesis not as well understood.

HC4N is of particular interest to experimental and the-oretical groups alike because it is one of the smallest sys-tems calculated to possess multiple low-lying, highly po-lar carbene isomers.12 McCarthy and co-workers generatedtwo singlet isomers of HC4N in a supersonic molecularbeam and detected them with Fourier transform microwave(FTM) spectroscopy: a cyanovinylidene carbene structure,NC(H)C=C=C:,13 and a ring-chain structure.14 A number oftheoretical works have been aimed at understanding the rela-tive stability of various isomers and electronic states;6, 12, 15–18

for the most part, these studies found that the ring-chain (cy-clopropenylidene) singlet isomer of HC4N is lower in energy

0021-9606/2011/135(20)/204307/9/$30.00 © 2011 American Institute of Physics135, 204307-1

Downloaded 31 Jan 2012 to 128.138.107.159. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

Page 3: Photoelectron spectroscopy of HC4N− · and dense molecular clouds; because nitrogen is a major component of interstellar dust, cyanopolyynes are also im-portant constituents of

204307-2 Vogelhuber et al. J. Chem. Phys. 135, 204307 (2011)

SCHEME 1.

than the linear chain triplet isomer, but the results were de-pendent upon the method employed.

There is also question as to whether the structure of thetriplet ground state of the chain isomer of HC4N is linear orbent; it is useful to compare HCC–C–CN with several othertriplet carbenes (Scheme 1). In the case of triplet HCC–C–CCH, experiment and theory agree that the triplet is axiallysymmetric.19 Sanov et al. concluded that triplet NC–C–CN islinear or quasilinear.20 Tang et al. reported the first observa-tion of the triplet ground state of the chain isomer of HCC–C–CN in the laboratory using FTM spectroscopy; with the aidof calculations, they concluded that triplet HC4N has a verysmall barrier to linearity (25 cm−1 at CCSD(T)/6-311G(d,p))and thus can be considered a linear molecule.21 Aoki et al.,also using CCSD(T), predict a linear triplet.16 Later, Cer-nicharo et al. reported the observation of the linear tripletform of HC4N in the carbon-rich IRC + 10216 envelope.22

Here, we present a photoelectron spectroscopy studyof the HC4N− anion. From the 364-nm photoelectronspectrum of HC4N−, we obtain the first experimental mea-surement of the electron affinity (EA) and �EST of the chainisomer of HC4N. For comparison and to aid in our under-standing of the HC4N− spectrum, we also report the 364-nmphotoelectron spectra of HCCN− and DCCN− with improvedsignal-to-noise over the previously published spectrum.5 Theimprovement in the HCCN− and DCCN− photoelectronspectra allows for better resolution of the vibrational pro-gressions and a more precise determination of the EAs and�ESTs of HCCN and DCCN. We compare the HC4N− andHCCN− (DCCN−) spectra and consider them in the contextof a broader set of carbene carbon chains.

II. EXPERIMENTAL METHOD

The negative ion photoelectron spectrometer used in thisexperiment has been described in detail in Refs. 23–25. Theapparatus consists of four main sections: an ion source, a massfilter, an interaction region with crossed laser and ion beams,and an electrostatic electron kinetic energy analyzer. Nega-tive ions are formed in a flowing afterglow ion source. A mi-crowave discharge containing trace amounts of O2 gas in Hebuffer gas (∼0.4 Torr) generates atomic oxygen radical an-ion, O−. To generate HC4N− anions, tetrolonitrile (H3C4N) isadded downstream of O− (Scheme 2). Synthetic details for thepreparation of tetrolonitrile are presented in supplementarymaterial.26 To make HCCN− anions, acetonitrile (H3CCN) isadded downstream of O−: O− + H3CCN → HCCN− + H2O.We produce DCCN− using the analogous reaction and begin-ning with D3CCN.

SCHEME 2.

Collisions with He buffer gas cool the ions to ∼300 K.The flow tube can be further cooled with a liquid nitrogenjacket to obtain a “cold spectrum” of ions with tempera-tures near 150 K. Anions are extracted into a differentiallypumped region and are accelerated to 735 eV before enter-ing a Wien velocity filter with a mass resolution of m/�m ∼60.27 With this resolution, we cannot cleanly resolve HC4N−

and H2C4N−, both of which are produced in the flowing after-glow ion source; however, we select the low-mass side of them/z ∼ 63–64 peak to obtain the spectrum of HC4N−. We alsoobtained the photoelectron spectrum of neat H2C4N− (gener-ated by first reacting O− with CH4 to produce OH−, whichthen reacts with H3C4N to produce H2C4N−) to ensure thatthere was no contribution from H2C4N− in the HC4N− spec-tra. The photoelectron spectrum of H2C4N− is a broad, fea-tureless progression that appears at lower binding energy thanHC4N−, as shown in Fig. S1.26

The mass-selected ion beam (typically 70 pA) is decel-erated to 35 eV and focused into the laser interaction region.Here, the 1 W output from a single-mode continuous-waveargon ion laser operating at 364 nm (3.40814 eV) is built upto ∼100 W of circulating power in an optical buildup cav-ity located within the vacuum system. Photodetachment ofan electron yields the neutral species, as depicted for HC4N−

photodetachment in Scheme 3. Photoelectrons ejected in thedirection orthogonal to both the laser and ion beams entera hemispherical energy analyzer. The photoelectron signalis recorded as a function of electron kinetic energy with aposition-sensitive detector. The energy analyzer has a reso-lution of ∼11 meV under the conditions used for the presentexperiments.

The electron kinetic energy (eKE) can be convertedto electron binding energy (eBE) through the relationship

SCHEME 3.

Downloaded 31 Jan 2012 to 128.138.107.159. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

Page 4: Photoelectron spectroscopy of HC4N− · and dense molecular clouds; because nitrogen is a major component of interstellar dust, cyanopolyynes are also im-portant constituents of

204307-3 Photoelectron spectroscopy of HC4N− J. Chem. Phys. 135, 204307 (2011)

eBE = hν – eKE. The absolute kinetic energy scale iscalibrated23, 24, 28 before and after each experiment using thewell-known EA of atomic oxygen.29 Additionally, the energyscale is corrected for a slight linear compression (<1%)23

using the photoelectron spectrum of O2−, which provides a

number of known transitions spanning the photoelectron en-ergy range.30, 31 After making these corrections and account-ing for the resolution of the spectrometer and rotational peakprofiles, absolute electron binding energies can be determinedwith an accuracy of better than 5 meV.

A rotatable half-wave plate positioned outside thebuildup cavity varies the polarization of the photodetachmentradiation in order to control the angle θ between the electricfield vector of the laser beam and the photoelectron collectionaxis. The photoelectron angular distribution is described bythe equation32

I (θ ) = σ0

4π(1 + βP2(cos θ )), (1)

where σ 0 is the total photodetachment cross section, β is theanisotropy parameter, and P2(cos θ ) is the second Legendrepolynomial. We measure the anisotropy parameter explicitlyby recording the photoelectron signal at the kinetic energyof one suitable intense peak in the photoelectron spectrum asa function of θ (between θ = 0◦ and θ = 360◦ in steps of10◦). The photoelectron angular distribution is fit with Eq. (1),and full spectra collected at θ = 0◦ and θ = 90◦ are scaledto match β at the energy at which it was measured. Sepa-rately, we collect a photoelectron spectrum at θ = 54.7◦ (theso-called magic angle), where the photoelectron intensity isindependent of β and directly reflects the relative photode-tachment cross section.

III. THEORETICAL METHODS

All electronic structure calculations were performed us-ing the GAUSSIAN 03 program package.33 Optimized geome-tries, harmonic vibrational frequencies, and normal mode co-ordinates were calculated at the ROMP2 level of theory34

with the 6-311++G(d,p) basis set35 for the X̃2A′′ states ofHC4N− and HCCN− (DCCN−) as well as for the X̃3A′′ andã 1A′ states of the neutrals HC4N and HCCN (DCCN). Allmolecules were constrained to Cs symmetry. For the reportedscans along the CCC (HC4N) or HCC (HCCN) angle, Figs. 3and S4 (Ref. 26), the geometries were further constrained sothat all but the scanned angles were 180◦, and the bond lengthswere constrained to their optimized values on the anion sur-face when all but one of the angles was 180◦.

We employ a Franck-Condon analysis of the vibrationalstructure in the photoelectron spectra to identify the activevibrational modes and the geometry change upon photode-tachment. The Franck-Condon profiles of the photoelectronspectra are simulated with the PESCAL program,31 using thecalculated geometries, normal mode vectors, and vibrationalfrequencies of the anion and neutral states. The normalmodes and the Duschinsky J′′ matrix and K′′ displacementsare calculated. The Franck-Condon factors are computed inthe harmonic oscillator approximation including Duschinskyrotation using the Sharp-Rosenstock-Chen method.36 The

individual vibronic peak contours are simulated by a Gaus-sian function with a full-width half-maximum of 11 meV,consistent with instrumental resolution.

Activity of a large-amplitude bending vibration uponphotodetachment of the bent HC4N− anion to the quasi-linear X̃3A′′ state of HC4N renders our standard indepen-dent harmonic oscillator approximation unsuitable for simu-lating X̃3A′′ ← X̃2A′′ photodetachment of HC4N−. A moreappropriate treatment of X̃3A′′ ← X̃2A′′ photodetachmentwould utilize a curvilinear HCC–C–CN bending coordinateand a vibration-rotation Hamiltonian that allows for large dis-placements of the bending coordinate.37 However, this moresophisticated treatment is beyond the scope of this work.In our standard Franck-Condon analysis using the Sharp-Rosenstock-Chen method, nonphysical K′′ displacements arecomputed that result in an inadequate simulation of X̃3A′′

← X̃2A′′ photodetachment. For this reason, we will onlypresent partial simulations of the HC4N− and HCCN−

(DCCN−) photoelectron spectra; full simulations are includedin Figs. S3 and S5.26

IV. RESULTS

A. HC4N−

1. Photoelectron spectra of HC4N−

It is instructive to first consider the valence electronicstructure of HC4N (Fig. 1) and its implications for the pho-toelectron spectrum of HC4N−. Like HCCN−,5 the groundstate of HC4N− is X̃2A′′. The excess charge resides primarilyon the third carbon atom, and the anion is bent. If an elec-tron is photodetached from the in-plane σ orbital to form theX̃3A′′ ground state of HC4N, the molecule straightens to be-come quasilinear. However, if an electron is removed fromthe out-of-plane π orbital to form the ã 1A′ excited state ofHC4N, the molecule remains bent in much the same geom-etry as the anion. Intuitively, we expect the large geometrychange upon photodetachment of the bent anion to the quasi-linear triplet ground state to result in an extended vibrationalprogression at low binding energy in the photoelectron spec-trum of HC4N−. We expect multiple vibrational modes—especially bending modes involving the carbene center—to

FIG. 1. Generalized valence bond diagrams of the X̃2A′′ state of HCCN−,the X̃3A′′ and ã 1A′ states of HCCN,5 as well as the X̃2A′′ state of HC4N−and the X̃3A′′ and ã 1A′ states of HC4N.

Downloaded 31 Jan 2012 to 128.138.107.159. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

Page 5: Photoelectron spectroscopy of HC4N− · and dense molecular clouds; because nitrogen is a major component of interstellar dust, cyanopolyynes are also im-portant constituents of

204307-4 Vogelhuber et al. J. Chem. Phys. 135, 204307 (2011)

FIG. 2. Photoelectron spectra of HC4N−. The 364-nm spectra were collectedat the magic angle at 150 K (black trace) and 300 K (blue trace). The partialFranck-Condon simulation of the 150 K HC4N− photoelectron spectrum isshown in purple, with gray sticks representing individual vibronic transitions.The calculated geometries, frequencies, and K′′ displacements used in theã 1A′ ← X̃2A′′ photodetachment simulation are listed in Tables II and III.The simulation uses the experimental binding energy of the ã 1A′ state ofHC4N. The equilibrium structures of the X̃3A′′ and ã 1A′ states of HC4N arealso shown.

be active upon photodetachment and to contribute intensity tothe triplet manifold. Since both the anion and the neutral sin-glet are bent, we expect significant Franck-Condon overlapbetween the ground vibrational state of HC4N− and the sin-glet state of HC4N. This will result in an intense “origin” peakcorresponding to the 0-0 transition, and it will appear higherin energy than the triplet state.

The 364-nm, magic angle photoelectron spectrum ofHC4N− is shown in Fig. 2. The spectrum was collected at bothroom temperature (300 K, blue trace) and at 150 K (black

trace). As anticipated, we observe a broad progression atlow binding energy, corresponding to photodetachment to theX̃3A′′ ground state of HC4N. The large geometry change thatoccurs upon X̃3A′′ ← X̃2A′′ photodetachment precludes theobservation of the origin peak. Additionally, because there aremultiple active modes upon photodetachment—particularlythe low frequency HCC—C–CN bend—we observe a con-gested spectrum and are unable to resolve individual vibronicpeaks in the triplet envelope. Therefore, we cannot measurevibrational frequencies of the X̃3A′′ ground state of HC4N.Though we do not resolve an origin peak from which we candetermine the EA, comparison of the 300 K and 150 K spectrashows that the EA of HC4N is 2.05(8) eV (Table I). The 300 Kspectrum displays a sharp onset at 2.00 eV; in the 150 K spec-trum, the intensity between 2.00 and 2.05 eV is suppressed.This indicates that the intensity observed between 2.00 and2.05 eV in the 300 K spectrum results from hot bands or vi-bronic transitions arising from vibrationally excited anions. Inthe 150 K spectrum, we observe intensity above the baselinebeginning at ∼2.05 eV, and we begin to resolve two featuresbetween 2.00 and 2.15 eV (see inset of Fig. 2). Unfortunately,the substantial geometry change in coupled modes means thatthe independent mode Franck-Condon simulation is not evenqualitatively reliable4, 38 and cannot help in estimating a lowerbound for the origin (EA) of the triplet state. Because the fea-tures in the origin region are not well resolved, we assign theEA of HC4N to be 2.05 eV—where we see appreciable in-tensity above the baseline in the cooled spectrum—with anuncertainty that encompasses both the shelf in the 300 K spec-trum and the two small features in the 150 K spectrum. Thus,the EA of HC4N is measured to be 2.05(8) eV.

At higher binding energy, a sharp feature appears at eBE= 2.809(4) eV. The polarization dependence of the photoelec-tron spectrum of HC4N− is given in Table I and is shownin Fig. S2;26 the difference in the β anisotropy parametersbetween the broad progression beginning at 2.05(8) eV andthe sharp feature at eBE = 2.809(4) eV indicates that thesefeatures result from two different electronic states of HC4N.The peak at eBE = 2.809(4) eV corresponds to the origin ofthe ã 1A′ excited state of HC4N. Because the geometry of theX̃2A′′ state of HC4N− and the ã 1A′ excited state of HC4N

TABLE I. Experimental and calculated electron affinities (EA), vertical detachment energies (VDE), and singlet-tripletsplittings (�EST) of HCCN and HC4N given in eV.

X̃3A′′ ← X̃2A′′ ã 1A′ ← X̃2A′′

EA VDE β (0-0) VDE β �EST

HC4N Experimenta 2.05(8) 2.57(3) +0.2 2.809(4) 2.809(4) −0.3 0.76(8)Calculatedb 2.229 2.533 2.771 2.784 0.542

HCCN Experimenta 2.001(15) 2.36(2) −0.25 2.511(4) 2.512(4) −0.62 0.510(15)Experimentc 2.003(14) −0.37d 2.518(8) −0.77 0.515(16)Calculatedb 1.746 2.185 2.512 2.562 0.765

DCCN Experimenta 1.998(15) 2.37(2) −0.28 2.506(4) 2.506(4) −0.63 0.508(15)Experimentc 2.009(20) −0.33d 2.527(18) −0.78 0.518(27)Calculatedb 1.749 2.184 2.509 2.560 0.761

aExperiment, this work.bROMP2/6-311++G(d,p).cExperiment, Nimlos et al. (Ref. 5).dβ value listed is the average of the β values of the peaks in the triplet progression.

Downloaded 31 Jan 2012 to 128.138.107.159. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

Page 6: Photoelectron spectroscopy of HC4N− · and dense molecular clouds; because nitrogen is a major component of interstellar dust, cyanopolyynes are also im-portant constituents of

204307-5 Photoelectron spectroscopy of HC4N− J. Chem. Phys. 135, 204307 (2011)

TABLE II. Calculated (ROMP2/6-311++G(d,p)) equilibrium geometry of X̃2A′′ HC4N− and the calculated net ge-ometry change upon photodetachment to the X̃3A′′ and ã 1A′ states of HC4N. Bond lengths are given in units ofangstrom (Å), and bond angles are given in units of degree (◦). Boldfaced entries highlight the internal coordinatesthat undergo significant change upon photodetachment.

HC4N− Geometry change

Internal coordinate X̃2A′′ X̃3A′′ ← X̃2A′′ ã 1A′ ← X̃2A′′

H–C1 1.062 +0.003 +0.005C1–C2 1.223 −0.003 +0.011C2–C3 1.377 −0.059 − 0.008C3–C4 1.402 −0.023 +0.004C4–N 1.167 −0.027 +0.017� H–C1–C2 178.6 +0.4 − 1.9� C1–C2–C3 173.9 +1.5 − 2.8� C2–C3–C4 119.6 +28.2 − 1.7� C3–C4–N 173.6 +4.5 − 1.6

are both bent, we observe a very prominent ã 1A′ origin peak.We measure the singlet-triplet splitting (�EST) of HC4N to be0.76(8) eV (Table I).

The most striking—and most curious—difference be-tween the 300 K and 150 K HC4N spectra shown in Fig. 2 isthe appearance of an additional sharp feature at 3.106(4) eVin the 150 K spectrum. The intensity of this peak, which isseparated from the ã 1A′ origin by 2395 cm−1, fluctuated withrespect to the X̃3A′′ and ã 1A′ features throughout the day andfrom day to day, and it was never observed at room tempera-ture. The eBE of this feature suggests that it is a CN stretchof the singlet state, but such a temperature dependence of acombination band is not typically seen in photoelectron spec-tra. A measurement of the β of this peak would lend insightinto the source of this feature, but unfortunately the intensityfluctuations of this peak made it difficult to obtain a reliable

β measurement of the peak (we measure β(3.106 eV) ∼ 0.3).As discussed in Sec. IV A 2., we employ calculations and sim-ulations to confirm that this feature does indeed correspond toactivation of CN stretch in the ã 1A′ state of HC4N.

2. Electronic structure calculations for HC4N−

As discussed in Sec. III, our standard Franck-Condonanalysis of the HC4N− photoelectron spectrum is inade-quate for modeling X̃3A′′ ← X̃2A′′ photodetachment. Us-ing ROMP2/6-311++G(d,p), we calculate the EA(HC4N) tobe significantly greater than the measured EA of 2.05(8) eV(Table I); Kalcher’s ACPF calculation (using the averagedcoupled pair functional in conjunction with the ROHF wavefunction for the anion and triplet state of the neutral, andthe CAS(2,2) wave function for the singlet state of the

TABLE III. Calculated (ROMP2/6-311++G(d,p)) unscaled, harmonic frequencies of X̃2A′′ HC4N− and the X̃3A′′and ã 1A′ states of HC4N. The measured CN stretch frequency is listed in italics below the calculated value. ComputedK′′ displacements (PESCAL) are also listed. The computed X̃3A′′ ← X̃2A′′ K′′ displacements do not accurately rep-resent the geometry change that takes place upon photodetachment and result in a Franck-Condon simulation that doesnot reproduce the observed triplet progression (Fig. S3) (Ref. 26).

HC4N− X̃3A′′ ← X̃2A′′ ã 1A′ ← X̃2A′′

Mode Description 2A′′, cm−1 �K′′ 3A′′, cm−1 �K′′ 1A′, cm−1

A′ ν1 CH stretch 3521.5 −0.0093727 3494.4 0.0025897 3486.5ν2 CN stretch 2159.8 −0.12460 2309.5 0.045273 2217.6

2395(40)a

ν3 CCC asym stretch 1937.7 −0.052916 1714.3 0.021984 2111.4ν4 CCC asym stretch 1215.3 −0.044035 1308.1 −0.027163 1247.1ν5 CCC sym stretch 812.1 −0.48678 651.6 0.033921 831.3ν6 HCC–C–CN bend 580.3 0.48465 353.7 −0.065517 579.5ν7 HCC–C–CN rocking 407.6 0.23376 214.0 0.0093059 409.5ν8 H wag 283.2 0.29055 589.1 −0.027086 507.8ν9 HCCC–C–N bend 153.6i 1.7090 286.0i 0.073406 142.4

A′′ ν10 H wag 403.4 0 491.8 0 808.2ν11 N–C–C bend 336.3 0 177.7 0 346.4ν12 NCC–C–CH bend 155.7 0 297.2 0 291.9

aExperiment, this work.

Downloaded 31 Jan 2012 to 128.138.107.159. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

Page 7: Photoelectron spectroscopy of HC4N− · and dense molecular clouds; because nitrogen is a major component of interstellar dust, cyanopolyynes are also im-portant constituents of

204307-6 Vogelhuber et al. J. Chem. Phys. 135, 204307 (2011)

neutral) also overestimates the EA by ∼250 meV.17 The cal-culated geometry change upon photodetachment is illustratedin Table II, and the calculated vibrational frequencies arelisted in Table III. As expected, a large geometry change iscalculated to occur upon X̃3A′′ ← X̃2A′′ photodetachment,which results in the observed extended vibrational progres-sion. Specifically, the C2–C3 bond length decreases by 0.059Å, while the � C2–C3–C4 increases by 28.2◦ as the moleculebecomes nearly linear. Previous studies seem to favor a lin-ear X̃3A′′ structure,16, 21, 22 while our calculations predict aslightly bent equilibrium geometry. Though we calculate theminimum to occur at a � C2–C3–C4 of 147.8◦ (Table II), aone-dimensional potential cut along the � C2–C3–C4 bendingcoordinate (Fig. 3) reveals that the X̃3A′′ state of HC4N is ex-tremely floppy with a barrier to linearity of only ∼24 meV.The one-dimensional cuts also reveal that as C2–C3–C4 be-comes more bent, the triplet and singlet states cross. The un-resolved nature of the triplet vibrational progression preventsus from determining the geometry of the X̃3A′′ state fromthe photoelectron spectrum. Activity of this large-amplitudebending motion is the source of the breakdown of the indepen-dent harmonic oscillator analysis (see Fig. S3 for the X̃3A′′

← X̃2A′′ photodetachment spectrum).26 Such failure of theharmonic, normal mode analysis has been observed in otherfloppy systems undergoing large amplitude geometry changesupon photodetachment.5, 38–40

The partial simulated HC4N− spectrum, which showsonly those features arising from ã 1A′ ← X̃2A′′ photodetach-ment, is shown in Fig. 2 in purple, with gray sticks represent-ing individual calculated vibronic transitions. The simulationuses the calculated geometries, K′′ displacements, and vibra-tional frequencies (Tables II and III). The calculated ã 1A′

origin, using either ROMP2 or ACPF,17 is lower than themeasured value of 2.809(4) eV (Table I); therefore, we usethe observed ã 1A′ origin in the simulation. The simulationof ã 1A′ ← X̃2A′′ photodetachment is very successful. Boththe HC4N− anion and the ã 1A′ excited state of HC4N arebent ( � C2–C3–C4 ∼ 120◦). None of the internal coordinatesof HC4N− undergo a significant change upon photodetach-ment to the ã 1A′ state, resulting in a very prominent originpeak and a modest vibrational progression. The simulationreproduces the temperature-dependent peak that appears at3.106(4) eV. The simulation confirms that this feature arisesprimarily from activation of the ν2 CN stretch as well as ac-tivity of ν3 (C1–C2–C3 asymmetric stretch).

We find no evidence of an open-shell singlet state in ourspectra. In the case of NCCCN, Sanov et al. calculate theopen-shell 1�g to lie 0.1–0.3 eV higher in energy than theclosed-shell 1A1, depending on the level of theory.20 The twosinglet states are degenerate at the linear NCCCN geome-try, but upon distortion Renner-Teller coupling results in en-ergy splitting to make the closed-shell singlet more stable. Inthe case of the oxyallyl diradical, the open-shell 1B2 is cal-culated to lie 1.346 eV higher in adiabatic energy than theclosed-shell X̃1A1.1, 41 To our knowledge, the energy of theopen-shell singlet of HC4N has not been computed. Thereis no indication of the open-shell singlet state in our spec-tra, though it is possible that it lies close in energy to theclosed-shell 1A′.

FIG. 3. One-dimensional potential cut along the HCC–C–CN bend coordi-nate. Energies are relative to the minimum on the anion surface. All internalangles except � C2–C3–C4 are constrained to be linear, while the remain-ing internal coordinates are set to the values that minimize the energy of theanion under these constraints. All curves were calculated using ROMP2/6-311++G(d,p) level of theory/basis set.

B. HCCN− and DCCN−

The new 364-nm, magic angle photoelectron spectra ofHCCN− and DCCN− are shown in Fig. 4. The data shown inFig. 4 display significantly improved signal-to-noise over theprevious data.5 As a result, we more clearly resolve transitionswithin the vibrational progressions. As briefly mentioned inconnection with HC4N−, it is extremely difficult to obtaina precise, reliable measurement of the EA of a moleculeexhibiting an extended vibrational progression. In such cases,the intensity of the origin peak is weak or even unobservable.Furthermore, normal mode analysis often fails to accuratelyrepresent the photoelectron spectrum, and we thereforecannot use calculated Franck-Condon factors to help as-sign the origin or to anticipate its relative intensity. LikeHC4N−, the photoelectron spectra of HCCN− and DCCN−

display extended vibrational progressions in the groundstate. Through comparison of the HCCN− and DCCN−

spectra, we assign the origins of HCCN− and DCCN− to thepeaks marked with arrows in Fig. 4. We measure the EA ofHCCN to be 2.001(15) eV and of DCCN to be 1.998(15) eV,in accord with the measurements of the previous work(Table I). We measure the �EST of HCCN as 0.510(15) eVand the �EST of DCCN to be 0.508(15) eV. The stated

Downloaded 31 Jan 2012 to 128.138.107.159. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

Page 8: Photoelectron spectroscopy of HC4N− · and dense molecular clouds; because nitrogen is a major component of interstellar dust, cyanopolyynes are also im-portant constituents of

204307-7 Photoelectron spectroscopy of HC4N− J. Chem. Phys. 135, 204307 (2011)

FIG. 4. Partial Franck-Condon simulations of the photoelectron spectra ofHCCN− and DCCN−. Green arrows mark the origin peaks, and insets ofthe origin regions show the measured EAs and uncertainties. The equilib-rium structures of the X̃3A′′ and ã 1A′ states of HCCN (DCCN) are alsoshown. The new photoelectron spectra collected in the present work dis-play improved signal-to-noise, enabling us to resolve additional peaks. Thecalculated geometries, frequencies, and K′′ displacements used in the ã 1A′← X̃2A′′ simulation are listed in Tables S1–S3.26 The simulations use themeasured binding energy of the ã 1A′ states of HCCN and of DCCN.

uncertainties assume that the origin peaks have been assignedcorrectly. However, it is notoriously difficult to assign theorigin of an extended progression, and the uncertainties givenin the previous study5 were definitely optimistic.

The photoelectron spectra of HCCN− and DCCN− ex-hibit several similarities to the HC4N− spectrum. Both HC4Nand HCCN (DCCN) possess a X̃3A′′ ground state that is sig-nificantly less bent than the anion. Upon X̃3A′′ ← X̃2A′′

photodetachment, the � H–C1–C2 increases by 35.0◦, and theC1–C2 bond decreases by 0.050 Å (Table S1).26 The large ge-ometry change from the bent anion to the quasilinear triplet42

results in an extended vibrational progression beginning at2.0 eV (Table I). The peaks in the X̃3A′′ HCCN progressionare spaced by roughly 400 cm−1, corresponding to activity ofthe ν5 HCC bending mode.5

As in the case of HC4N−, the large geometry changeleads to the failure of the harmonic normal mode analysis toadequately model the X̃3A′′ ← X̃2A′′ photodetachment pho-toelectron spectrum. One-dimensional potential cuts along theH–C–CN bend coordinate are presented in Fig. S4.26 The sim-

ulated X̃3A′′ ← X̃2A′′ spectra of HCCN− and DCCN− areshown in Fig. S5; the calculated geometries, K′′ displace-ments, and vibrational frequencies are given in Tables S1–S3.26 In the HCCN− (DCCN−) case, the large change inthe � H–C1–C2 is represented in the harmonic normal modeanalysis in Cartesian coordinates by a nonphysical changein the H–C1 bond length.38 This is manifested in the ap-pearance of a CH (CD) stretch progression in the predictedspectrum.

Like HC4N, the origin of the excited ã 1A′ state is themost prominent feature in the spectrum of HCCN. The� H–C1–C2 of both HCCN− and the ã 1A′ state of HCCNare strongly bent (∼110◦), resulting in an intense originpeak and a short vibrational progression. At higher bindingenergy, we observe an additional sharp feature at eBE= 2.747 eV, or 1903 cm−1 higher in energy than the ã 1A′

HCCN origin. As in the case of HC4N−, the eBE of this peaksuggests it is a CN stretch of the ã 1A′ state. In the HCCN−

and DCCN− spectra, this feature does not display the samestrong temperature dependence as the analogous peak in theHC4N− 150 K spectrum; however, the intensity of the peakdoes vary with time with respect to the ã 1A′ origin peak andthe X̃3A′′ progression. The photoelectron spectra shown inFig. 4 indicate the average intensity of the eBE ∼ 2.75 eVpeak over the data-sets collected.

In the previous anion photoelectron spectroscopy studyof HCCN− and DCCN−, Nimlos et al. also report the appear-ance of a peak at eBE ∼ 2.75 eV.5 In contrast to our findings,however, they observe this peak in the cold HCCN− spectrumbut not at room temperature. Furthermore, they do not reportthe appearance of this peak in either the cold or room tem-perature DCCN− spectra. The photoelectron spectrum of theisomer HCNC− is also shown in the 2002 work. There is noindication of a peak at eBE ∼ 2.75 eV due to the HCNC−

isomer, confirming that the eBE ∼ 2.75 eV peak is indeeddue to HCCN− (DCCN−). In the 2002 study, the peak is leftunassigned.

One important outcome of our reexamination of theHCCN− and DCCN− spectra is the identification of the eBE∼ 2.75 eV peak, which was previously left unassigned. Theharmonic normal mode analysis successfully predicts theã 1A′ ← X̃2A′′ photodetachment photoelectron spectra ofHCCN− and DCCN− (Fig. 4) to verify the identity of theeBE ∼ 2.75 eV peak. The simulation shows a peak at eBE= 2.77 eV resulting from excitation of CN stretch in theã 1A′ state of HCCN. From our photoelectron spectra, wemeasure the frequency of the ν2 CN stretch in the ã 1A′ stateof HCCN to be 1903(40) cm−1 (Table S2),26 and in DCCNto be 1936(40) cm−1 (Table S3).26 We resolve additionalfeatures in the DCCN− 150 K spectrum (Fig. 4(a)) between2.55 and 2.65 eV that are reproduced in the simulation, givingus further confidence in the Franck-Condon simulations ofthe ã 1A′ ← X̃2A′′ photoelectron spectra of these species.

The fact that this condition-dependent peak appears at asimilar energy relative to the ã 1A′ origin in both the HC4N−

and HCCN− (DCCN−) spectra (2395 and ∼1920 cm−1, re-spectively) and that there is no significant isotope shift upondeuteration of HCCN− confirms that these peaks result fromCN stretch excitation of the ã 1A′ state of the neutral.

Downloaded 31 Jan 2012 to 128.138.107.159. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

Page 9: Photoelectron spectroscopy of HC4N− · and dense molecular clouds; because nitrogen is a major component of interstellar dust, cyanopolyynes are also im-portant constituents of

204307-8 Vogelhuber et al. J. Chem. Phys. 135, 204307 (2011)

V. DISCUSSION

We observe several similarities between the HC4Nand HCCN (DCCN) systems, but it is useful to alsoexamine them in the context of a broader series ofmolecules.

First, we observe that the EAs of HCCN and HC4N aresimilar, but they are very different from the EA of HC3N.The EAs of HCCN and HC4N are both ∼2.0 eV, whereasthe EA of HC3N is considerably lower: ∼0.5 eV.43 This isperhaps not surprising, as trends in abundance, structure, andreactivity that depend on whether n is even or odd have beennoted among HCnN cyanopolyynes. This difference may bedue in part to the carbene nature of the even cyanopolyynesas compared to the closed-shell singlet odd cyanopolyynes.The neutral n = odd cyanopolyyne molecules are consideredto be more stable than the even ones; thus, the greater EAof the even cyanopolyynes implies that—unlike their neutralcounterparts—the odd cyanopolyyne anions are not more sta-ble than the even cyanopolyyne anions.

We also observe that exchanging an alkyne group fora nitrile increases the vertical detachment energy (VDE) by∼1 eV, which is clearly a result of the CN group shiftingthe spectrum to higher eBE. The VDE is the energy at whichthe greatest Franck-Condon overlap occurs, and it appears asthe maximum of the vibrational progression. The VDE of thetriplet ground state of HCCCH is 1.714 eV,44 of HCCN is2.36 eV, and of NCCCN (Ref. 20) is 3.53 eV. As a pseudo-halogen, the CN group acts to stabilize negative ions.20, 43, 45, 46

Because the CN group stabilizes the negative charge of theanion more than it stabilizes the corresponding neutral, theaddition of CN groups has the effect of increasing the VDE.

VI. CONCLUSIONS

We report the 364-nm photoelectron spectrum of HC4N.At lower binding energy, we observe a broad progression cor-responding to photodetachment from the X̃2A′′ HC4N− an-ion to the X̃3A′′ ground state of HC4N. The large geometrychange upon photodetachment—specifically, an increase of28.2◦ in the � C2–C3–C4—precludes the observation of anorigin peak. The adiabatic EA of HC4N is measured to be2.05(8) eV. At 2.809(4) eV, we observe the ã 1A′ origin ofHC4N. The �EST of HC4N is measured to be 0.76(8) eV, withan uncertainty primarily due to our inability to resolve the ori-gin of the triplet state.

For comparison, we also report the photoelectron spec-tra of HCCN− and DCCN−. The photoelectron spectra ofHCCN− and DCCN− also exhibit an extended progressioncorresponding to X̃3A′′ ← X̃2A′′ photodetachment. In thiscase, however, the origin transition can be resolved, giv-ing EA(HCCN) = 2.001(15) eV, which is very similar butmore accurate than the EA(HC4N). The ã 1A′ origin ofHCCN appears at 2.511(4) eV, yielding a �EST of HCCNof 0.510(15) eV. Like HC4N−, the photoelectron spectra ofHCCN− and DCCN− reveal peaks ∼2000 cm−1 higher in en-ergy than the singlet origin that vary in intensity dependingon experimental conditions. In each case, these peaks resultfrom CN stretch excitation of the ã 1A′ state of the neutral.

The HC4N and HCCN (DCCN) systems display sev-eral similarities, but through comparison with other organicchains, we gain insight into how they relate to a broader se-ries of molecules. While the EA of HC4N is essentially thesame as that of HCCN, the EA of HC3N is much lower. Inaddition, exchanging an alkyne group for a nitrile raises theVDE by ∼1 eV.

ACKNOWLEDGMENTS

We are pleased to acknowledge generous support for thiswork from the National Science Foundation (NSF) (GrantNo. CHE-0809391) and the Air Force Office of ScientificResearch (AFOSR) (FA9550-09-1-0046). A.B.M. is pleasedto acknowledge support from NSF Grant No. CHE-0848242.R.J.M. gratefully acknowledges support from NSF Grant No.CHE-1011959. K.M.V. and S.W.W. would like to extendthanks to Eric Green for his help in collecting data and toJeff Gazaille for NMR assistance.

1W. C. Lineberger and W. T. Borden, Phys. Chem. Chem. Phys. 13, 11792(2011).

2C. Wentrup, Reactive Molecules (Wiley, New York, 1984).3R. L. Schwartz, G. E. Davico, T. M. Ramond, and W. C. Lineberger, J.Phys. Chem. A 103, 8213 (1999).

4S. W. Wren, K. M. Vogelhuber, K. M. Ervin, and W. C. Lineberger, Phys.Chem. Chem. Phys. 11, 4745 (2009).

5M. R. Nimlos, G. Davico, C. M. Geise, P. G. Wenthold, W. C. Lineberger,S. J. Blanksby, C. M. Hadad, G. A. Petersson, and G. B. Ellison, J. Chem.Phys. 117, 4323 (2002).

6J. Y. Qi, M. D. Chen, W. Wu, Q. E. Zhang, and C. T. Au, Chem. Phys. 364,31 (2009).

7M. B. Bell, P. A. Feldman, M. J. Travers, M. C. McCarthy, C. A. Gottlieb,and P. Thaddeus, Astrophys. J. 483, L61 (1997).

8M. C. McCarthy, J. U. Grabow, M. J. Travers, W. Chen, C. A. Gottlieb, andP. Thaddeus, Astrophys. J. 494, L231 (1998).

9R. Nagarajan and J. P. Maier, Int. Rev. Phys. Chem. 29, 521 (2010).10M. C. McCarthy and P. Thaddeus, J. Mol. Spectrosc. 232, 351 (2005).11C. D. Ball, M. C. McCarthy, and P. Thaddeus, J. Chem. Phys. 112, 10149

(2000).12K. Aoki, S. Ikuta, and A. Murakami, Chem. Phys. Lett. 209, 211 (1993).13M. C. McCarthy, A. J. Apponi, V. D. Gordon, C. A. Gottlieb, P. Thaddeus,

T. D. Crawford, and J. F. Stanton, J. Chem. Phys. 111, 6750 (1999).14M. C. McCarthy, J. U. Grabow, M. J. Travers, W. Chen, C. A. Gottlieb, and

P. Thaddeus, Astrophys. J. 513, 305 (1999).15K. Aoki and S. Ikuta, J. Chem. Phys. 98, 7661 (1993).16S. Ikuta, T. Tsuboi, and K. Aoki, J. Mol. Struct. THEOCHEM 528, 297

(2000).17J. Kalcher, Chem. Phys. Lett. 403, 146 (2005).18M. Z. Kassaee, J. Azarnia, and S. Arshadi, J. Mol. Struct. THEOCHEM

686, 115 (2004).19N. P. Bowling, R. J. Halter, J. A. Hodges, R. A. Seburg, P. S. Thomas,

C. S. Simmons, J. F. Stanton, and R. J. McMahon, J. Am. Chem. Soc. 128,3291 (2006).

20D. J. Goebbert, K. Pichugin, D. Khuseynov, P. G. Wenthold, and A. Sanov,J. Chem. Phys. 132, 224301 (2010).

21J. A. Tang, Y. Sumiyoshi, and Y. Endo, Chem. Phys. Lett. 315, 69 (1999).22J. Cernicharo, M. Guelin, and J. R. Pardo, Astrophys. J. 615, L145 (2004).23K. M. Ervin and W. C. Lineberger, in Advances in Gas Phase Ion Chem-

istry, edited by N. G. Adams and L. M. Babcock (JAI, Greenwich, 1992),Vol. 1, p. 121.

24D. G. Leopold, K. K. Murray, A. E. S. Miller, and W. C. Lineberger, J.Chem. Phys. 83, 4849 (1985).

25K. M. Ervin, J. Ho, and W. C. Lineberger, J. Chem. Phys. 91, 5974(1989).

26See supplementary material at http://dx.doi.org/10.1063/1.3663617 for de-tails of the synthesis of H3C4N; the photoelectron spectrum of H2C4N−(Fig. S1); the polarization dependence of the photoelectron spectrum ofHC4N− (Fig. S2); the full simulated photoelectron spectrum of HC4N−

Downloaded 31 Jan 2012 to 128.138.107.159. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

Page 10: Photoelectron spectroscopy of HC4N− · and dense molecular clouds; because nitrogen is a major component of interstellar dust, cyanopolyynes are also im-portant constituents of

204307-9 Photoelectron spectroscopy of HC4N− J. Chem. Phys. 135, 204307 (2011)

(Fig. S3); one-dimensional potential cuts along the H–C–CN bend co-ordinate of HCCN (Fig. S4); the full simulated photoelectron spectra ofDCCN− and HCCN− (Fig. S5); and for tables of the calculated equilib-rium geometries of HCCN− (DCCN−) and HCCN (DCCN) (Table S1) andfrequencies of HCCN (Table S2) and DCCN (Table S3).

27S. W. Wren, K. M. Vogelhuber, J. Garver, S. Kato, L. Sheps, V. Bierbaum,and W. C. Lineberger, “Nitrogen proximity effects in azines: How nitrogenatoms affect C–H bond strengths and anion stability,” J. Am. Chem. Soc.(submitted).

28C. S. Feigerle, Ph.D. dissertation, University of Colorado, 1983.29D. M. Neumark, K. R. Lykke, T. Andersen, and W. C. Lineberger, Phys.

Rev. A 32, 1890 (1985).30K. M. Ervin, W. Anusiewicz, P. Skurski, J. Simons, and W. C. Lineberger,

J. Phys. Chem. A 107, 8521 (2003).31K. M. Ervin, PESCAL, Fortan program, 2010.32J. Cooper and R. N. Zare, J. Chem. Phys. 48, 942 (1968).33M. J. Frisch, G. W. Trucks, H. B. Schlegel et al., GAUSSIAN 03, Revision

B.05, Gaussian, Inc., Pittsburgh, PA, 2003.34C. Moller and M. S. Plesset, Phys. Rev. 46, 618 (1934).35R. Krishnan, J. S. Binkley, R. Seeger, and J. A. Pople, J. Chem. Phys. 72,

650 (1980).

36K. M. Ervin, T. M. Ramond, G. E. Davico, R. L. Schwartz, S. M. Casey,and W. C. Lineberger, J. Phys. Chem. A 105, 10822 (2001).

37J. T. Hougen, P. R. Bunker, and J. W. C. Johns, J. Mol. Spec. 34, 136 (1970).38K. M. Vogelhuber, S. W. Wren, A. B. McCoy, K. M. Ervin, and

W. C. Lineberger, J. Chem. Phys. 134, 184306 (2011).39J. R. Reimers, J. Chem. Phys. 115, 9103 (2001).40R. Borrelli and A. Peluso, J. Chem. Phys. 125, 194308 (2006).41T. Ichino, S. M. Villano, A. J. Gianola, D. J. Goebbert, L. Velarde,

A. Sanov, S. J. Blanksby, X. Zhou, D. A. Hrovat, W. T. Borden, and W. C.Lineberger, J. Phys. Chem. A 115, 1634 (2011).

42E. T. Seidl and H. F. Schaefer, J. Chem. Phys. 96, 4449 (1992).43D. J. Goebbert, D. Khuseynov and A. Sanov, J. Phys. Chem. A 114, 2259

(2010).44D. L. Osborn, K. M. Vogelhuber, S. W. Wren, E. M. Miller,

Y.-J. Lu, A. S. Case, R. J. McMahon, and W. C. Lineberger,“Photoelectron spectroscopy of the propargylene anion”(unpublished).

45D. J. Goebbert, D. Khuseynov, and A. Sanov, J. Chem. Phys. 131, 161102(2009).

46D. J. Goebbert, L. Velarde, D. Khuseynov, and A. Sanov, J. Phys. Chem.Lett. 1, 792 (2010).

Downloaded 31 Jan 2012 to 128.138.107.159. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions