Skip to main content
Nature Communications logoLink to Nature Communications
. 2019 Nov 15;10:5199. doi: 10.1038/s41467-019-13039-y

The dicarbon bonding puzzle viewed with photoelectron imaging

B A Laws 1,, S T Gibson 1, B R Lewis 1, R W Field 2
PMCID: PMC6858380  PMID: 31729361

Abstract

Bonding in the ground state of C2 is still a matter of controversy, as reasonable arguments may be made for a dicarbon bond order of 2, 3, or 4. Here we report on photoelectron spectra of the C2 anion, measured at a range of wavelengths using a high-resolution photoelectron imaging spectrometer, which reveal both the ground X1Σg+ and first-excited a3Πu electronic states. These measurements yield electron angular anisotropies that identify the character of two orbitals: the diffuse detachment orbital of the anion and the highest occupied molecular orbital of the neutral. This work indicates that electron detachment occurs from predominantly s-like (3σg) and p-like (1πu) orbitals, respectively, which is inconsistent with the predictions required for the high bond-order models of strongly sp-mixed orbitals. This result suggests that the dominant contribution to the dicarbon bonding involves a double-bonded configuration, with 2π bonds and no accompanying σ bond.

Subject terms: Chemical physics, Atomic and molecular physics


In spite of its apparent simplicity, the dicarbon molecule has a bonding structure which is matter of debate. Here the authors measure high-resolution spectra of the C2 anion by photoelectron imaging, revealing a bonding configuration dominated by a double π bond, with no accompanying σ bond.

Introduction

Despite the relative simplicity of homonuclear diatomic molecules, the bonding nature of dicarbon, C2, has long been a topic of debate. This discussion has been driven by recent advances in computational methods, with various studies suggesting the carbon–carbon bond may have a bond order of 2, 3, or even 4, with the latter from ab-initio studies1. Standard qualitative theories predict different values for the dicarbon bond order2,3. From a simple Lewis electron-pair repulsion description, the 8 valence electrons in C2 are predicted to form a quadruple bond. However, this bonding assignment seems unlikely, with stable quadruple bonds typically only found between transition metal elements that have partially filled d-orbitals4. In contrast, a hybrid-orbital (HO) style approach invokes sp1 hybridisation to predict a triple bond between the carbons. However, if molecular orbital (MO) theory is used, a ground-state valence electron configuration for C2 of KK(σ2s)2(σ2s*)2(π2p)4 is predicted, yielding a bond order of 2, with the unusual situation of a π double bond with no accompanying σ bond.

From an ab-initio approach, standard Hartree–Fock-based calculations support a dicarbon double bond; however, more advanced theoretical studies have suggested that the C–C bond may be better described by a higher bond order58. A recent high-level full configuration–interaction calculation, combined with valence bond (VB) theory, identified four distinct contributions to the bonding in C21. This included contributions from a σ bond, 2×π bonds, plus an interaction between the outward pointing sp1 hybrid orbitals1. The strength of this inverted bond between the sp1 orbitals has since been calculated at various levels of theory and is estimated to contribute ~50–80 k mol−11,5,9,10. This high bond-order model is further supported by a subsequent quantum chemistry calculation, which found higher magnetic shielding in C2 compared with C2H2, supporting a bulkier C–C bond11. However, not all of the recent studies are in agreement, with some research preferring the notion of a double12, triple3, or quasi double–triple13 bond, whereas other studies note that the theoretical approaches are not sufficient to definitively discern between the different bonding models1416.

Despite advances in spectroscopic techniques, experimental studies have not been able to confirm any of the suggested bonding structures, with the majority of the debate currently driven by the results of ab-initio calculations2. Due to the highly reactive nature of C2, most experimental studies involve flame-emission1720 or plasma-discharge21,22 spectroscopy. These studies support a bond order between 2 and 3, with a measured C–C bond length of 1.243 Å, longer then a typical alkyne triple bond, but shorter then a typical alkene double bond19. Likewise, a measured C–C bond dissociation energy of 602 kJ mol−1, as well as a calculated bond restoring force of 12 N14, also lie between double and triple bond limits23.

As the dicarbon anion C2 is stable, photoelectron spectroscopy may be used to probe the reactive C2 neutral molecule24,25. The first dicarbon photoelectron experiment was performed by Ervin et al.24, on C2 anions produced in an O/HCCH afterglow ion source. This source produced hot anions, with multiple hot bands present in the spectrum, and defined an accurate value for the electron affinity of C2 of 3.269(6) eV24. A later study by Bragg et al.26 probed the low-lying excited states of C2, in a single wavelength measurement at 264 nm (as part of a larger study, employing time-resolved photoelectron spectroscopy to examine transitions from excited anion states). This study provided the first experimental anisotropy measurement of photodetachment from the C2 anion.

In this work, the photoelectron spectrum of C2 is revisited using a high-resolution photoelectron imaging (HR-PEI) spectrometer. Although it is now well established that the photoelectron angular distributions indicate the character of the relevant molecular anion orbitals27, electric dipole selection rules also influence the electron anisotropy via the character of the final neutral states, providing insight to understand the nature of the bonding between carbon atoms in C2. Here we report on photoelectron angular distributions measured at a range (290 to 355 nm) of wavelengths, to probe the character of the C2 orbitals. This reveals detachment occurs from pure s-like (3σg) and p-like (1πu) orbitals, suggesting that the dominant contribution to the dicarbon bonding involves a double-bonded configuration, with 2π bonds and no accompanying σ bond.

Results and Discussion

Theoretical discrepancies

The simple idea of a bond order, defined as

BondingelectronsAntibondingelectrons2 1

is often a useful way for chemists to describe the bonding nature in molecules. However, this can be an oversimplification in systems where the contribution from different bonding/anti-bonding electrons are unequal. Furthermore, the concept breaks down entirely in systems, which are intrinsically multi-reference in nature. Thus, when discussing the bonding in a molecule as subtle as C2, it is important to keep in mind that there are multiple configurations all contributing to the overall state.

Ab-initio descriptions of the bonding are also complicated by the multi-reference nature of C2. Specifically, it is the quasi-degeneracy of the 2σu*, 1πu and 3σg molecular orbitals, responsible for the numerous low-lying excited electronic states of C2, which muddle the bonding picture. Much of this uncertainty is influenced by the nature of the 2σu* orbital, which is predicted to be a very weakly anti-bonding orbital1,28. To account for this behaviour, an alternative orbital scheme has been proposed, involving hybridisation of the 2σu*3σg molecular orbitals, forming sp1-like singly occupied hybrid orbitals (2σu*+3σg) and (2σu*3σg)5, as depicted in Fig. 1.

Fig. 1.

Fig. 1

Orbital descriptions of C2 using different approaches, highlighting the uncertain bond nature of dicarbon. a Hybrid-orbital picture of C2, involving sp1 hybridisation on the carbon atoms. This approach suggests a triple-bonded dicarbon molecule, with one σ and 2π bonds, as represented by solid black lines. b Molecular orbital diagram for C2, predicting a double bond between the C atoms consisting of 2π bonds, without an accompanying σ bond. The additional electron for the anion C2 is shown in green. c Orbital diagram for C2 from a valence-bond viewpoint. Due to the quasi-degenerate nature of orbitals 2σu,3σg and 1πu, the 2σu* and 3σg orbitals mix to form sp1-like hybrid orbitals on each of the carbons. Valence bond theory suggests a fourth bond involving the interaction between the outward pointing φ orbitals

To visualise the differences between these possible bonding pictures, ab-initio calculations were performed in this work using NWChem29 and Q-Chem30 software. Most standard methods for approximately solving the Schrödinger equation for molecules are based on Hartree–Fock [or self-consistent field (SCF)] theory, utilising a mean-field approximation. However, for molecules with quasi-degenerate or low-lying excited states, a multiconfigurational complete active space (CASSCF) approach may be required. This approach accounts for states that are a linear combination of several quasi-degenerate configurations, allowing for non-integer orbital occupation numbers.

The molecular orbitals for C2 were calculated with both SCF and CASSCF methods, using a cc-pVTZ Dunning basis set31, as shown in Fig. 2. The key difference between the two approaches is the ordering of the 2σu* orbital, with the multiconfigurational calculation increasing the orbital energy due to possible mixing between the 2σu*3σu* and 2σg3σg orbitals. This mixing is also illustrated in the occupation numbers, with a substantial (~0.4) occupation found in the 3σg orbital. As the dominant weight in this mixed orbital lies with the anti-bonding 2σu*, these occupation numbers suggest a bond order of three, consistent with a pure hybrid-orbital theory argument6. However, if this 2σu*3σu* and 2σg3σg mixing is allowed to occur before the calculation, using hybrid orbitals φL=2σu+3σg and φR=2σu3σg, the predominant configuration is for each orbital to be singly occupied, with one electron on each of the carbons5. From Hund’s Rule, one may expect the lowest energy configuration to be the triplet state, with the odd electrons occupying the degenerate hybrid orbitals in the same spin state. However, generalized valence bond calculations, using the principal of maximum coupling between overlapping atomic orbitals, have shown that the energy gain from the coupling interaction between the singlet-paired electrons in each hybrid orbital is greater than the energy cost from violating Hund’s Rule1,32. This is consistent with the VB picture (Fig. 1c), used to suggest a dicarbon quadruple bond, whereas the SCF results are consistent with the double-bonded MO picture (Fig. 1b). Orbitals of the C2 anion were also calculated using the same (cc-pVTZ) basis. The addition of an electron has a minimal impact on the structure of the inner orbitals, with a calculated overlap integral of the πu anion and neutral orbitals giving 85% similarity, with the 15% difference accounted for by the slight change in the C2 bond length.

Fig. 2.

Fig. 2

Valence molecular orbitals of C2 calculated using NWChem software29. a Orbitals calculated using an SCF (self-consistent field) approach, recreating the molecular orbital diagram in Fig. 1b. b Orbitals calculated using a multiconfigurational CASSCF (complete active space self-consistent field) approach, which agree with the valance bond orbital diagram (Fig. 1c). Orbital occupation numbers are also given

High-resolution photoelectron imaging

To investigate the uncertainty regarding the bonding nature of dicarbon, photoelectron spectra of C2 were measured, using the HR-PEI spectrometer at the Australian National University. Dicarbon anions are produced in a single pulsed-jet discharge source of pure ethylene, with the subsequent ion mass separated by time-of-flight. Ions are then photodetached at 355 nm using the third harmonic of a Nd:YAG laser, with the electrons mapped onto a detector using velocity-map imaging (VMI). This allows for energetic and angular information to be measured simultaneously. A raw velocity-mapped image of C2 photodetachment, corresponding to 414,511 electrons, is presented in Fig. 3. The imagecontains two rings, corresponding to the C2X1Σg+ + e C2X2Σg++hν and C2a3Πu + e C2X2Σg++hν photodetachment transitions. Because of the quasi-degeneracy of the 2σu*, 1πu and 3σg molecular orbitals, the dicarbon neutral molecule has many low-lying electronic states, which leads to the unusual property whereby the term energy of the first excited state [718.318(2) cm−133] is smaller than the vibrational frequency [ωe=1854.5(8)  cm−124]. Consequently, at a detachment wavelength of 355 nm we observe two electronic transitions, but only the vibrational original of each state.

Fig. 3.

Fig. 3

Velocity-map image of C2 at 355 nm. Fast electrons are mapped to the detector edge with slow electrons at the centre, whereas the laser polarisation defines the vertical axis. Two transitions are observed, with detachment to the ground X1Σg+ state observed as an outer ring and detachment to the first excited a3Πu state represented by the inner ring. The two transitions display opposite anisotropies, with the 1Σg+ electrons preferentially distributed at the poles of the image and the a3Πu electrons weakly focused around the horizontal axis

Rotational band models for the C2(X1Σg+) C2(X2Σg+) and C2(a3Πu) C2(X2Σg+) electronic transitions have been constructed using methods similar to those developed by Buckingham et al.34. A detailed derivation of the molecular rotational model may be found in the Supplementary Discussion. The Hunds case-(b) to case-(b) and case-(b) to case-(a) rotational models were fitted the the ground (X1Σg+) and excited (a3Πu) photoelectron transitions, respectively, as shown in Fig. 4. The model fit yields an anion rotational temperature of T = 197(2) K, whereas the Gaussian full-width half-maximum of each transition gives an energy resolution of ΓX1Σ=15(1) cm−1 and Γa3Π=18(1) cm−1, which are consistent with expectations for the spectrometer at electron kinetic energies of ϵ~0.3 eV.

Fig. 4.

Fig. 4

Photoelectron spectrum of C2 at  355 nm. Two electronic transitions C2(X1Σg+) C2(X2Σg+) and C2(a3Πu) C2(X2Σg+) are observed, separated by 612 cm−1. The experimental spectrum is shown in black, with a rotational band model (XX shown in blue, aX shown in green) fitted to each transition. This highlights the presence of a hot band (1, 1)

The rotational model also shows additional weak transitions to the right of the XX peak and to the left of the aX peak. These signatures are assigned to the (1,1) hot-band transition. In the earlier measurement of Ervin et al.24, strong (0,1) hot-band transitions were observed; however, only the (1,1) transitions appear to be present in the photoelectron spectrum presented here, possibly due to a lower source temperature. To validate the assignment of the (1,1) bands in Fig. 4, Franck–Condon factors were calculated. Potential energy curves of the anion X2Σg+ and neutral X1Σg+ and a3Πu states were constructed using the Rydberg–Klein–Rees inversion method, with the resulting vibrational wavefunction overlap integrals presented in Table 1.

Table 1.

Franck–Condon factors for the vibrational transitions in C2 photodetachment

X1Σg+X2Σg+ a3ΠuX2Σg+
(0, 1) 0.090 0.231
(0, 0) 0.902 0.756
(1, 0) 0.096 0.209
(1, 1) 0.726 0.365
(2, 1) 0.177 0.287
(2, 2) 0.198 0.089

Table 1 confirms that photodetachment from vibrationally hot anions yielding C2(X1Σg+) will have a maximum intensity at (1,1), explaining the absence of the (0,1) hot band in the spectrum in Fig. 4. Photodetachment yielding C2(X3Πu) also has a maximum intensity at (1,1): however, this transition has a much smaller Franck–Condon factor than the corresponding transition in the ground state. Consequently, this transition has less intensity in the experimental spectrum. From the relative intensities of the (0,0) and (1,1) transitions in the photoelectron spectrum, the vibrational temperature of the C2 anion may be defined. This analysis reveals a temperature of Tvib~900K, which is significantly higher than the rotational temperature derived from the model fit in Fig. 4 [Trot=197(2)K]. This is also higher than the normal operating conditions of the spectrometer, suggesting that the high-voltage discharge source preferentially produces dicarbon anions in vibrationally excited states.

The photoelectron spectrum in Fig. 4 may be calibrated using the term energy [Te=718.318(1) cm−1] of the excited a3Πu state, which is well-defined from neutral flame-emission spectroscopy studies35. The rotational band model provides an accurate position for the band origin, allowing for the precise determination of molecular constants (listed in the Supplementary Table 1). From the calibration of the high-resolution photoelectron spectrum presented here, a precise value for the electron affinity of EA = 3.2727(4) eV is determined, in agreement with the previously accepted value of Lineberger et al.24 of EA = 3.269(6) eV. The anion ground-state rotational [B=1.746(1)cm1] and vibrational [ωe=1782(2)cm1] constants are also extracted.

To investigate the relationship between the photoelectron spectrum of C2 and the laser detachment energy, additional measurements were recorded by pumping a Sunlite Optical Parametric Oscillator (OPO) with the third harmonic of a Nd:Yag laser. To achieve photon energies greater than the electron affinity [3.2727(4) eV], the OPO output was doubled using a Continuum FX-Doubler. This allowed for photoelectron spectra of C2 to be measured at a range of wavelengths (290–325 nm), as shown in Fig. 5. The OPO measurements show both the origin (0, 0) and first excited (1, 0) vibrational transitions for the ground X1Σg+ and excited a3Πu electronic states.

Fig. 5.

Fig. 5

Photoelectron spectra of C2 at a range (290–355 nm) of detachment wavelengths. The vibrational origin transitions are observed for the ground X1Σg+ and first excited a3Πu electronic states. The OPO measurements (290–325 nm) also show the (1, 0) vibrational transition for each state

Bonding insights from the anisotropy

The angular distribution of the photoelectrons, as measured by the VMI lens, is typically not isotropic and is related to the character of the parent anion molecular orbital. For detachment using linearly polarised light, the differential cross-section is given by

dσdΩ=σtotal4π1+βP2(cosθ), 2

where θ is the angle between the ejected electron and the (vertical) laser polarisation, and P2 is the second-order Legendre polynomial. The anisotropy parameter (β) may take values ranging from −1 to 2, for purely perpendicular and parallel electronic transitions, respectively34. β is a quantitative measure of how anisotropic the electron distribution is, with β=0 corresponding to a perfectly isotropic distribution.

A qualitative description of the anisotropy parameter may be determined by visual inspection of the velocity-map image in Fig. 3. Noticeably, the two electronic transitions have different photoelectron angular distributions, with the electrons from detachment to X1Σg+ preferentially distributed at the poles of the image signalling a strong positive anisotropy parameter, whereas the a3Πu electrons appear to have a slight negative anisotropy parameter, with the distribution skewed towards the horizontal. Quantitative values are obtained by fitting Eq. (2) to radially integrated transition intensities, as a function of angle. Applying this to the inverse Abel transformed velocity-mapped image of Fig. 3 gives anisotropy parameters β(X1Σg+) =+1.75(5) and β(a3Πu) =0.35(1).

The measured anisotropy parameters are directly related to the interference of detachment partial waves by the well-known Cooper–Zare anisotropy formula, applicable for a central potential,

β=(1)χ,12+(+1)(+2)χ,+126(+1)χ,+1χ,1cos(δ+1δ1)(2+1)[χ,12+(+1)χ,+12], 3

where χ,±1 is the radial transition dipole matrix element for the ±1 partial wave emitted from the initial atomic orbital and δ±1 are the associated phase shifts36. Qualitatively, Eq. (3) describes the variation of the anisotropy of an electron distribution, with kinetic energy ϵ, for detachment from an orbital with angular momentum . Therefore, a measurement of the anisotropy of C2 detachment is sensitive to the character of the relevant anion molecular orbitals.

For detachment from an s-orbital, =0, Eq. (3) becomes

βs=2χ0,12χ0,12=2. 4

A strong positive anisotropy close to +2, such as β(X1Σg+)=+1.75(5), is indicative of a dominant s orbital character. Conversely, for a p orbital with =1, Eq. (3) becomes

βp=2A2ϵ24Aϵcos(δ1δ0)1+2A2ϵ2 5

where

Aϵ=χ1,2χ1,0 6

and where the Hanstorp coefficient A is used to represent the ratio of radial matrix elements37. Equation (6) shows that as ϵ0, β0. A negative β parameter close to threshold, such as β(a3Πu) =0.35(1), is typical of detachment from a p-like orbital, as seen in similar diatomic molecules BN38 and NO27,39 where an electron is detached from a π-orbital.

Assuming that the relevant molecular orbitals in C2 may be accurately described by a mixture of s and p character, photodetachment will be governed by the modified Cooper–Zare equation40,

βsp=2Zϵ+2Aϵ24ϵcos(δ2δ0)1A+2Aϵ2+Zϵ, 7

where Z=(1f)Bd(fAd), for an orbital 1fs+fp. Parameters Ad and Bd from Eq. (7) represent the scaling of the different radial dipole integrals, with the ratio BdAd=83 for 2s/2p mixing, whereas cos(δ2δ0), the phase shift between the outgoing waves, is 140. This leaves the parameter f, associated with the fractional percentage of s and p character (f=0 for pure s, f=1 for pure p), and the Hanstorp coefficient A in Eq. (7), as the only fitting parameters.

The experimental anisotropy parameters for C2 photodetachment at 355–295 nm (to both the ground and first-excited neutral states) are plotted in Fig. 6, together with the values of Bragg et al.26 measured at 264 nm. By fitting the anisotropies for each wavelength to the sp character model Eq. (7), values for the Hanstorp coefficients, A, and fractional character percentage of the detachment orbital, f, may be determined. This fitting process gives values of A=1.88(8)eV1 and f=0.081(9) for C2X1Σg+ C2X2Σg+ detachment and values of A=0.86(6)eV1 and f=0.87(1) for the excited a3Πu X2Σg+ transition. This corresponds to a ground-state detachment orbital with ~92%s character, whereas the excited state corresponds to detachment from an orbital with ~87%p character.

Fig. 6.

Fig. 6

Anisotropy parameters for C2 photodetachment. Experimental values are show for detachment at 295–355 nm [from this work (circles/squares)] and at 264 nm [from ref. 26 (stars/crosses)]. Detachment to the ground X1Σg+ state has a Hanstorp coefficient of A=1.88(8)eV1 with an orbital character of f=92%s (orange curve), whereas detachment to the excited a3Πu state has a Hanstorp coefficient of A=0.86(6)eV1 and an orbital character of f=13%s (blue curve). Anisotropy curves calculated from Dyson orbitals are also shown (red and blue dashed lines). Error bars represent 1 SD, defined by the covariance matrix of the least-squares fit to Eq. (2)

Modified Cooper–Zare curves (Eq. (7)) are shown in Fig. 6 for different values of sp orbital character percentages, with f=0.081,0.15,0.8, and 0.87. This graph highlights a general rule of sp mixing, with a higher percentage of s character associated with a β close to +2, whereas orbitals with more p character produce β parameters close to 1. From comparison to the experimental data points, it can be seen that f(X1Σg+)<0.15 and f(a3Πu)>0.8, giving lower limits on the amount of sp character of each detachment orbital.

Another useful comparison is to construct Dyson orbitals ϕd, which are defined as the overlap between the initial ϕi(n) and final ϕf(n1) states of the molecule,

ϕd=Nϕi(n)(1,,n)ϕf(n1)(2,,n)dn, 8

so that the photoelectron dipole moment Dk may be written in terms of the Dyson orbital ϕd and the outgoing electron plane wave ψk41,

Dkϕderψk. 9

As Dyson orbitals may be constructed using quantum chemistry software, this will provide a direct link between the ab-initio theory and the experimental results. Dyson orbitals corresponding to C2 photodetachment to both the ground and first excited state were calculated using Q-Chem software30 at the CCSD(T) level of theory with a 6-311 + G(4+)(3df) basis set. Anisotropy parameters were then calculated over a range of electron kinetic energies using ezDyson software42. As anisotropy calculations are sensitive to the asymptotic tails of the Dyson orbitals, it has been shown in previous studies that the contribution from the diffuse Gaussian orbitals may need to be adjusted to get better agreement between experiment and theory43. In this case it was found that increasing the contribution from the diffuse orbitals by a factor of two was required. The resulting anisotropy curves for detachment from the 3σg and 1πu orbitals are also shown in Fig. 6.

Each of the bonding schemes, double, triple and quadruple, may be compared with the above limits to deduce which one best aligns with the orbital information gained from the present experimental measurements. The double-bonded structure suggested by the MO diagram in Fig. 1b predicts that the C2X1Σg+ +e C2X2Σg++hν transition involves detaching an electron out of the 3σg bonding orbital and the transition to form the lowest triplet state (a3Πu) would involve detachment from the 1πu bonding orbital. The 3σg orbital possesses predominant s-character and the 1πu orbital possesses significant p-character (Fig. 2), as confirmed by the Dyson orbital calculations. Therefore, this orbital scheme is consistent with the angular distributions measured in Fig. 3, satisfying the orbital character requirements of at least 85% s and 80% p character, for the singlet and triplet transitions, respectively.

The triple-bonded structure, predicted by HO theory (Fig. 1a), suggests that C2 possesses σ bond, 2π bonds, and singly occupied sp1 orbitals on each of the carbons. However, as the sp1 hybrid orbitals will have close to 50% s and 50% p character, they are not suitable for either the singlet or triplet detachment. Multiconfigurational CASSCF calculations have also been used to predict a C2 bond order of 3, with a calculated orbital configuration of KK(2σg)2(1πu)2(1πu)2(2σu*)1.6(3σg)0.4 (Fig. 2). Detachment to the X1Σg+ ground state would most likely occur out of what is primarily a 3σg orbital. However, as this orbital is mixed with both the 2σu* and higher lying 1πg* orbitals, achieving an s character purity of over 85% would seem unlikely. Furthermore, the triplet transition may involve detachment from what is primarily a 2σu* orbital, which, especially when orbital mixing with 3σg is accounted for, is unlikely to satisfy the 80% p character requirement.

Likewise, the suggestion of a quadruple bond faces a similar problem. In this scenario, it is suggested there are one σ and two π bonds, as before, as well as a weak bond-like interaction between the singly occupied sp1 hyrbid orbitals on the carbons. However, even if the highest occupied molecular orbitals/lowest unoccupied molecular orbitals are not pure 50:50sp1 orbitals, it is very unlikely that they would represent over 85% s purity, as is required for the C2(X1Σg+) C2(X2Σg+) transition.

It is important to note that, as an intrinsically multi-reference system, it is not possible to exclusively assign the structure of C2 to any one bonding configuration. However, the anisotropy results presented here strongly suggest that the dominant contribution arises from a double bond-like configuration, with the unusual case of having 2π bonds without an accompanying σ bond.

Ab-initio calculations give a complete multi-configuration picture, with the possibility of many different configurations contributing to the overall state. In this work the experimental photoelectron angular distribution decomposes the detachment orbital as a linear combination of atomic-like orbitals (s, p, d). The measurements indicate almost pure atomic character, which supports a molecular orbital configuration for the C2(X2Σg+) anion of KK(2σg)2(2σu*)2(1πu)4(3σg)1. The anisotropy parameters reported here are consistent with a double-bond configuration for the neutral C2 molecule, with 2π bonds and no accompanying σ bond. This suggests that although multiple configurations may contribute to the overall state, the 2π bond configuration is likely the dominant contributor to the overall bonding structure. This would also appear to be in agreement with the other experimental observations that have been made, namely the measured C–C bond length of 1.243 Å19, which lies between the usual lengths for double and triple bonds. The fact that this bond differs from a standard C=C double bond, may be due to the unique nature of having two π bonds, as opposed to the usual σ+π bond configuration. This work suggests that triple and quadruple bond configurations, based on the hybridisation of sp orbitals, will only have a small influence on the overall bonding nature, as opposed to what some calculations and theoretical approaches have suggested recently.

Methods

High-resolution photoelectron imaging spectrometer

A velocity-map imaging lens, a modified version of the original concept of Eppink and Parker44, is incorporated co-axially into a fast-beam spectrometer. Details of the apparatus are given in refs. 39,45. Dicarbon anions C2 are produced by passing pure ethylene gas (C2H4) through a pulsed-valve nozzle at a stagnation pressure of ~2 atm, with supersonic expansion through a pulsed high-voltage discharge. Negative ions are extracted, accelerated to 500 eV and focused into an ion gating, bunching and potential re-referencing unit46. Anion mass separate travelling over a 2 m time-of-flight tube, with each mass packet bunching to an ~2 mm3 volume at the interaction region, where an electrostatic gate selects the mass packet of interest. This ion packet is crossed with a detachment laser beam, generated from a Continuum Powerlite 9010 Nd:YAG laser, operated at its third harmonic, 355 nm. The laser produces between 20 and 50 mJ per pulse at 10 Hz, measured near the interaction region. A Continuum Sunlite EX OPO combined with an FX-Doubler was used to produce wavelengths at 295, 300, 305, 315 and 325 nm, with laser powers of 2–10 mJ. The precise wavelength of the laser light is measured using a wavemeter (HighFinesse WS7 UV), giving calibrated wavelengths of 289.612(1), 299.593(1), 304.596(1), 314.581(1), 324.548(1) and 354.8071(9) nm.

Photodetached electrons are velocity mapped to a 75 mm diameter MCP/phosphor screen detector (Burle), which is imaged by a 2048×2048 pixel monochrome CCD camera (PCO 2000). Each camera frame is transferred to a computer at a 10 Hz repetition rate and is processed in real time to identify individual electron events, centroiding position to a sub-pixel accuracy. The electron positions are written to a data file for subsequent analysis.

Image analysis

An image of the velocity-mapped photodetached electrons at the detector is then obtained through binning of the centroided electron-event positions into a rectangular pixel-grid image, of arbitrary pixel number, based on signal-to-noise ratio. The velocity-map image is centred and then circularised by an angular-dependent radial scaling determined by comparing adjacent radial slice intensity profiles47. An inverse Abel transformation of the VMI, based on the algorithm of Hansen and Law48,49, returns a slice image of the 3D electron source distribution.

Absolute energy calibration of the photoelectron spectra is achieved using published measurements of species, including O45 and O250, which have been studied under similar conditions as used for the C2 measurements.

Supplementary information

Supplementary Information (163.1KB, pdf)
Peer Review File (250.8KB, pdf)

Acknowledgements

This research was supported by the Australia Research Council Discovery Project Grant DP160102585.

Author contributions

B.A.L. performed the experiments and ab-initio calculations. B.A.L. and S.T.G. were involved in the data analysis, the experiment design, and wrote the computer codes used to examine the experimental data. B.A.L., S.T.G., B.R.L. and R.W.F. interpreted the results in relation to bonding in C2. B.A.L. wrote the manuscript, with input, comments and discussion from S.T.G., B.R.L. and R.W.F.

Data availability

All data supporting the findings of this study are available within the Article and from the corresponding author upon reasonable request. Experimental data are available at https://physics.anu.edu.au/research/portal/vmi.

Code availability

All of the computer code supporting the findings of this study are available from the corresponding author upon reasonable request. The code used in the velocity-map image analysis is available at https://github.com/PyAbel/PyAbel.

Competing interests

The authors declare no competing interests.

Footnotes

Peer review information Nature Communications thanks the anonymous reviewers for their contribution to the peer review of this work. Peer reviewer reports are available.

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary information

Supplementary information is available for this paper at 10.1038/s41467-019-13039-y.

References

  • 1.Shaik S, et al. Quadruple bonding in C2 and analogous eight-valence electron species. Nat. Chem. 2012;4:195–200. doi: 10.1038/nchem.1263. [DOI] [PubMed] [Google Scholar]
  • 2.Macrae RM. Puzzles in bonding and spectroscopy: the case of dicarbon. Sci. Prog. 2016;99:1–58. doi: 10.3184/003685016X14509452393033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.de Sousa DWO, Nascimento MAC. Is there a quadruple bond in C2 ? J. Chem. Theory Comput. 2016;12:2234–2241. doi: 10.1021/acs.jctc.6b00055. [DOI] [PubMed] [Google Scholar]
  • 4.Radius U, Breher F. To boldly pass the metal–metal quadruple bond. Angew. Chem. Int. Ed. 2006;45:3006–3010. doi: 10.1002/anie.200504322. [DOI] [PubMed] [Google Scholar]
  • 5.Zhong R, Zhang M, Xu H, Su Z. Latent harmony in dicarbon between VB and MO theories through orthogonal hybridization of 3[sigma]g and 2[sigma]u. Chem. Sci. 2016;7:1028–1032. doi: 10.1039/C5SC03437J. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Huntley DR, Markopoulos G, Donovan PM, Scott LT, Hoffmann R. Squeezing C–C bonds. Angew. Chem. Int. Ed. 2005;44:7549–7553. doi: 10.1002/anie.200502721. [DOI] [PubMed] [Google Scholar]
  • 7.Wu CJ, Carter EA. Ab initio thermochemistry for unsaturated C2 hydrocarbons. J. Phys. Chem. 1991;95:8352–8363. doi: 10.1021/j100174a058. [DOI] [Google Scholar]
  • 8.Hay PJ, Hunt WJ, Goddard WA. Generalized valence bond description of simple alkanes, ethylene, and acetylene. J. Am. Chem. Soc. 1972;94:8293–8301. doi: 10.1021/ja00779a002. [DOI] [Google Scholar]
  • 9.Shaik S, Danovich D, Braida B, Hiberty PC. A Response to a Comment by G. Frenking and M. Hermann on: "The Quadruple Bonding in C2 Reproduces the Properties of the Molecule". Chem. Eur. J. 2016;22:18977–18980. doi: 10.1002/chem.201602840. [DOI] [PubMed] [Google Scholar]
  • 10.Shaik S, Danovich D, Braida B, Hiberty PC. The quadruple bonding in C. Chem. Eur. J. 2016;22:4116–4128. doi: 10.1002/chem.201600011. [DOI] [PubMed] [Google Scholar]
  • 11.Karadakov PB, Kirsopp J. Magnetic shielding studies of C2 and C2 H2 ” support higher than triple bond multiplicity in C2. Chem. Eur. J. 2017;23:12949–12954. doi: 10.1002/chem.201703051. [DOI] [PubMed] [Google Scholar]
  • 12.Kepp KP. Trends in strong chemical bonding in C2, CN, CN-, CO, N2, NO, NO., and O2. J. Phys. Chem. A. 2017;121:9092–9098. doi: 10.1021/acs.jpca.7b08201. [DOI] [PubMed] [Google Scholar]
  • 13.Zou W, Cremer D. C2 in a box: determining its intrinsic bond strength for the X1Σg. ground state. Chem. Eur. J. 2016;22:4087–4099. doi: 10.1002/chem.201503750. [DOI] [PubMed] [Google Scholar]
  • 14.Grunenberg J. Quantum chemistry: Qudruply bonded carbon. Nat. Chem. 2012;4:154–155. doi: 10.1038/nchem.1274. [DOI] [PubMed] [Google Scholar]
  • 15.Frenking G, Hermann M. Comment on "the quadruple bonding in C2 reproduces the properties of the molecule". Chem. Eur. J. 2016;22:18975–18976. doi: 10.1002/chem.201601382. [DOI] [PubMed] [Google Scholar]
  • 16.Cooper DL, Penotti FE, Ponec R. Why is the bond multiplicity in C2 so elusive? Comput. Theor. Chem. 2015;1053:189–194. doi: 10.1016/j.comptc.2014.08.024. [DOI] [Google Scholar]
  • 17.Lloyd GM, Ewart P. High resolution spectroscopy and spectral simulation of C2 using degenerate four-wave mixing. J. Chem. Phys. 1999;110:385–392. doi: 10.1063/1.478070. [DOI] [Google Scholar]
  • 18.Kaminski CF, Hughes IG, Ewart P. Degenerate four-wave mixing spectroscopy and spectral simulation of C2 in an atmospheric pressure oxy-acetylene flame. J. Chem. Phys. 1997;106:5324–5332. doi: 10.1063/1.473563. [DOI] [Google Scholar]
  • 19.Brewer L, Hicks WT, Krikorian OH. Heat of sublimation and dissociation energy of gaseous C2. J. Chem. Phys. 1962;36:182–188. doi: 10.1063/1.1732293. [DOI] [Google Scholar]
  • 20.Douay M, Nietmann R, Bernath PF. New observations of the A1ΠuX1Σg. transition (phillips system) of C2. J. Mol. Spectrosc. 1988;131:250–260. doi: 10.1016/0022-2852(88)90236-6. [DOI] [Google Scholar]
  • 21.Martin M. C2 spectroscopy and kinetics. J. Photochem. Photobiol. A. 1992;66:263–289. doi: 10.1016/1010-6030(92)80001-C. [DOI] [Google Scholar]
  • 22.Chan M-C, Yeung S-H, Wang N, Cheung AS-C. Laser absorption spectroscopy of the d3Πgc3Σu. transition of C2. J. Phys. Chem. A. 2013;117:9578–9583. doi: 10.1021/jp3122769. [DOI] [PubMed] [Google Scholar]
  • 23.Szwarc M. The determination of bond dissociation energies by pyrolytic methods. Chem. Rev. 1950;47:75–173. doi: 10.1021/cr60146a002. [DOI] [PubMed] [Google Scholar]
  • 24.Ervin KM, Lineberger WC. Photoelectron spectra of dicarbon(1-) and ethynyl(1-) J. Phys. Chem. 1991;95:1167–1177. doi: 10.1021/j100156a026. [DOI] [Google Scholar]
  • 25.Arnold DW, Bradforth SE, Kitsopoulos TN, Neumark DM. Vibrationally resolved spectra of C2 C11 by anion photoelectron spectroscopy. J. Chem. Phys. 1991;95:8753–8764. doi: 10.1063/1.461211. [DOI] [Google Scholar]
  • 26.Bragg AE, Wester R, Davis AV, Kammrath A, Neumark DM. Excited-state detachment dynamics and rotational coherences of C2âĹŠ via time-resolved photoelectron imaging. Chem. Phys. Lett. 2003;376:767–775. doi: 10.1016/S0009-2614(03)01060-1. [DOI] [Google Scholar]
  • 27.Khuseynov D, Blackstone CC, Culberson LM, Sanov A. Photoelectron angular distributions for states of any mixed character: An experiment-friendly model for atomic, molecular, and cluster anions. J. Chem. Phys. 2014;141:124312. doi: 10.1063/1.4896241. [DOI] [PubMed] [Google Scholar]
  • 28.Shaik S, Rzepa HS, Hoffmann R. One molecule, two atoms, three views, four bonds? Angew. Chem. Int. Ed. 2013;52:3020–3033. doi: 10.1002/anie.201208206. [DOI] [PubMed] [Google Scholar]
  • 29.Valiev M, et al. Nwchem: a comprehensive and scalable open-source solution for large scale molecular simulations. Comput. Phys. Commun. 2010;181:1477–1489. doi: 10.1016/j.cpc.2010.04.018. [DOI] [Google Scholar]
  • 30.Shao Y, et al. Advances in molecular quantum chemistry contained in the Q-Chem 4 program package. Mol. Phys. 2015;113:184–215. doi: 10.1080/00268976.2014.952696. [DOI] [Google Scholar]
  • 31.Dunning TH. Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen. J. Chem. Phys. 1989;90:1007–1023. doi: 10.1063/1.456153. [DOI] [Google Scholar]
  • 32.Su P, et al. Bonding conundrums in the C2 molecule: a valence bond study. J. Chem. Theory Comput. 2011;7:121–130. doi: 10.1021/ct100577v. [DOI] [PubMed] [Google Scholar]
  • 33.Herzberg, G. Molecular Spectra and Molecular Structure - Spectra of Diatomic Molecules, Vol. I, pp. 110–140, 168–212 (Van Nostrand Reinhold Company, 1950).
  • 34.Buckingham AD, Orr BJ, Sichel JM. A Discussion on photoelectron spectroscopy - angular distributions and intensity in molecular photoelectron spectroscopy I. General theory for diatomic molecules. Phil. Trans. R. Soc. A. 1970;268:147–157. doi: 10.1098/rsta.1970.0068. [DOI] [Google Scholar]
  • 35.Amiot C, Chauville J, Maillard J-P. New analysis of the C2 Ballik-Ramsay system from flame emission spectra. J. Mol. Spectrosc. 1979;75:19–40. doi: 10.1016/0022-2852(79)90143-7. [DOI] [Google Scholar]
  • 36.Cooper J, Zare RN. Angular distribution of photoelectrons. J. Chem. Phys. 1968;48:942–943. doi: 10.1063/1.1668742. [DOI] [Google Scholar]
  • 37.Hanstorp D, Bengtsson C, Larson DJ. Angular distributions in photodetachment from O−. Phys. Rev. A. 1989;40:670–675. doi: 10.1103/PhysRevA.40.670. [DOI] [PubMed] [Google Scholar]
  • 38.Asmis KR, Taylor TR, Neumark DM. Anion photoelectron spectroscopy of BN−. Chem. Phys. Lett. 1998;295:75–81. doi: 10.1016/S0009-2614(98)00943-9. [DOI] [Google Scholar]
  • 39.Laws BA, Cavanagh SJ, Lewis BR, Gibson ST. NOO peroxy isomer exposed with velocity-map imaging. J. Phys. Chem. Lett. 2017;8:4397–4401. doi: 10.1021/acs.jpclett.7b02183. [DOI] [PubMed] [Google Scholar]
  • 40.Sanov A, Grumbling ER, Goebbert DJ, Culberson LM. Photodetachment anisotropy for mixed s-p states: 8/3 and other fractions. J. Chem. Phys. 2013;138:054311. doi: 10.1063/1.4789811. [DOI] [PubMed] [Google Scholar]
  • 41.Gozem S, et al. Photoelectron wave function in photoionization: plane wave or Coulomb wave? J. Phys. Chem. Lett. 2015;6:4532–4540. doi: 10.1021/acs.jpclett.5b01891. [DOI] [PubMed] [Google Scholar]
  • 42.Gozem, S. & Krylov, A. I. “ezDyson”. https://iopenshell.usc.edu/downloads/ezdyson (2016).
  • 43.Oana CM, Krylov AI. Cross sections and photoelectron angular distributions in photodetachment from negative ions using equation-of-motion coupled-cluster dyson orbitals. J. Chem. Phys. 2009;131:124114. doi: 10.1063/1.3231143. [DOI] [PubMed] [Google Scholar]
  • 44.Eppink ATJB, Parker DH. Velocity map imaging of ions and electrons using electrostatic lenses: application in photoelectron and photofragment ion imaging of molecular oxygen. Rev. Sci. Instrum. 1997;68:3477–3484. doi: 10.1063/1.1148310. [DOI] [Google Scholar]
  • 45.Cavanagh SJ, et al. High resolution velocity-map imaging photoelectron spectroscopy of the O fine-structure transitions. Phys. Rev. A. 2007;76:052708. doi: 10.1103/PhysRevA.76.052708. [DOI] [Google Scholar]
  • 46.Dedman CJ, Roberts EH, Gibson ST, Lewis BR. Fast 1 kV MOSFET switch. Rev. Sci. Instrum. 2001;72:3718–3720. doi: 10.1063/1.1389488. [DOI] [Google Scholar]
  • 47.Gascooke JR, Gibson ST, Lawrance WD. A "circularisation" method to repair deformations and determine the centre of velocity map images. J. Chem. Phys. 2017;147:013924. doi: 10.1063/1.4981024. [DOI] [PubMed] [Google Scholar]
  • 48.Hansen EW, Law P-L. Recursive methods for computing the Abel transform and its inverse. J. Opt. Soc. Am. A. 1985;2:510–520. doi: 10.1364/JOSAA.2.000510. [DOI] [Google Scholar]
  • 49.Hickstein DD, Gibson ST, Yurchak R, Das DD, Ryazanov M. A direct comparison of high-speed methods for the numerical abel transform. Rev. Sci. Instrum. 2019;90:065115. doi: 10.1063/1.5092635. [DOI] [PubMed] [Google Scholar]
  • 50.Van Duzor M, et al. Vibronic coupling in the superoxide anion: the vibrational dependence of the photoelectron angular distribution. J. Chem. Phys. 2010;133:174311. doi: 10.1063/1.3493349. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary Information (163.1KB, pdf)
Peer Review File (250.8KB, pdf)

Data Availability Statement

All data supporting the findings of this study are available within the Article and from the corresponding author upon reasonable request. Experimental data are available at https://physics.anu.edu.au/research/portal/vmi.

All of the computer code supporting the findings of this study are available from the corresponding author upon reasonable request. The code used in the velocity-map image analysis is available at https://github.com/PyAbel/PyAbel.


Articles from Nature Communications are provided here courtesy of Nature Publishing Group

RESOURCES