Parity non-conservation (PNC) due to the weak interaction is predicted to give rise to enantiomer dependent vibrational constants in chiral molecules, but the phenomenon has so far eluded experimental observation. The enhanced sensitivity of molecules to physics beyond the Standard Model (BSM) has led to substantial advances in molecular precision spectroscopy, and these may be applied to PNC searches as well. Specifically, trapped molecular ion experiments leverage the universality of trapping charged particles to optimize the molecular ion species studied toward BSM searches, but in searches for PNC, only a few chiral molecular ion candidates have been proposed so far. Importantly, viable candidates need to be internally cold, and their internal state populations should be detectable with high quantum efficiency. To this end, we focus on molecular ions that can be created by near threshold resonant two-photon ionization and detected via state-selective photo-dissociation. Such candidates need to be stable in both charged and neutral chiral versions to be amenable to these methods. Here, we present a collection of suitable chiral molecular ion candidates we have found, including CHDBrI+ and CHCaBrI+, that fulfill these conditions according to our ab initio calculations. We find that organo-metallic species have low ionization energy as neutrals and relatively high dissociation thresholds. Finally, we compute the magnitude of the PNC values for vibrational transitions for some of these candidates. An experimental demonstration of state preparation and readout for these candidates will be an important milestone toward measuring PNC in chiral molecules for the first time.

Following the observation of parity non-conservation (PNC) in β decay1 and atomic spectroscopy,2,3 the symmetry between the two mirror configurations of a chiral molecule was also predicted to be broken by the weak interaction.4,5 In most chiral molecules, the weak interaction is supposed to make one enantiomer slightly more energetically stable. Consequently, the effect alters the tunneling dynamics between enantiomers5–7 and creates enantiomer specific vibrational transition frequencies.8 This effect may also pertain to the mystery of how life originated with a specific handedness, with many examples of a single naturally occurring enantiomer.4,9–11 Observation of PNC in molecules, which has so far not been achieved in the laboratory, will improve our understanding of such phenomena and is a fundamentally important question in chemistry.

The majority of efforts to measure PNC in molecules have focused on neutral species,12–18 despite the advantages associated with charged molecules. However, recent advances in precision spectroscopy with trapped molecular ions that leverage long coherence times encourage the development of this platform.19–23 Moreover, the generality of ion trapping allows the selection of a molecular ion with optimized sensitivity toward the question at hand. This capability has been exploited in searches for CP (charge and parity) violations,19,21 searches for dark matter,24 and precision rotational spectroscopy,25 for example. Furthermore, methods for the production of molecular ions that are cold internally and externally have been demonstrated.21,26 Finally, chiral molecules with particularly asymmetric electronic wavefunctions, such as radical states, are predicted to have enhanced PNC by reducing the cancellation of PNC contributions.9,27 Such radicals have improved lifetimes as radical ions due to the suppression of reactions between the molecules.

The advantages associated with trapped ions have led to a few proposals to utilize chiral molecular ions to observe PNC.5,9,28–30 However, the scarcity of existing theoretical modeling for chiral molecular ions makes it challenging to choose a suitable candidate for such an experiment. One challenge is that molecular ions are often prepared through the ionization of neutrals and thus have a low dissociation threshold. The susceptibility to predissociation may shorten the interrogation times of vibrationally excited states.

Here, we conduct an ab initio study in search of suitable chiral molecular ions that can be prepared through state-selective near threshold photo-ionization (STPI) and have a sufficiently high dissociation threshold to facilitate vibrational precision spectroscopy with long interrogation times. We propose STPI as a pathway to create internally cold, chiral molecular ions, as it has been applied successfully to a few diatomic species.19,26,31,32 Using STPI on a cold neutral molecular beam only allows molecules in certain quantum states to enter the trapped ion ensemble. This step is important since the complexity of such polyatomic molecules makes them particularly difficult to cool, and the hope for a cycling transition diminishes when fewer symmetries exist.33–36 

In a quantum projection, noise limited statistical uncertainty on the measured frequency is given by
δf=1CτN,
(1)
where C is the contrast, τ is the coherence time, and N is the total number of molecules measured. Many details may cause C, τ, and N to vary, and technical noise may cause the observed experimental uncertainty to be far above this limit. The quality of state preparation and state resolved detection pertains to the contrast attained in C. The quantum efficiency of the detection process relates to N. Finally, the molecular ion stability is crucial to estimate since it may limit τ, eliminating one of the most important advantages of using charged, chiral molecules. The estimated statistical uncertainty δf compared to the vibrational PNC shift governs the experiment’s feasibility, along with consideration of sources of systematic uncertainty.

The following four separate stages are envisioned for the future experiment:

  1. Generation and cooling of neutral molecules in a supersonic expansion or other generic beam method.

  2. Ionization of the molecules via 1 + 1′—two-photon resonant ionization.

  3. Ion trapping and spectroscopy experiment.

  4. State-selective photo-dissociation spectroscopy with time-of-flight mass spectrometry for detection.

This paper focuses on finding suitable chiral molecular ion candidates for stages 2 and 4. A plan for step 3 is discussed in Ref. 37, but many other alternatives could be applied. Step 1 is straightforward for some of the candidates discussed in this work and more involved for others. Since the scheme addresses the same molecule in neutral and ionized forms, we refer to the neutral molecule and its companion charged molecule interchangeably. However, the precision spectroscopy stage is focused on the cation version of the molecule.

To limit the complexity of molecules, we study chiral molecules with up to five-atoms in a tetrahedral structure. Four-atom chiral molecules are not considered here as these are less likely to be chiral in both the neutral and charged configuration simultaneously, often having low transition state barriers between enantiomers.5,38

We find several suitable candidates with various spin multiplicities of the form CHXBrI+ where X ∈ {D, Ca} and isotopically chiral CHX79Br81Br+ where X ∈ {D, Ca, Yb}. All of these candidates are chiral in both neutral and charged forms. We discuss how our results indicate that these molecules fit stages 2 and 4. Additionally, we compute the magnitude of the expected PNC for some of these candidates here and in Ref. 39. We show, for example, that candidates where X ∈ {Li, Na} fit most but not all of the requirements and thus are not suitable candidates. Despite not explicitly investigating CHYbBrI+, we can infer that it is also amenable to the experiment by considering similar molecules such as CHYbBr2+ examined in the study.

Polyatomic molecules have many degrees of freedom, making a general method to internally cool them challenging to find. Most cold polyatomic molecule experiments resort to generic collision-based cooling methods such as supersonic expansion40 or buffer gas cooled beams,41 but these methods lower the internal temperature of molecules to a few kelvins only. Recent advances in laser cooling schemes suggest a promising pathway toward cold polyatomic molecules.33–35,42,43 In fact, the Franck–Condon factors for some isotopically chiral molecules have been found to be favorable for laser cooling,33 which places cold and controlled neutral chiral molecules within reach. For molecular ions that originate as neutrals, there is an alternative avenue by which they may be prepared, populating a low number of states through a filtering step discussed here.

The proposed scheme begins with a cold beam of the candidate molecules in neutral form. When the neutral molecules enter the ion trap, they are ionized by a two-color resonant process. The first photon frequency is resonant with a specific resolved molecular transition. Therefore, it excites only molecules in certain quantum states. The second photon ionizes the molecules that populate this intermediate state, either directly on the continuum near the ionization threshold or through an autoionizing Rydberg state. The molecular ion can only minimally heat up internally due to the low energy at which the electron is ejected or the selectivity of the autoionization process.

We discuss two avenues to achieve state selectivity. The first is to use a ro-vibronic (rovibrational and electronic) excitation for the first photon, followed by ionization by a second photon (Fig. 1, left). This places both of these photon energies in the UV (ultraviolet) or deep UV (DUV) range. The main risk with this method is that many organic molecules are vulnerable to dissociation in electronically excited states, which may occur before the second photon completes the ionization process. A second way of achieving cold molecular ions is to use a vibrational excitation in the mid-IR along the C–H stretch where the molecule is stable as the selective filtering step, followed by a second photon in the DUV or VUV (vacuum UV) to ionize the molecule (Fig. 1, right). This avoids the pre-dissociation of the molecule in the intermediate state since the process is energetically forbidden and the state density is significantly lower.

FIG. 1.

Resonant two-color near threshold ionization procedure schematically outlined for CHDBrI+ (STPI). The first photon excitation is resonant with a transition that ideally excites a single rotational state of the molecule. This can be a ro-vibronic transition, as shown on the left, or a rovibrational transition, as shown on the right. The selective filtering excitation is followed by a second (vacuum) ultraviolet photon that ionizes the molecule directly or through an autoionizing state such that the molecule’s final quantum state suffers minimal change.

FIG. 1.

Resonant two-color near threshold ionization procedure schematically outlined for CHDBrI+ (STPI). The first photon excitation is resonant with a transition that ideally excites a single rotational state of the molecule. This can be a ro-vibronic transition, as shown on the left, or a rovibrational transition, as shown on the right. The selective filtering excitation is followed by a second (vacuum) ultraviolet photon that ionizes the molecule directly or through an autoionizing state such that the molecule’s final quantum state suffers minimal change.

Close modal

Once the molecules are ionized and internally cold, they are still moving at hundreds of meters per second due to the neutral beam velocity. The ensemble can be decelerated by applying a uniform electric field to the molecular ion cloud.20,21 Subsequently, the ions are trapped by the same electrodes and separated from the rest of the neutral molecules. The ionization process serves as a selective step for specific quantum states since the molecules that are not excited by the first photon remain neutral and will not be decelerated and trapped, leading to a cold sample of molecular ions. This method is utilized for the HfF+ experiment20,21 following state-selective 1 + 1′ ionization and produces 40k ions in fewer than 4 quantum rotational states with a translational temperature of 2 K. This same method is also used to create state-selected N2+ with the 2 + 1′ scheme, where the ions are subsequently trapped by sympathetic cooling with laser cooled atomic ions.26 Throughout the manuscript, we refer to this state-selective near threshold photoionization step as STPI.

While the first photon transitions may be challenging to fully resolve for complex polyatomic molecules, any partial selectivity that arises in this step is likely to produce molecular ions that are very cold. Since the spectral width of nanosecond pulsed dye lasers is on the order of GHz, this tool is excellent for addressing individual rotational states, but the hyperfine structure will not be resolved. Population in hyperfine states is unlikely to harm the initial forms of our precision measurement and can be overcome by depletion or tailored pulses and polarization selections.44,45 Another advantage of trapped molecular ions is that there is sufficient time to perform multiple depletion processes, even when the available power is low.

Our scheme requires that chiral molecular ion candidates must be stable both as neutrals and as ions. For the 5-atom molecules we consider, the removal of an electron from such a small chiral molecule often makes the molecular ion very weakly bound, with a tendency to dissociate, particularly when the molecule is vibrationally excited (Sec. V A). In fact, as we will show, many candidates that are amenable to STPI dissociate far too easily to support vibrational spectroscopy. We, therefore, search for candidates with sufficiently high dissociation thresholds (>1.2 eV). We find that adding a metal atom to halogen substituted methane substantially increases the dissociation threshold while simultaneously reducing the ionization threshold of the neutral version of the species. For example, we find that for CHCaBrI+, the most favorable dissociation channel is I + CHCaBr+ at an energy of 2.55 eV as compared to CHDBrI+, which dissociates to I + CHDBr+ at 1.29 eV (Sec. V A). Simultaneously, the ionization threshold of CHCaBrI is more than 4 eV lower than that of its non-metallic counterparts (Sec. IV).

To detect the internal state of our chiral molecular ions, we turn to photodissociation due to the relatively low dissociation threshold. In an ion trap, state-selective photodissociation (SPD) serves as a highly efficient method for probing the internal states of the molecular ions by detecting the fragments via mass spectrometry.20,46–48 An alternative we pursue involves the detection of the internal state of the molecule using single-photon photo-dissociation. Upon dissociation, any internal rotational state of the molecule will be translated into a different kinetic energy of the photo-fragments. Detection of the photofragments may be performed by coupling a velocity map imaging setup to an ion trap, for example. Our initial design is promising, but details will be reported elsewhere. In this work, we check that the wavelengths associated with the dissociation processes would be feasible by calculating the dissociation energies for the various dissociation channels (Sec. V A).

Both STPI (Sec. IV) and SPD (Sec. V A) need to be developed experimentally, but this paper aims to rule out many candidates that would not fit these schemes and suggest some that do based on ab initio calculations.

The computed properties presented in the next sections are the vertical and adiabatic ionization energies and several excitation energies (EEs) of the neutral systems, in addition to the dissociation channel energies, isotopic vibrational modes, rotational constants, and transition state energies of the cation systems. Below are the details for all of these computations.

The energy differences we are interested in are the ionization potentials, adiabatic and vertical, the low-energy dissociations, the activation energy between the S and R enantiomers, and the low-level EEs of the neutral systems. All the electronic configurations considered herein have either a closed-shell structure with a singlet multiplicity or a radical configuration with a doublet multiplicity.

Radical systems are frequently studied using single reference methods. In particular, density functional theory (DFT) is often used to determine the geometrical structures and energy differences of reactions that include radicals.49–54 Sometimes, coupled clusters with singles, doubles, and perturbative triples [CCSD(T)] are used to benchmark the proper exchange-correlation functional for geometry optimization.52 Configuration interaction and Møller Plesset perturbation theory from second to fourth order can also be used to optimize structures.55,56 Notice that in Ref. 50, it was shown that DFT performs significantly worse than the unrestricted second order Møller Plesset perturbation method in both geometry optimization and interaction energy calculations for radicals. However, all the above-mentioned methods represent a single-reference approach in which the radicals’ treatment is based on an unrestricted wave-function (Hartree Fock or DFT). The unrestricted approximation for the radical wave-functions suffers, for example, from spin-contamination. Note that only in some studies is the very important ⟨S2⟩ expectation value, which gauges the spin-contamination of the wave-function, reported.56 

Therefore, in this study, we examine the following scheme for describing the closed-shell configurations as well as the radicals: the geometry is calculated at the second-order Møller Plesset (MP2) level, while the energies are obtained at the CCSD(T) level using the same basis sets [i.e., CCSD(T)//MP2]. However, since these open-shell radical doublet state calculations are based on the unrestricted Hartree–Fock wave-function with the well-known problematic behavior that manifests in spin-contamination, we carefully monitor this approximation. In addition, we perform equation-of-motion CCSD (EOM-CCSD) calculations to verify the validity of the CCSD(T)//MP2 scheme. Indeed, we find that for some radicals, the MP2 geometries are inconsistent with the EOM-CCSD ones (see Table S3 in the supplementary material). On the other hand, EOM-CCSD for ionization potential (EOM-IP-CCSD) or electron affinity (EOM-EA-CCSD), depending on the molecule under investigation, is a proper approach for describing radicals. It is a single-reference but multi-configuration approach that operates in Fock space, in which the reference and target states are treated in a balanced fashion, and as a result, it provides a well-defined spin state.57 Notice that EOM-CCSD analytic gradients are available within the Q-Chem program.58 

The eight low-lying EEs are calculated for the neutral molecules. For the molecules with a singlet state, we use EOM-CCSD for excitation energies (EOM-EE-CCSD), and for the system with a doublet neutral state, we use EOM-EA-CCSD. As discussed earlier, the EOM-CCSD energies are calculated at the EOM-CCSD geometry using the same basis set, which is suitable for doublet states, whereas for closed-shell singlet states, we use the CCSD energies at the MP2 geometry (in a few cases, we used the CCSD geometry since MP2 did not converge, and the difference between the two is negligible, as shown before in Ref. 59).

The harmonic frequencies of the optimized structures are calculated using EOM-IP-CCSD for doublet state cations and CCSD for singlets. The isotopic effect is explicitly considered in these calculations. In systems with two hydrogens, we use the masses of hydrogen and deuterium, and in the case of Br2, we use the different Br masses, 79Br and 81Br. In addition, we calculate these frequencies using the DFT ωB97M-V functional. The agreement between the two methods suggests that this functional is suitable for the frequency analysis of the systems at hand. Moreover, the ωB97M-V ⟨S2⟩ expectation values are found in the 0.75–0.756 range for all the reported doublet states. Therefore, we employ ωB97M-V to calculate the transition states between the S and R enantiomers in order to evaluate if the enantiomer states are time invariant. All the non-relativistic calculations are performed with the Q-Chem electronic-structure package.60 

We consider two basis sets of triple-zeta (TZ) and quadruple-zeta (QZ) quality. The frozen-core scheme is used within the coupled-cluster based calculations. The Dunning correlation consistent cc-pVXZ61,62 basis sets are used for all the light atoms (H, Li, C, O, Na, and Ca) within the XZ sets, where X = T or Q. For EEs, we use the augmented versions of these basis sets,63 aug-TZ and aug-QZ. For the heavier atoms (Cl, Br, and I), we use these basis sets paired with small-core pseudopotentials, i.e., cc-pVXZ-PP.64,65 For Yb, we use def2-XZVPP paired with a small-core pseudopotential.66 

These pseudopotentials were optimized to provide an accurate description of the Pauli repulsion of the cores, their Coulomb and exchange effects on the valence space, and scalar-relativistic corrections.64,65 Therefore, we also examined the spin-orbit effect by taking the difference between the fully relativistic calculations using the Dirac-Coulomb Hamiltonian including the Gaunt interaction, and the spin-free version,67 which provides results without spin-orbit coupling for the four-component Hamiltonian. The spin-orbit contributions to the adiabatic ionization potential (AIP) and the vertical ionization potential (VIP) of the ten molecules were calculated using the dyall.cv2z basis set within EOM-IP-CCSD. The eight low-lying EEs of the neutral molecules (2CHCaBr2 and 2CHCaBrI) with doublet multiplicity were calculated via EOM-EA-CCSD and the dyall.av2z basis set, except for Ca, for which we used dyall.v2z. For neutral molecules with singlet multiplicity, we used EOM-EE-CCSD/dyall.v2z; these include 1CH2Br2 (four excited states), 1CHLiBr2 (four excited states), 1CHLiBrI (four excited states), 1CHNaBr2 (five excited states), and 1CHNaBrI (two excited states). All in all, we calculated the spin-orbit effects for ten VIPs and AIPs, twenty-four EEs using EOM-EA-CCSD, and nineteen EEs using EOM-EE-CCSD. The maximal spin-orbit contribution among the molecules reduces the AIPs by 0.02 eV, the VIPs by 0.08 eV, the EEs using EOM-EA-CCSD by 0.003 eV, and the EOM-EE-CCSD EEs by 0.05 eV. The average shifts are at least fourfold smaller and have the same sign for each molecule. We conclude that the spin-orbit effect is negligible and cancels out since we investigated the energy difference between very similar chemical systems. Therefore, we report the values obtained by Q-Chem.60 All the relativistic calculations are performed using the relativistic electronic structure package DIRAC22.68 

Finally, we evaluate the stability of the selected cations by calculating their low lying dissociation channel energies. For fragments with singlet multiplicity, we use CCSD, and for doublet fragments, we use EOM-IP-CCSD or EOM-EA-CCSD using the QZ basis set and the Stuttgart/Cologne pseudopotentials. The spin-orbit corrections in this case are not negligible and are added to the non-relativistic dissociation energies. The spin-orbit effect was calculated only for the lowest dissociation channel of each system using the dyall.cv2z basis set with the same methodology. The average spin-orbit effect for the 11 dissociation energies is −0.15 eV, with the maximum value −0.27 eV, constituting less than 10% of the calculated values that neglect relativistic effects.

Benchmark calculations are presented in the supplementary material. In summary, the presented values and geometries were obtained using the following schemes: EOM-CCSD/QZ is used for the ionization energies (Sec. IV), and EOM-CCSD/aug-QZ for the excitation energies (Sec. IV). The cation’s vibrational frequencies using EOM-CCSD/TZ and ωB97M-V/TZ (Sec. V B). Transition states and their vibrational analysis using ωB97M-V/TZ (Sec. V B). Dissociation energies are calculated using EOM-CCSD/QZ; in addition, spin-orbit corrections are added via EOM-CCSD/dyall.cv2z (Sec. V A).

State-selective ionization combined with control of the excess energy in the process can be achieved using a resonant 1 + 1′ process, where 1 and 1′ denote the number of photons involved in each of the two stages in the process, and the “prime” tag indicates that the two stages use different photon energies. This notation is common for such REMPI (resonance enhanced multi-photon ionization) methods. The advantage of a 1 + 1′ process is that the power in each stage can be carefully tuned to avoid the competing 1 + 1 process. This is in contrast to the 2 + 1′ and 3 + 1′ processes, where the simultaneous interaction of multiple photons in the first stage usually necessitates focusing the laser beam, which precludes fine control of the first photon beam power since the ionization volume and power are coupled. To approximate the amenability of our candidates to 1 + 1′ resonant photo-ionization, we calculate the ionization threshold of the molecule as well as the electronically excited states for the resonant transition of the first photon.

The ionization threshold has an adiabatic and vertical component, and we calculate both for each of the initial sets of candidates: CHXBrI, where X ∈ {D, Li, Na, Ca}, and CHX79Br81Br, where X ∈ {D, Li, Na, Ca, Yb} (Table I). Between the AIP and the VIP, we expect to find resonant autoionizing states that are often used to control the emitted electron energy and quantum states while maintaining a large coupling to the continuum. Controlled ionization is also possible by choosing a near threshold energy for the 1′ photon.

TABLE I.

Vertical ionization potentials (VIPs) and adiabatic ionization potentials (AIPs) in eV calculated at the EOMCCSD/QZ level.

MoleculeVIPAIPMoleculeVIPAIP
CHD79Br81Br 10.68 10.22 CHDBrI 9.98 9.64 
CHLi79Br81Br 8.27 7.17 CHLiBrI 8.12 7.05 
CHNa79Br81Br 7.85 6.91 CHNaBrI 7.74 6.92 
CHCa79Br81Br 5.56 5.38 CHCaBrI 5.53 5.35 
CHYb79Br81Br 5.60 5.43    
MoleculeVIPAIPMoleculeVIPAIP
CHD79Br81Br 10.68 10.22 CHDBrI 9.98 9.64 
CHLi79Br81Br 8.27 7.17 CHLiBrI 8.12 7.05 
CHNa79Br81Br 7.85 6.91 CHNaBrI 7.74 6.92 
CHCa79Br81Br 5.56 5.38 CHCaBrI 5.53 5.35 
CHYb79Br81Br 5.60 5.43    

To estimate the resonant, filtering, and transition energies, we calculate the energies of the first eight vertical electronic excitations. For each candidate, these are listed in Table II. The first photon’s (1) frequency would need to be resonant with one of these states. Naturally, there are many specific rovibrational transitions associated with each state to choose from, but these energies provide a rough estimate of where these transitions with maximal Frank–Condon factors are expected.

TABLE II.

The eight lowest EEs calculated via EOM-EE-CCSD/aug-QZ for singlet states and via EOM-EA-CCSD/aug-QZ for doublets, where for singlets, we use the MP2/QZ geometry and for doublets EOM-EA-CCSD/QZ. For brevity, Br2 and H2 are used, which correspond to 79Br81Br and HD.

1CH2BrI1CHLiBrI1CHNaBrI2CHCaBrI1CH2Br21CHLiBr21CHNaBr22CHCaBr22CHYbBr2
4.93 4.55 3.41 1.67 5.99 3.91 3.72 1.80 1.97 
4.96 4.53 4.20 1.93 6.06 4.66 4.33 1.92 2.11 
6.05 4.68 4.63 1.94 6.33 5.18 5.03 1.97 2.18 
6.06 5.03 4.89 2.46 6.51 5.61 5.46 2.51 2.87 
6.84 5.44 5.27 2.53 7.47 5.82 5.53 2.56 2.92 
6.91 5.68 5.44 2.75 7.52 5.97 5.65 2.92 3.04 
6.95 5.80 5.52 3.33 7.56 5.98 5.66 3.39 3.61 
7.23 5.85 5.54 3.37 7.61 6.32 5.98 3.43 3.62 
1CH2BrI1CHLiBrI1CHNaBrI2CHCaBrI1CH2Br21CHLiBr21CHNaBr22CHCaBr22CHYbBr2
4.93 4.55 3.41 1.67 5.99 3.91 3.72 1.80 1.97 
4.96 4.53 4.20 1.93 6.06 4.66 4.33 1.92 2.11 
6.05 4.68 4.63 1.94 6.33 5.18 5.03 1.97 2.18 
6.06 5.03 4.89 2.46 6.51 5.61 5.46 2.51 2.87 
6.84 5.44 5.27 2.53 7.47 5.82 5.53 2.56 2.92 
6.91 5.68 5.44 2.75 7.52 5.97 5.65 2.92 3.04 
6.95 5.80 5.52 3.33 7.56 5.98 5.66 3.39 3.61 
7.23 5.85 5.54 3.37 7.61 6.32 5.98 3.43 3.62 

Once excited to the intermediate state, we can use the AIP and VIP to estimate the energy of the 1′ photon that is needed to ionize the molecules by subtracting the respective electronic state energy. Figure 2 shows the predicted photon energies needed in the 1 + 1′ process for each candidate. The energies of the electronically excited states are depicted by circles [first photon energy (1)], and the ionizing photon energy (1′) is depicted by an error bar that stretches between the AIP and the VIP. The color-coding of the markers of the 1 and 1′ photons corresponds to different electronic excitations (Table II). Excitations one through eight are plotted in ascending order and colored black, red, light green, blue, orange, purple, magenta, and dark green, respectively. The excitation energy and first photon energy are equal, which serves as a key to help distinguish between the different intermediate states.

FIG. 2.

Photon energies for 1 + 1′ REMPI ionization of CHXBrI where X ∈ {D, Li, Ca, Na} (a) and CHX79Br81Br where X ∈ {D, Li, Ca, Na, Yb} (b) are shown. The electronic excitation energies that approximately represent the first photon energy are depicted by the circles. After the electronic excitation, the remaining energy in the AIP and the VIP is depicted by the left and right edges of the error bar, which correspond to the second photon energy range. The electronic excitation symbol and remaining energy for ionization markings are color-coded per excitation. The vertical, dashed lines denote the maximal photon energy reachable with a tunable dye laser directly (blue), frequency doubled (red), and frequency tripled (yellow). For brevity, Br2 and H2 are used, which correspond to 79Br81Br and HD.

FIG. 2.

Photon energies for 1 + 1′ REMPI ionization of CHXBrI where X ∈ {D, Li, Ca, Na} (a) and CHX79Br81Br where X ∈ {D, Li, Ca, Na, Yb} (b) are shown. The electronic excitation energies that approximately represent the first photon energy are depicted by the circles. After the electronic excitation, the remaining energy in the AIP and the VIP is depicted by the left and right edges of the error bar, which correspond to the second photon energy range. The electronic excitation symbol and remaining energy for ionization markings are color-coded per excitation. The vertical, dashed lines denote the maximal photon energy reachable with a tunable dye laser directly (blue), frequency doubled (red), and frequency tripled (yellow). For brevity, Br2 and H2 are used, which correspond to 79Br81Br and HD.

Close modal

To understand if it is feasible to realize the wavelengths in the lab, we compare the energies of the photons to the tuning range of a dye laser pumped by a 532 nm source. The choice of dye lasers originates in their broad tunability over the range and relatively narrow linewidth that commonly allows the resolution of individual rotational lines. The maximal lasing frequencies for tunable dye lasers are depicted by vertical dashed lines including their frequency conversion add-ons such as doubling and tripling.

In the second approach discussed to achieve the filtering step, the molecule would be excited along a specific rovibrational transition (Fig. 1 right). The vibrational excitation needs to be sufficiently high to ensure that the intermediate state is not thermally populated. The relevant energies for the second transition are the AIPs and VIPs (Table I) minus the vibrational excitation. This means that the second photon would need to be in the VUV range for most of the molecules except for the Ca and Yb substituted candidates, where the transition is reachable with a tripled dye laser.

A clear trend emerges from Table II and Fig. 2. Substituting the deuterium with an alkali atom reduces the energies of the excited electronic states by more than 1 eV. Replacing deuterium with Ca or Yb further reduces this energy since these candidates are now radicals in the neutral state. Furthermore, these substitutions cause the ionization energy to dramatically drop (see Table I). This makes this subset of candidates particularly appealing from a state preparation perspective.

For the scheme to work, one of the important steps is for the intermediate transition to be resolved. This naturally becomes challenging with the growth of molecular complexity, and certainly, the molecules discussed here are far from simple. However, any sort of partially resolved transition should already assist in reducing the target molecular ion’s internal temperature in the process. Moreover, rovibrational transitions tend to be resolved for similar molecules.69 

In this section, we consider the candidates in charged form to test their amenability for precision spectroscopy in search of PNC. We discuss the binding energy of the candidates (Sec. V A) as well as the structure of the molecules (Sec. V B).

Since the candidate cations are produced by shedding an electron, as we describe in Sec. IV, we suspect that their binding energy might be relatively low. For the chiral molecular ion to be a viable candidate, we must verify that it will not fragment too easily, and certainly not when undergoing vibrational excitation. This is because one of the main avenues to search for PNC is through vibrational spectroscopy and comparison between enantiomers.8 If the molecular ion unintentionally dissociates during the spectroscopy stage, the precision of the measurement will be substantially reduced, which lowers its appeal as a candidate.

Here, we calculate the dissociation threshold for the various dissociation channels for the candidates. The resulting thresholds for the lowest energy dissociative channel are given in Table III. The dissociation channels were calculated by taking the difference between the cation energy and the sum of the fragment energies. EOM-CCSD/QZ//EOM-CCSD/QZ (IP, ionization potential, or EA, electron attachment) is used for systems with doublet multiplicity and CCSD/QZ//MP2/QZ for singlet state molecules. The relativistic spin-orbit coupling correction is added on top of the non-relativistic results. This correction is obtained by taking the difference between results obtained using the full relativistic Hamiltonian and its spin-free version; the same approach but a smaller basis set was used for this calculation (see Sec. III for details). For CHLiBr2+ and CHNaBr2+, we find two channels with similar dissociation energies, and both are listed. For CHYbBr2+1, the EOM-CCSD calculation of 2CHYbBr+ did not converge, so we present the CCSD(T) results instead. The spin multiplicity of the system, which is either singlet or doublet, is noted in a superscript before the molecular formula.

TABLE III.

Lowest dissociation energies in eV of the cations in their ground states, doublet states calculated using the EOM-CCSD/QZ at the EOM-CCSD/QZ geometry, and singlets using the CCSD/QZ at the MP2/QZ geometry (values marked with an “∗” are calculated at the CCSD(T)/QZ//MP2/QZ level instead of EOM-CCSD/QZ). spin-orbit corrections are obtained using the same approach as with the dyall.cv2z basis set. For brevity, Br2 and H2 are used, which correspond to 79Br81Br and HD.

Dissociation channelEOMCCSpin-orbitTotal
2CH2BrI+1CH2Br+ + 21.50 −0.21 1.29 
2CHLiBrI+1CHLiBr+ + 22.24 −0.27 1.97 
2CHNaBrI+1CHNaBr+ + 22.54 −0.28 2.26 
1CHCaBrI+2CHCaBr+ + 22.82 −0.27 2.55 
CH22Br2+ 1CH2Br+ + 2Br 1.43 −0.14 1.29 
2CHLiBr2+ 1CLiBrH+ + 2Br 2.72 −0.16 2.56 
2CHLiBr2+ 2CLiBr+ + 1HBr 2.68 −0.04 2.62 
2CHNaBr2+ 2CNaBr+ + 1HBr 2.79 −0.04 2.75 
2CHNaBr2+ 1CHNaBr+ + 2Br 2.96 −0.15 2.81 
1CHCaBr2+ 2CHCaBr+ + 2Br 3.25 −0.14 3.11 
1CHYbBr2+ 2CHYbBr+ + 2Br *3.26 −0.17 3.09 
Dissociation channelEOMCCSpin-orbitTotal
2CH2BrI+1CH2Br+ + 21.50 −0.21 1.29 
2CHLiBrI+1CHLiBr+ + 22.24 −0.27 1.97 
2CHNaBrI+1CHNaBr+ + 22.54 −0.28 2.26 
1CHCaBrI+2CHCaBr+ + 22.82 −0.27 2.55 
CH22Br2+ 1CH2Br+ + 2Br 1.43 −0.14 1.29 
2CHLiBr2+ 1CLiBrH+ + 2Br 2.72 −0.16 2.56 
2CHLiBr2+ 2CLiBr+ + 1HBr 2.68 −0.04 2.62 
2CHNaBr2+ 2CNaBr+ + 1HBr 2.79 −0.04 2.75 
2CHNaBr2+ 1CHNaBr+ + 2Br 2.96 −0.15 2.81 
1CHCaBr2+ 2CHCaBr+ + 2Br 3.25 −0.14 3.11 
1CHYbBr2+ 2CHYbBr+ + 2Br *3.26 −0.17 3.09 

Additional, higher dissociation channel energies are listed in Table VI.

To estimate the minimal acceptable dissociation threshold, we can consider, for example, the worst-case scenario of a vibrational excitation along the C–H stretch mode where the photon energy is roughly 0.4 eV (explicit calculations in Sec. V B). Therefore, the energy of the first vibrational excitation for the C–H stretch is 0.4 eV above the minimum electronic energy of CHDBrI+. In addition, the presence of the excitation laser may dissociate the molecule in the first excited vibrational state if it couples between that state and the continuum. This would cause a loss of population during the spectroscopy. However, this loss is manageable since, in a Ramsey spectroscopy experiment, this dissociation would be limited to the π2 pulses37 but could also occur from black body radiation. We would nonetheless prefer that the dissociation threshold be higher than 0.8 eV, or about double the highest vibrational mode energy. Alternatively, using a lower energy vibrational mode is also possible. For a dissociation channel such as CH2BrI+ → CH2Br+ + I, the zero point energies (ZPEs) of CH2BrI+ and the CH2Br+ fragment are similar since the iodine is associated with the lowest energy vibrational modes. Therefore, we neglect the contribution of the ZPEs to the reduction of the dissociation threshold of the molecules since their difference is small relative to the threshold. The energies for the vibrational modes for the cation-candidates are shown in Table IV. For the metal substituted candidates, we can use the same cutoff of 0.8 eV as the molecular properties and dissociation channels are similar to the other systems. The vibrational energies were calculated at the coupled-cluster level as well as using density functional theory with the ωB97M-V functional. The agreement between the two approaches suggests that ωB97M-V is also suitable for calculating the energies and frequencies of transition states between the two enantiomers.

TABLE IV.

Vibrational transition frequencies of the cations in cm−1. EOM-IP-CCSD/TZ (EOMCC) for doublet state cations and CCSD/TZ (CCSD) for singlet. In addition, ωB97M-V/TZ calculations are presented as DFT. Here, the isotopic effect is explicitly considered using the mass of deuterium and the different Br masses, i.e., Br2 = 79Br81Br. The left column represents the mode number. Vibrational scaling factors of 0.942–0.946 for coupled clusters and 0.946–0.949 for ωB97M-V should be applied,70,71 as verified by our comparison of the computed vibrational transition frequencies for neutral CHDBrI and CHDBr2 to experimental results in Table S8 of supplementary material.

2CHDBrI+2CHDBr2+1CHCaBrI+1CHCaBr2+1CHYbBr2+
#EOMCCDFTEOMCCDFTCCSDDFTCCSDDFTCCSDDFT
151.2 155.9 190.9 193.8 129.3 132.0 149.4 147.5 113.5 110.6 
523 535 561 551 145.9 144.7 150.4 150.3 118.8 118.8 
616 620 642 649 205 205 218 216 182.8 180.2 
716 714 760 755 428 442 462 470 445 436 
794 791 771 796 479 500 489 509 484 479 
1104 1097 1057 1118 511 520 519 525 489 488 
1284 1267 1295 1276 912 894 956 932 953 938 
2362 2349 2365 2353 1095 1091 1116 1110 1111 1105 
3225 3212 3228 3212 3202 3182 3208 3191 3206 3189 
2CHDBrI+2CHDBr2+1CHCaBrI+1CHCaBr2+1CHYbBr2+
#EOMCCDFTEOMCCDFTCCSDDFTCCSDDFTCCSDDFT
151.2 155.9 190.9 193.8 129.3 132.0 149.4 147.5 113.5 110.6 
523 535 561 551 145.9 144.7 150.4 150.3 118.8 118.8 
616 620 642 649 205 205 218 216 182.8 180.2 
716 714 760 755 428 442 462 470 445 436 
794 791 771 796 479 500 489 509 484 479 
1104 1097 1057 1118 511 520 519 525 489 488 
1284 1267 1295 1276 912 894 956 932 953 938 
2362 2349 2365 2353 1095 1091 1116 1110 1111 1105 
3225 3212 3228 3212 3202 3182 3208 3191 3206 3189 

All the molecules CHXBrI+ where X ∈ {D, Li, Ca, Na} and CHXBr2+ where X ∈ {D, Li, Ca, Na, Yb} have dissociation thresholds above 0.8 eV. The higher the threshold, the less susceptible the molecule will be to spontaneous dissociation by absorbing a photon from the black body. This is in contrast to the other chiral molecular ions we have computed, such as CHCl2F+, CHBrClF+, CHBr2F+, and CHIBrCl+ (Table V), which range from instability with a negative energy for dissociation to weakly bound molecular ions with thresholds below the 0.8 eV cutoff. These molecules would likely not support spectroscopy of the C–H stretch transition and other lower energy modes without compromising the molecule’s lifetime. Notably, we find that all the metal substituted molecules are very stable with a high dissociation threshold. This high stability, combined with the predicted convenient wavelengths for STPI (Sec. IV), makes them particularly appealing for precision spectroscopy.

TABLE V.

The lowest CCSD(T) dissociation energies at the MP2 geometries using the TZ and QZ basis sets for the unsuitable candidates. In this case, CHIBrCl+ is close to the stability cutoff, but notice that the inclusion of the spin-orbit contribution will reduce its threshold further.

MoleculeTZQZMoleculeTZQZ
2CHCl2F+ → 2CHBrClF+ → 
2CCl2+ + 1HF −0.06 −0.11 1CClHF+ + 2Br 0.42 0.53 
2CClHF+ + 2Cl −0.17 −0.09    
2CHBr2F+ → 2CHIBrCl+ → 
1CBrHF+ + 2Br 0.01 0.10 1CHBrCl+ + 20.62 0.74 
MoleculeTZQZMoleculeTZQZ
2CHCl2F+ → 2CHBrClF+ → 
2CCl2+ + 1HF −0.06 −0.11 1CClHF+ + 2Br 0.42 0.53 
2CClHF+ + 2Cl −0.17 −0.09    
2CHBr2F+ → 2CHIBrCl+ → 
1CBrHF+ + 2Br 0.01 0.10 1CHBrCl+ + 20.62 0.74 

The dissociation energies for the candidates (Table V) are computed at the CCSD(T) at the MP2 geometries. Since these values are well below our 0.8 eV cutoff and these candidates are unlikely to be suitable, we did not perform the EOM-CCSD calculations as performed for the candidates in Table III.

With molecule stability established, we turn to discuss dissociation for state-selective detection. Using dissociation for state-selective detection is common in trapped molecular ion experiments due to the low dissociation thresholds.20,46–48 Since the fragments have a different mass than the parent molecular ion, they are straightforward to distinguish with high quantum efficiency. The rise of ion traps that are coupled to mass spectrometers72,73 facilitates a high quantum efficiency avenue to detect these fragments and thus measure the internal state distribution of the parent molecular ion.

It is challenging to predict the dominant dissociation channel for these complicated five-atom molecules that have a large density of states originating in multiple channels above the dissociation threshold. However, if we consider the different dissociation channels, we see that the wavelengths for a two photon resonant dissociation process should be feasible. For example, for CHDBrI+, the dissociation energy is 1.29 eV, which is comparable to commercially available diode lasers in the near infrared, which range from 0.8 to 1.8 eV. These lasers can be combined with a selective excitation step realized with microwaves or a second diode laser to create the state-selective 1 + 1′ dissociation process, similarly to the STPI described in Sec. IV. The 1′ photon can also be tuned to the most favorable dissociation process for detection, which is not necessarily immediately above the dissociation threshold. In Table VI, we show additional dissociation channel thresholds for the promising candidates to help guide the experiment. In this case, the spin-orbit contribution is not included; however, we expect a similar shift of the dissociation energies to the shifts noted in Table III. Therefore, these values represent an estimate for experimental studies; a more refined computation will be necessary in the future.

TABLE VI.

Multiple dissociation channel energies are listed in eV. The doublet states are calculated at the EOM-CCSD/QZ level and the singlet states at CCSD/QZ. Values marked with an are calculated at the CCSD(T)/QZ level.

Dissociation channelEOMCCDissociation channelEOMCCDissociation channelEOMCC
2CH2Br2+  2CHLiBr2+  2CHNaBr2+ 
1CH2Br+ + 2Br 1.43 2CLiBr+ + 1HBr 2.67 1CHNaBr+ + 2Br 2.96 
2CHBr+ + 1HBr 2.48 1CHLiBr+ + 2Br 2.71 2CNaBr+ + 1HBr 2.79 
  2CHBr+ + 1LiBr 3.51 2CHBr+ + 1NaBr 3.56 
2CH2BrI+ →  2CHLiBrI+ →  2CHNaBrI+ → 
1CH2Br+ + 21.50 1CHLiBr+ + 22.24 1CHNaBr+ + 22.54 
1CH2I+ + 2Br 1.76 1CHLiI+ + 2Br 2.69 1CHNaI+ + 2Br 3.01 
2CHI+ + 1HBr 2.74 2CLiI+ + 1HBr 2.68 2CNaI+ + 1HBr 2.90 
2CHBr+ + 1HI 3.12 2CLiBr+ + 1HI 2.77 2CNaBr+ + 1HI 2.94 
1CHCaBrI+ →  1CHCaBr2+  1CHYbBr2+ 
2CHCaBr+ + 22.82 2CHCaBr+ + 2Br 3.25 2CHYbBr+ + 2Br 3.26 
2CHCaI+ + 2Br 3.27 1CCaBr+ + 1HBr 4.42 1CHBr2+ + 1Yb 4.70 
  1CBr2 + 1CaH+ 4.81 1CBr2 + 1YbH+ 4.73 
  1CHBr2+ + 1Ca 4.96 1CYbBr+ + 1HBr 4.44 
Dissociation channelEOMCCDissociation channelEOMCCDissociation channelEOMCC
2CH2Br2+  2CHLiBr2+  2CHNaBr2+ 
1CH2Br+ + 2Br 1.43 2CLiBr+ + 1HBr 2.67 1CHNaBr+ + 2Br 2.96 
2CHBr+ + 1HBr 2.48 1CHLiBr+ + 2Br 2.71 2CNaBr+ + 1HBr 2.79 
  2CHBr+ + 1LiBr 3.51 2CHBr+ + 1NaBr 3.56 
2CH2BrI+ →  2CHLiBrI+ →  2CHNaBrI+ → 
1CH2Br+ + 21.50 1CHLiBr+ + 22.24 1CHNaBr+ + 22.54 
1CH2I+ + 2Br 1.76 1CHLiI+ + 2Br 2.69 1CHNaI+ + 2Br 3.01 
2CHI+ + 1HBr 2.74 2CLiI+ + 1HBr 2.68 2CNaI+ + 1HBr 2.90 
2CHBr+ + 1HI 3.12 2CLiBr+ + 1HI 2.77 2CNaBr+ + 1HI 2.94 
1CHCaBrI+ →  1CHCaBr2+  1CHYbBr2+ 
2CHCaBr+ + 22.82 2CHCaBr+ + 2Br 3.25 2CHYbBr+ + 2Br 3.26 
2CHCaI+ + 2Br 3.27 1CCaBr+ + 1HBr 4.42 1CHBr2+ + 1Yb 4.70 
  1CBr2 + 1CaH+ 4.81 1CBr2 + 1YbH+ 4.73 
  1CHBr2+ + 1Ca 4.96 1CYbBr+ + 1HBr 4.44 

A second avenue we consider for detection is dissociation by a single photon process. In this approach, the fingerprints of the internal state of the molecular ion will appear in the kinetic energies and angular distributions of the photo-fragments. The parent molecule’s internal state will affect the kinetic energy of the photofragments due to energy conservation since a single photon is absorbed. We aim to probe two rotational states in this manner.37 For distinguishing between vibrational states, this method may not be as straightforward since the energy difference may remain within one of the photofragments. While measurement of the kinetic energy of photo-fragments has been achieved when dissociating trapped molecular ions,20 the energy resolution needed to resolve individual rotational states of the parent molecule would need to be below 10 m/s for the candidates considered. An ion trap that is optimized for the detection of photofragment energies by coupling to a velocity map imaging detector should be able to reach the required resolutions. Currently, we are building such an apparatus. The ion trajectory simulations, which will be reported elsewhere, indicate that this is feasible.

As indicated in Table III for the single photon dissociation process, realizing the required wavelengths is also possible with a single photon for a wide range of diode and pulsed dye lasers.

The geometry of the molecular cation is one of the most important properties that must be verified to support a measurement of PNC through a comparison of vibrational transition frequencies. For many 4-atom candidates, such as hydrogen peroxide, the chiral geometry does not survive the removal of an electron because the cation has a planar structure.38 However, even the neutral form of hydrogen peroxide does not support time-invariant chiral states since it has a relatively low transition state between enantiomers relative to twist mode energy splitting. The eigenstates of such a system are the symmetric and anti-symmetric superposition states of the S and R molecule configurations.5 

For our candidates, we search for stable chiral geometry, which can be determined if the transition state between the two enantiomers is sufficiently high. First, we calculate the minimal-energy geometry of the molecular ion candidates to verify that they are indeed chiral. These geometries can be seen in Fig. 3. While all of these molecules have a chiral structure, the geometries of both the lithium and sodium substituted molecules are very close to planar. This near planar geometry hints at a low barrier between the two enantiomers. Indeed, a search for a planar transition state (TS) reveals that its energy is only 0.01 and 0.03 eV for CHLiBrI+ and CHNaBrI+, respectively. The transition states are planar and have a single imaginary frequency out of the nine normal modes. Therefore, we can conclude through comparison with the zero point energies of the molecules that the chiral ground state of these molecules will not be time invariant.

FIG. 3.

Geometries, EOM-CC/QZ for doublet cations and MP2/QZ for singlets. The top five molecules have a significant transition state (TS) energy examined at the ωB97M-V/TZ level. The Li and Na substituted molecules have near planar geometries with a very low TS energy.

FIG. 3.

Geometries, EOM-CC/QZ for doublet cations and MP2/QZ for singlets. The top five molecules have a significant transition state (TS) energy examined at the ωB97M-V/TZ level. The Li and Na substituted molecules have near planar geometries with a very low TS energy.

Close modal

On the other hand, the rest of the candidates whose chiral structure is more acute survived this test. Their geometries and transition state energies are shown in Fig. 3. For CHDBrI+ and CHDBr2+, the transition state energies at 1.30 and 1.39 eV are significantly higher than all the vibrational mode energies in the system, including the first excited state of the C–H stretch mode. For the other metal substituted candidates, the transition states are all above 0.7 eV, which is higher than the zero point energy for all the molecules. In particular, all the bending modes that overlap with the enantiomer mutation coordinate have at least 5 states below the transition state barrier, and this barrier scale is approximately equal to the energy of the first excited state of the C–H stretch mode.

Another effect that might limit the lifetime of excited vibrational states is known as intra-molecular vibrational redistribution (IVR).74 To a certain extent, the rate of IVR is determined by the vibrational state density at the excited energy. Taking all the different vibrational combinations, we find that the state density at the energy of the first excited state of the C–H stretch mode (v9 = 1, where v9 is the vibrational quantum number of mode No. 9) is 1.3 and 0.9 states per cm−1 for CHDBrI+ and CHDBr2+, respectively. At the energy of the excited state of the C–D stretch mode (v8 = 1), the densities are 3 times lower. For CHCaBrI+ and CHCaBr2+, the density at v9 = 1 is significantly higher, ranging from 35 to 25 states per cm−1, but for the C–H bend modes (v8 = 1), the density drops below 1 state per cm−1 for both calcium substituted molecules. These densities are much higher relative to the effective density that should be considered as redistribution is less likely to proceed to combinations of more than two modes. An accidental overlap is unlikely for these modes when compared to the natural linewidth of the vibrational excited state, even when considering the densities resulting from combinations of all modes.

These are good indications for long-lived excited vibrational states. We do not explicitly estimate the state lifetimes in the current manuscript. However, spectroscopy using lower energy vibrational modes would enhance the lifetimes with respect to dissociation.

Additionally, in order to support 3-wave mixing schemes such as the one presented in Ref. 37, we also need the rotational constants of these molecular ions. These are shown in Table VII. The two deuterated molecules have approximately prolate symmetric tops, while the Yb substituted molecule may be approximated as having an oblate top.

TABLE VII.

Rotational constants of the cations in MHz. Here, the isotopic effect is explicitly considered using the mass of deuterium and the two different Br isotope masses, i.e., Br2 = 79Br81Br.

CHDBrI+CHDBr2+CHCaBrI+CHCaBr2+CHYbBr2+
14 750 15 610 2550 2705 1170 
1170 1645 830 1175 1070 
1095 1510 660 870 585 
CHDBrI+CHDBr2+CHCaBrI+CHCaBr2+CHYbBr2+
14 750 15 610 2550 2705 1170 
1170 1645 830 1175 1070 
1095 1510 660 870 585 

Finally, we discuss the sensitivity of our candidates to Zeeman shifts. The chiral molecular ions have two spin multiplicities for the candidates. Closed-shell molecules such as CHCaBrI+ will have a small magnetic moment, which makes them favorably immune to magnetic field drifts.

Although we have skipped the computations for CHYbBrI+ in this work, we can compare the results between BrI-containing molecules and Br2-containing molecules. In particular, we observe the similarities between CHCaBr2+ and CHCaBrI+ as well as between CHCaBr2+ and CHYbBr2+ with respect to the ionization wavelengths, dissociation threshold, and transition state energy. These similarities in molecules that are explicitly examined lead us to infer that CHYbBrI+ is also a promising candidate from an experimental perspective.

Sections IV and V discussed pathways to create cold molecules, which affect the contrast and the stability of the molecular ions, which relate to the accessible coherence time in Eq. (1). However, to estimate the number of molecules that need to be measured in a precision measurement assuming that the quantum projection limit is achieved,37 we must compare the magnitude of the expected shift due to PNC in the vibrational frequency to the expected precision δf.

Here, we present the PNC calculations for the different vibrational modes of CHCaBrI+. Table VIII shows the PNC shifts expected for the various modes, with 100 and 29 mHz shifts for the most relevant modes for precision spectroscopy, the C–H bend and stretch modes (No. 8 and 9 in Table IV, respectively). We also computed the PNC shifts for the doubly isotopically chiral CHD79Br81Br+, which are significantly lower, probably due to the relatively low mass of its constituents. However, its symmetric structure may simplify the molecule’s spectrum, promoting other aspects of the experiment. In contrast, the PNC frequency shift in CHDBrI+ is very large, on the order of 1 Hz for most of its higher energy modes, and is fully reported in Ref. 39.

TABLE VIII.

PNC frequency difference between enantiomers for each vibrational mode of CHCaBrI+ and CHD79Br81Br+ in Hz.

Mode #CHCaBrI+CHD79Br81Br+
−2.8 × 10−2 4.3 × 10−5 
−5.6 × 10−2 −3.1 × 10−6 
1.5 × 10−1 3.3 × 10−3 
6.6 × 10−2 −3.0 × 10−3 
−2.0 × 10−1 2.9 × 10−3 
−2.0 × 10−3 −7.7 × 10−3 
1.9 × 10−1 −4.8 × 10−3 
−1.0 × 10−1 1.0 × 10−2 
−2.9 × 10−2 −1.3 × 10−2 
Mode #CHCaBrI+CHD79Br81Br+
−2.8 × 10−2 4.3 × 10−5 
−5.6 × 10−2 −3.1 × 10−6 
1.5 × 10−1 3.3 × 10−3 
6.6 × 10−2 −3.0 × 10−3 
−2.0 × 10−1 2.9 × 10−3 
−2.0 × 10−3 −7.7 × 10−3 
1.9 × 10−1 −4.8 × 10−3 
−1.0 × 10−1 1.0 × 10−2 
−2.9 × 10−2 −1.3 × 10−2 

For the PNC calculations, the molecular geometry was optimized on the ωB97M-V/Def2-TZVPP level of theory using Q-Chem 5.2.2.75 

In order to obtain the PV contributions to the total energies, we carried out single point relativistic DFT calculations using the DIRAC23 program.76 In order to conserve computational effort, we replaced the 4-component Dirac Hamiltonian by the exact 2-component (X2C) Hamiltonian, where the large and small components are exactly decoupled and the positive energy spectrum of the 4-component Hamiltonian is reproduced within numerical precision. In this scheme, the spin-same-orbit interactions are introduced in a mean-field fashion by using the AMFI procedure.77 We used the CAM-B3LYP* functional, whose parameters were adjusted by Thierfelder et al., to reproduce the PV energy shifts obtained using coupled cluster calculations.78 Dyall’s v4z basis sets were used for all the elements.79 Alongside the relativistic absolute energy at each geometry, these calculations also yield the PV energy contribution, EPV.

To calculate the vibrational parity violating frequency shifts, the parity violating shifts of the vibrational ground and first excited states are needed, which we obtained as follows. We performed relativistic single-point calculations at 11 equally spaced points between −0.5 and 0.5 Å along the selected normal mode. This yields the potential energy and the parity violating energy as a function of the normal coordinate q; we fitted polynomials to these points in order to create smooth potential and parity violating energy curves V(q) and VPV(q).

Next, we numerically solve the Schrödinger equation for V(q) using the Numerov–Cooley procedure as implemented by Bast and obtain the vibrational wavefunctions |n⟩.80–82 The parity violating shift of the nth vibrational level in the first order of perturbation theory is then
EnPV=n|VPV(q)|n.
(2)
The difference between the enantiomers in the frequency of a transition from level m to level n is then given by
ΔνmnPV=2h(EnPVEmPV),
(3)
with h the Planck constant. The factor 2 arises since when in one enantiomer the energy shifts up by EPV, it shifts down by the same amount in the other enantiomer.

Any candidate discussed here requires a pathway for its creation if it is to be used in a precision measurement. For some of the candidates, such as CHDBrI and CHDBr2, the non-chiral counterparts, CH2BrI and CH2Br2, are commercially available. The natural abundance between the two bromine isotopes is 1:1, leaving 50% of the molecules in the pro-chiral mixed isotope form. The molecules with different bromine isotopes can be chosen through mass selection. For the deuterated molecules, selection is also an option, but the very low natural abundance means that a synthesis method is preferable.

The Ca substituted molecules would need to be generated in the vacuum chamber. For example, it may be possible to generate CHCaBrI by creating a Ca plasma by laser ablation near a supersonic expansion seeded with CH2BrI or CHBr2I. A similar scheme83 has been used to generate CH3Ca by ablation of Ca near CH3Cl. Another similar scheme has been proposed to create Yb substituted methanes,84 which may be a pathway to generate CHYbBrI.

The search for PNC in molecules can benefit from the long interrogation times accessible in trapped chiral molecular ions as well as the enhanced PNC shifts they are predicted to exhibit. However, for a successful precision spectroscopy experiment with chiral molecular ions, a favorable candidate must also fulfill other criteria, including efficient state preparation, high quantum efficiency in detection, and resistance to predissociation when vibrationally excited. In this work, we investigate these properties for several five-atom, carbon-center, tetrahedral chiral cations via ab initio calculations and estimate the magnitude of the PNC shift for some of the candidates.

To this end, we calculate several electronic properties mainly using coupled-cluster based methods. We validate the chirality of the optimized cation geometries and calculate vertical and adiabatic ionization energies, dissociation channel energies, isotopic vibrational modes, rotational constants, transition state energies, several excitation energies of the neutral systems, and PNC frequency shifts where relevant.

Our in-depth study in search of candidates for trapped chiral-molecular-ion precision spectroscopy has revealed that CHXBrI+, where X ∈ {D, Ca}, and isotopically chiral CHX79Br81Br+, where X ∈ {D, Ca, Yb}, are favorable candidates. These candidates have promising avenues toward their preparation with internally cold temperatures and are stable in the charged form. Moreover, the magnitude of vibrational frequency shifts due to PNC is shown to be significant for selected candidates.

The supplementary material Tables S1–S6 contain benchmark calculations for IPs, optimized geometries, and dissociation energies, which demonstrate the need to use EOM-CC for calculating doublet states. Table S7 demonstrates the importance of augmenting the basis sets for excitation energy calculations. Table S8 compares vibrational modes for the neutral, achiral CH2BrI and CH2Br2 with experimental measurements and introduces vibrational scaling factors. These factors can be used to retrieve realistic cation vibrational frequencies, for which experimental values do not exist to the best of our knowledge.

We acknowledge I. Gilary for useful discussions. E.E. acknowledges Peter Schwerdtfeger for useful discussions. This research was supported by the Israel Science Foundation under Grant No. 1661/19 and the Israel Science Foundation under Grant No. 1142/21. This research project was partially supported by the Helen Diller Quantum Center at the Technion. E.E. wishes to acknowledge the Indonesia Endowment Fund for Education/Lembaga Pengelola Dana Pendidikan (LPDP) for research funding. A.B., E.E., and L.F.P. acknowledge the Center for Information Technology of the University of Groningen for their support and for providing access to the Peregrine high-performance computing cluster.

The authors have no conflicts to disclose.

Arie Landau: Conceptualization (equal); Investigation (lead); Software (lead); Writing – original draft (lead); Writing – review & editing (lead). E. Eduardus: Investigation (lead); Methodology (equal); Software (equal); Writing – review & editing (equal). Doron Behar: Formal analysis (supporting); Investigation (supporting); Writing – review & editing (supporting). Eliana Ruth Wallach: Formal analysis (supporting); Investigation (supporting); Writing – review & editing (supporting). Lukáš F. Pašteka: Investigation (equal); Methodology (equal); Supervision (equal); Writing – review & editing (equal). Shirin Faraji: Methodology (equal); Supervision (equal). Anastasia Borschevsky: Investigation (equal); Methodology (equal); Software (equal); Supervision (equal); Writing – review & editing (equal). Yuval Shagam: Conceptualization (lead); Funding acquisition (lead); Investigation (lead); Methodology (equal); Software (equal); Supervision (lead); Writing – original draft (lead); Writing – review & editing (lead).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
C.-S.
Wu
,
E.
Ambler
,
R. W.
Hayward
,
D. D.
Hoppes
, and
R. P.
Hudson
, “
Experimental test of parity conservation in beta decay
,”
Phys. Rev.
105
,
1413
(
1957
).
2.
M.
Bouchiat
,
J.
Guena
,
L.
Hunter
, and
L.
Pottier
, “
Observation of a parity violation in cesium
,”
Phys. Lett. B
117
,
358
364
(
1982
).
3.
C. S.
Wood
,
S. C.
Bennett
,
D.
Cho
,
B. P.
Masterson
,
J. L.
Roberts
,
C. E.
Tanner
, and
C. E.
Wieman
, “
Measurement of parity nonconservation and an anapole moment in cesium
,”
Science
275
,
1759
1763
(
1997
).
4.
Y.
Yamagata
, “
A hypothesis for the asymmetric appearance of biomolecules on earth
,”
J. Theor. Biol.
11
,
495
498
(
1966
).
5.
M.
Quack
,
J.
Stohner
, and
M.
Willeke
, “
High-resolution spectroscopic studies and theory of parity violation in chiral molecules
,”
Annu. Rev. Phys. Chem.
59
,
741
769
(
2008
).
6.
M.
Quack
,
G.
Seyfang
, and
G.
Wichmann
, “
Perspectives on parity violation in chiral molecules: Theory, spectroscopic experiment and biomolecular homochirality
,”
Chem. Sci.
13
,
10598
10643
(
2022
).
7.
F.
Hund
, “
Zur deutung der molekelspektren. III. Bemerkungen über das schwingungs- und rotationsspektrum bei molekeln mit mehr als zwei kernen
,”
Z. Phys.
43
,
805
826
(
1927
).
8.
V.
Letokhov
, “
On difference of energy levels of left and right molecules due to weak interactions
,”
Phys. Lett. A
53
,
275
276
(
1975
).
9.
M.
Senami
and
K.
Ito
, “
Asymmetry of electron chirality between enantiomeric pair molecules and the origin of homochirality in nature
,”
Phys. Rev. A
99
,
012509
(
2019
).
10.
M.
Quack
, “
How important is parity violation for molecular and biomolecular chirality?
,”
Angew. Chem., Int. Ed.
41
,
4618
4630
(
2002
).
11.
D. P.
Glavin
,
A. S.
Burton
,
J. E.
Elsila
,
J. C.
Aponte
, and
J. P.
Dworkin
, “
The search for chiral asymmetry as a potential biosignature in our solar system
,”
Chem. Rev.
120
,
4660
4689
(
2020
).
12.
C.
Daussy
,
T.
Marrel
,
A.
Amy-Klein
,
C. T.
Nguyen
,
C. J.
Bordé
, and
C.
Chardonnet
, “
Limit on the parity nonconserving energy difference between the enantiomers of a chiral molecule by laser spectroscopy
,”
Phys. Rev. Lett.
83
,
1554
(
1999
).
13.
L.
Satterthwaite
,
G.
Koumarianou
,
D.
Sorensen
, and
D.
Patterson
, “
Sub-Hz differential rotational spectroscopy of enantiomers
,”
Symmetry
14
,
28
(
2022
).
14.
S.
Albert
,
I.
Bolotova
,
Z.
Chen
,
C.
Fábri
,
M.
Quack
,
G.
Seyfang
, and
D.
Zindel
, “
High-resolution FTIR spectroscopy of trisulfane HSSSH: A candidate for detecting parity violation in chiral molecules
,”
Phys. Chem. Chem. Phys.
19
,
11738
11743
(
2017
).
15.
C.
Stoeffler
,
B.
Darquié
,
A.
Shelkovnikov
,
C.
Daussy
,
A.
Amy-Klein
,
C.
Chardonnet
,
L.
Guy
,
J.
Crassous
,
T. R.
Huet
,
P.
Soulard
, and
P.
Asselin
, “
High resolution spectroscopy of methyltrioxorhenium: Towards the observation of parity violation in chiral molecules
,”
Phys. Chem. Chem. Phys.
13
,
854
863
(
2011
).
16.
A.
Cournol
,
M.
Manceau
,
M.
Pierens
,
L.
Lecordier
,
D. B. A.
Tran
,
R.
Santagata
,
B.
Argence
,
A.
Goncharov
,
O.
Lopez
,
M.
Abgrall
,
Y.
Le Coq
,
R.
Le Targat
,
H.
Alvarez Martinez
,
W. K.
Lee
,
D.
Xu
,
P. E.
Pottie
,
R. J.
Hendricks
,
T. E.
Wall
,
J. M.
Bieniewska
,
B. E.
Sauer
,
M. R.
Tarbutt
,
A.
Amy-Klein
,
S. K.
Tokunaga
, and
B.
Darquié
, “
A new experiment to test parity symmetry in cold chiral molecules using vibrational spectroscopy
,”
Quantum Electron.
49
,
288
292
(
2019
).
17.
M. R.
Fiechter
,
P. A. B.
Haase
,
N.
Saleh
,
P.
Soulard
,
B.
Tremblay
,
R. W. A.
Havenith
,
R. G. E.
Timmermans
,
P.
Schwerdtfeger
,
J.
Crassous
,
B.
Darquié
,
L. F.
Pašteka
, and
A.
Borschevsky
, “
Toward detection of the molecular parity violation in chiral Ru(acac)3 and Os(acac)3
,”
J. Phys. Chem. Lett.
13
,
10011
10017
(
2022
).
18.
D.
Figgen
and
P.
Schwerdtfeger
, “
Structures, inversion barriers, and parity violation effects in chiral SeOXY molecules (X,Y=H, F, Cl, Br, or I)
,”
J. Chem. Phys.
130
,
054306
(
2009
).
19.
W. B.
Cairncross
,
D. N.
Gresh
,
M.
Grau
,
K. C.
Cossel
,
T. S.
Roussy
,
Y.
Ni
,
Y.
Zhou
,
J.
Ye
, and
E. A.
Cornell
, “
Precision measurement of the electron’s electric dipole moment using trapped molecular ions
,”
Phys. Rev. Lett.
119
,
153001
(
2017
).
20.
Y.
Zhou
,
Y.
Shagam
,
W. B.
Cairncross
,
K. B.
Ng
,
T. S.
Roussy
,
T.
Grogan
,
K.
Boyce
,
A.
Vigil
,
M.
Pettine
,
T.
Zelevinsky
,
J.
Ye
, and
E. A.
Cornell
, “
Second-scale coherence measured at the quantum projection noise limit with hundreds of molecular ions
,”
Phys. Rev. Lett.
124
,
053201
(
2020
).
21.
T. S.
Roussy
,
L.
Caldwell
,
T.
Wright
,
W. B.
Cairncross
,
Y.
Shagam
,
K. B.
Ng
,
N.
Schlossberger
,
S. Y.
Park
,
A.
Wang
,
J.
Ye
, and
E. A.
Cornell
, “
An improved bound on the electron’s electric dipole moment
,”
Science
381
,
46
50
(
2023
).
22.
D.
Hanneke
,
R. A.
Carollo
, and
D. A.
Lane
, “
High sensitivity to variation in the proton-to-electron mass ratio in O2+
,”
Phys. Rev. A
94
,
050101
(
2016
).
23.
M.
Fan
,
C. A.
Holliman
,
X.
Shi
,
H.
Zhang
,
M. W.
Straus
,
X.
Li
,
S. W.
Buechele
, and
A. M.
Jayich
, “
Optical mass spectrometry of cold RaOH+ and RaOCH3+
,”
Phys. Rev. Lett.
126
,
023002
(
2021
).
24.
T. S.
Roussy
,
D. A.
Palken
,
W. B.
Cairncross
,
B. M.
Brubaker
,
D. N.
Gresh
,
M.
Grau
,
K. C.
Cossel
,
K. B.
Ng
,
Y.
Shagam
,
Y.
Zhou
,
V. V.
Flambaum
,
K. W.
Lehnert
,
J.
Ye
, and
E. A.
Cornell
, “
Experimental constraint on axion-like particles over seven orders of magnitude in mass
,”
Phys. Rev. Lett.
126
,
171301
(
2021
).
25.
C. W.
Chou
,
A. L.
Collopy
,
C.
Kurz
,
Y.
Lin
,
M. E.
Harding
,
P. N.
Plessow
,
T.
Fortier
,
S.
Diddams
,
D.
Leibfried
, and
D. R.
Leibrandt
, “
Frequency-comb spectroscopy on pure quantum states of a single molecular ion
,”
Science
367
,
1458
1461
(
2020
).
26.
X.
Tong
,
A. H.
Winney
, and
S.
Willitsch
, “
Sympathetic cooling of molecular ions in selected rotational and vibrational states produced by threshold photoionization
,”
Phys. Rev. Lett.
105
,
143001
(
2010
).
27.
N.
Kuroda
,
T.
Oho
,
M.
Senami
, and
A.
Sunaga
, “
Enhancement of the parity-violating energy difference of H2X2 molecules by electronic excitation
,”
Phys. Rev. A
105
,
012820
(
2022
).
28.
J.
Stohner
, “
Parity violating effects in the molecular anion CBrClF
,”
Int. J. Mass Spectrom.
233
,
385
394
(
2004
).
29.
M.
Gottselig
,
M.
Quack
,
J.
Stohner
, and
M.
Willeke
, “
Mode-selective stereomutation tunneling and parity violation in HOClH+ and H2Te2 isotopomers
,”
Int. J. Mass Spectrom.
233
,
373
(
2004
).
30.
D. M.
Segal
,
V.
Lorent
,
R.
Dubessy
, and
B.
Darquié
, “
Studying fundamental physics using quantum enabled technologies with trapped molecular ions
,”
J. Mod. Opt.
65
,
490
500
(
2018
).
31.
H.
Loh
,
R. P.
Stutz
,
T. S.
Yahn
,
H.
Looser
,
R. W.
Field
,
E. A.
Cornell
, and
E. A.
Cornell
, “
REMPI spectroscopy of HfF
,”
J. Mol. Spectrosc.
276-277
,
49
56
(
2012
).
32.
J.
Schmidt
,
T.
Louvradoux
,
J.
Heinrich
,
N.
Sillitoe
,
M.
Simpson
,
J. P.
Karr
, and
L.
Hilico
, “
Trapping, cooling, and photodissociation analysis of state-selected H2+ ions produced by (3 + 1) multiphoton ionization
,”
Phys. Rev. Appl.
14
,
024053
(
2020
),
33.
T. A.
Isaev
and
R.
Berger
, “
Polyatomic candidates for cooling of molecules with lasers from simple theoretical concepts
,”
Phys. Rev. Lett.
116
,
063006
(
2016
).
34.
I.
Kozyryev
,
L.
Baum
,
K.
Matsuda
, and
J. M.
Doyle
, “
Proposal for laser cooling of complex polyatomic molecules
,”
ChemPhysChem
17
,
3641
3648
(
2016
).
35.
I.
Kozyryev
,
L.
Baum
,
K.
Matsuda
,
B. L.
Augenbraun
,
L.
Anderegg
,
A. P.
Sedlack
, and
J. M.
Doyle
, “
Sisyphus laser cooling of a polyatomic molecule
,”
Phys. Rev. Lett.
118
,
173201
(
2017
).
36.
D.
Mitra
,
Z. D.
Lasner
,
G. Z.
Zhu
,
C. E.
Dickerson
,
B. L.
Augenbraun
,
A. D.
Bailey
,
A. N.
Alexandrova
,
W. C.
Campbell
,
J. R.
Caram
,
E. R.
Hudson
, and
J. M.
Doyle
, “
Pathway toward optical cycling and laser cooling of functionalized arenes
,”
J. Phys. Chem. Lett.
13
,
7029
7035
(
2022
).
37.
I.
Erez
,
E. R.
Wallach
, and
Y.
Shagam
, “
Simultaneous enantiomer-resolved ramsey spectroscopy for chiral molecules
,” arXiv:2206.03699 (
2022
).
38.
P. B.
Changala
,
T. L.
Nguyen
,
J. H.
Baraban
,
G. B.
Ellison
,
J. F.
Stanton
,
D. H.
Bross
, and
B.
Ruscic
, “
Active thermochemical tables: The adiabatic ionization energy of hydrogen peroxide
,”
J. Phys. Chem. A
121
,
8799
8806
(
2017
).
39.
Eduardus
,
Y.
Shagam
,
A.
Landau
,
S.
Faraji
,
P.
Schwerdtfeger
,
A.
Borschevsky
, and
L. F.
Pašteka
, “
Large vibrationally induced parity violation effects in CHDBrI+—A promising candidate for future experiments
,” arXiv:2306.09763 (
2023
).
40.
M. D.
Morse
, “
Supersonic beam sources
,”
Exp. Methods Phys. Sci.
29
,
21
47
(
1996
).
41.
N. R.
Hutzler
,
H.-I.
Lu
, and
J. M.
Doyle
, “
The buffer gas beam: An intense, cold, and slow source for atoms and molecules
,”
Chem. Rev.
112
,
4803
4827
(
2012
).
42.
N. B.
Vilas
,
C.
Hallas
,
L.
Anderegg
,
P.
Robichaud
,
A.
Winnicki
,
D.
Mitra
, and
J. M.
Doyle
, “
Magneto-optical trapping and sub-Doppler cooling of a polyatomic molecule
,”
Nature
606
,
70
74
(
2022
).
43.
G.-Z.
Zhu
,
D.
Mitra
,
B. L.
Augenbraun
,
C. E.
Dickerson
,
M. J.
Frim
,
G.
Lao
,
Z. D.
Lasner
,
A. N.
Alexandrova
,
W. C.
Campbell
, and
J. R.
Caram
, “
Functionalizing aromatic compounds with optical cycling centres
,”
Nat. Chem.
14
,
995
999
(
2022
).
44.
J.
Lee
,
J.
Bischoff
,
A. O.
Hernandez-Castillo
,
B.
Sartakov
,
G.
Meijer
, and
S.
Eibenberger-Arias
, “
Quantitative study of enantiomer-specific state transfer
,”
Phys. Rev. Lett.
128
,
173001
(
2022
).
45.
M.
Leibscher
,
E.
Pozzoli
,
C.
Pérez
,
M.
Schnell
,
M.
Sigalotti
,
U.
Boscain
, and
C. P.
Koch
, “
Full quantum control of enantiomer-selective state transfer in chiral molecules despite degeneracy
,”
Commun. Phys.
5
,
110
(
2022
).
46.
K. K.
Ni
,
H.
Loh
,
M.
Grau
,
K. C.
Cossel
,
J.
Ye
, and
E. A.
Cornell
, “
State-specific detection of trapped HfF+ by photodissociation
,”
J. Mol. Spectrosc.
300
,
12
15
(
2014
).
47.
J. P.
Karr
,
A.
Douillet
, and
L.
Hilico
, “
Photodissociation of trapped
H2+
ions for REMPD spectroscopy
,”
Appl. Phys. B
107
,
1043
1052
(
2012
),
48.
R.
Carollo
,
A.
Frenett
, and
D.
Hanneke
, “
Two-photon vibrational transitions in 16O2+ as probes of variation of the proton-to-electron mass ratio
,”
Atoms
7
,
1
(
2018
),
49.
J.
Zádor
,
R. X.
Fernandes
,
Y.
Georgievskii
,
G.
Meloni
,
C. A.
Taatjes
, and
J. A.
Miller
, “
The reaction of hydroxyethyl radicals with o2: A theoretical analysis and experimental product study
,”
Proc. Combust. Inst.
32
,
271
277
(
2009
).
50.
X.-J.
Qi
,
L.
Liu
,
Y.
Fu
, and
Q.-X.
Guo
, “
Hydrogen bonding interactions of radicals
,”
Struct. Chem.
16
,
347
353
(
2005
).
51.
S.
Luo
,
L.
Gao
,
Z.
Wei
,
R.
Spinney
,
D. D.
Dionysiou
,
W.-P.
Hu
,
L.
Chai
, and
R.
Xiao
, “
Kinetic and mechanistic aspects of hydroxyl radical–mediated degradation of naproxen and reaction intermediates
,”
Water Res.
137
,
233
241
(
2018
).
52.
Y.
Xu
,
S.
Xi
,
F.
Wang
, and
X.
Li
, “
Theoretical study on reactions of alkylperoxy radicals
,”
J. Phys. Chem. A
123
,
3949
3958
(
2019
).
53.
R.
Xiao
,
L.
He
,
Z.
Luo
,
R.
Spinney
,
Z.
Wei
,
D. D.
Dionysiou
, and
F.
Zhao
, “
An experimental and theoretical study on the degradation of clonidine by hydroxyl and sulfate radicals
,”
Sci. Total Environ.
710
,
136333
(
2020
).
54.
A.
Lisovskaya
,
I.
Carmichael
, and
A.
Harriman
, “
Pulse radiolysis investigation of radicals derived from water-soluble cyanine dyes: Implications for super-resolution microscopy
,”
J. Phys. Chem. A
125
,
5779
5793
(
2021
).
55.
J. D.
Goddard
, “
A computational study of the HCCO and HCCS radicals
,”
Chem. Phys. Lett.
154
,
387
390
(
1989
).
56.
A.
Rauk
,
D.
Yu
, and
D. A.
Armstrong
, “
Carboxyl free radicals: Formyloxyl (HCOO.bul.) and acetyloxyl (CH3COO.bul.) revisited
,”
J. Am. Chem. Soc.
116
,
8222
8228
(
1994
).
57.
A. I.
Krylov
, “
Equation-of-motion coupled-cluster methods for open-shell and electronically excited species: The hitchhiker’s guide to Fock space
,”
Annu. Rev. Phys. Chem.
59
,
433
462
(
2008
).
58.
P. A.
Pieniazek
,
S. E.
Bradforth
, and
A. I.
Krylov
, “
Charge localization and Jahn–Teller distortions in the benzene dimer cation
,”
J. Chem. Phys.
129
,
074104
(
2008
).
59.
T.
Helgaker
,
P.
Jorgensen
, and
J.
Olsen
,
Molecular Electronic-Structure Theory
(
John Wiley & Sons
,
2014
).
60.
X.
Feng
,
E.
Epifanovsky
,
J.
Gauss
, and
A. I.
Krylov
, “
Implementation of analytic gradients for CCSD and EOM-CCSD using Cholesky decomposition of the electron-repulsion integrals and their derivatives: Theory and benchmarks
,”
J. Chem. Phys.
151
,
014110
(
2019
).
61.
T. H.
Dunning
, Jr
, “
Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen
,”
J. Chem. Phys.
90
,
1007
1023
(
1989
).
62.
D. E.
Woon
and
T. H.
Dunning
, Jr.
, “
Gaussian basis sets for use in correlated molecular calculations. III. The atoms aluminum through argon
,”
J. Chem. Phys.
98
,
1358
1371
(
1993
).
63.
R. A.
Kendall
,
T. H.
Dunning
, Jr
, and
R. J.
Harrison
, “
Electron affinities of the first-row atoms revisited. systematic basis sets and wave functions
,”
J. Chem. Phys.
96
,
6796
6806
(
1992
).
64.
K. A.
Peterson
, “
Systematically convergent basis sets with relativistic pseudopotentials. I. Correlation consistent basis sets for the post-d group 13–15 elements
,”
J. Chem. Phys.
119
,
11099
11112
(
2003
).
65.
K. A.
Peterson
,
D.
Figgen
,
E.
Goll
,
H.
Stoll
, and
M.
Dolg
, “
Systematically convergent basis sets with relativistic pseudopotentials. II. Small-core pseudopotentials and correlation consistent basis sets for the post-d group 16–18 elements
,”
J. Chem. Phys.
119
,
11113
11123
(
2003
).
66.
R.
Gulde
,
P.
Pollak
, and
F.
Weigend
, “
Error-balanced segmented contracted basis sets of double-ζ to quadruple-ζ valence quality for the lanthanides
,”
J. Chem. Theory Comput.
8
,
4062
4068
(
2012
).
67.
K. G.
Dyall
, “
An exact separation of the spin-free and spin-dependent terms of the Dirac–Coulomb–Breit Hamiltonian
,”
J. Chem. Phys.
100
,
2118
2127
(
1994
).
68.
T.
Saue
,
R.
Bast
,
A. S. P.
Gomes
,
H. J. A.
Jensen
,
L.
Visscher
,
I. A.
Aucar
,
R.
Di Remigio
,
K. G.
Dyall
,
E.
Eliav
, and
E.
Fasshauer
, “
The Dirac code for relativistic molecular calculations
,”
J. Chem. Phys.
152
,
204104
(
2020
).
69.
F. C.
Roberts
and
J. H.
Lehman
, “
Infrared frequency comb spectroscopy of CH2I2: Influence of hot bands and pressure broadening on the ν1 and ν6 fundamental transitions
,”
J. Chem. Phys.
156
,
114301
(
2022
).
70.
K. K.
Irikura
,
R. D.
Johnson
, and
R. N.
Kacker
, “
Uncertainties in scaling factors for ab initio vibrational frequencies
,”
J. Phys. Chem. A
109
,
8430
8437
(
2005
).
71.
NIST
, “
Computational chemistry comparison and benchmark database
,” https://cccbdb.nist.gov/vibscalejustx.asp,
2022
.
72.
S. J.
Schowalter
,
K.
Chen
,
W. G.
Rellergert
,
S. T.
Sullivan
, and
E. R.
Hudson
, “
An integrated ion trap and time-of-flight mass spectrometer for chemical and photo- reaction dynamics studies
,”
Rev. Sci. Instrum.
83
,
043103
(
2012
).
73.
P. C.
Schmid
,
J.
Greenberg
,
M. I.
Miller
,
K.
Loeffler
, and
H. J.
Lewandowski
, “
High resolution ion trap time-of-flight mass spectrometer for cold trapped ion experiments
,”
Rev. Sci. Instrum.
88
,
123107
(
2017
).
74.
D. J.
Nesbitt
and
R. W.
Field
, “
Vibrational energy flow in highly excited molecules: Role of intramolecular vibrational redistribution
,”
J. Phys. Chem.
100
,
12735
12756
(
1996
).
75.
E.
Epifanovsky
,
A. T.
Gilbert
,
X.
Feng
,
J.
Lee
,
Y.
Mao
,
N.
Mardirossian
,
P.
Pokhilko
,
A. F.
White
,
M. P.
Coons
,
A. L.
Dempwolff
,
Z.
Gan
,
D.
Hait
,
P. R.
Horn
,
L. D.
Jacobson
,
I.
Kaliman
,
J.
Kussmann
,
A. W.
Lange
,
K. U.
Lao
,
D. S.
Levine
,
J.
Liu
,
S. C.
McKenzie
,
A. F.
Morrison
,
K. D.
Nanda
,
F.
Plasser
,
D. R.
Rehn
,
M. L.
Vidal
,
Z. Q.
You
,
Y.
Zhu
,
B.
Alam
,
B. J.
Albrecht
,
A.
Aldossary
,
E.
Alguire
,
J. H.
Andersen
,
V.
Athavale
,
D.
Barton
,
K.
Begam
,
A.
Behn
,
N.
Bellonzi
,
Y. A.
Bernard
,
E. J.
Berquist
,
H. G.
Burton
,
A.
Carreras
,
K.
Carter-Fenk
,
R.
Chakraborty
,
A. D.
Chien
,
K. D.
Closser
,
V.
Cofer-Shabica
,
S.
Dasgupta
,
M.
de Wergifosse
,
J.
Deng
,
M.
Diedenhofen
,
H.
Do
,
S.
Ehlert
,
P. T.
Fang
,
S.
Fatehi
,
Q.
Feng
,
T.
Friedhoff
,
J.
Gayvert
,
Q.
Ge
,
G.
Gidofalvi
,
M.
Goldey
,
J.
Gomes
,
C. E.
González-Espinoza
,
S.
Gulania
,
A. O.
Gunina
,
M. W.
Hanson-Heine
,
P. H.
Harbach
,
A.
Hauser
,
M. F.
Herbst
,
M.
Hernández Vera
,
M.
Hodecker
,
Z. C.
Holden
,
S.
Houck
,
X.
Huang
,
K.
Hui
,
B. C.
Huynh
,
M.
Ivanov
,
Á.
Jász
,
H.
Ji
,
H.
Jiang
,
B.
Kaduk
,
S.
Kähler
,
K.
Khistyaev
,
J.
Kim
,
G.
Kis
,
P.
Klunzinger
,
Z.
Koczor-Benda
,
J. H.
Koh
,
D.
Kosenkov
,
L.
Koulias
,
T.
Kowalczyk
,
C. M.
Krauter
,
K.
Kue
,
A.
Kunitsa
,
T.
Kus
,
I.
Ladjánszki
,
A.
Landau
,
K. V.
Lawler
,
D.
Lefrancois
,
S.
Lehtola
,
R. R.
Li
,
Y. P.
Li
,
J.
Liang
,
M.
Liebenthal
,
H. H.
Lin
,
Y. S.
Lin
,
F.
Liu
,
K. Y.
Liu
,
M.
Loipersberger
,
A.
Luenser
,
A.
Manjanath
,
P.
Manohar
,
E.
Mansoor
,
S. F.
Manzer
,
S. P.
Mao
,
A. V.
Marenich
,
T.
Markovich
,
S.
Mason
,
S. A.
Maurer
,
P. F.
McLaughlin
,
M. F.
Menger
,
J. M.
Mewes
,
S. A.
Mewes
,
P.
Morgante
,
J. W.
Mullinax
,
K. J.
Oosterbaan
,
G.
Paran
,
A. C.
Paul
,
S. K.
Paul
,
F.
Pavošević
,
Z.
Pei
,
S.
Prager
,
E. I.
Proynov
,
Á.
Rák
,
E.
Ramos-Cordoba
,
B.
Rana
,
A. E.
Rask
,
A.
Rettig
,
R. M.
Richard
,
F.
Rob
,
E.
Rossomme
,
T.
Scheele
,
M.
Scheurer
,
M.
Schneider
,
N.
Sergueev
,
S. M.
Sharada
,
W.
Skomorowski
,
D. W.
Small
,
C. J.
Stein
,
Y. C.
Su
,
E. J.
Sundstrom
,
Z.
Tao
,
J.
Thirman
,
G. J.
Tornai
,
T.
Tsuchimochi
,
N. M.
Tubman
,
S. P.
Veccham
,
O.
Vydrov
,
J.
Wenzel
,
J.
Witte
,
A.
Yamada
,
K.
Yao
,
S.
Yeganeh
,
S. R.
Yost
,
A.
Zech
,
I. Y.
Zhang
,
X.
Zhang
,
Y.
Zhang
,
D.
Zuev
,
A.
Aspuru-Guzik
,
A. T.
Bell
,
N. A.
Besley
,
K. B.
Bravaya
,
B. R.
Brooks
,
D.
Casanova
,
J. D.
Chai
,
S.
Coriani
,
C. J.
Cramer
,
G.
Cserey
,
A. E.
Deprince
,
R. A.
Distasio
,
A.
Dreuw
,
B. D.
Dunietz
,
T. R.
Furlani
,
W. A.
Goddard
,
S.
Hammes-Schiffer
,
T.
Head-Gordon
,
W. J.
Hehre
,
C. P.
Hsu
,
T. C.
Jagau
,
Y.
Jung
,
A.
Klamt
,
J.
Kong
,
D. S.
Lambrecht
,
W.
Liang
,
N. J.
Mayhall
,
C. W.
McCurdy
,
J. B.
Neaton
,
C.
Ochsenfeld
,
J. A.
Parkhill
,
R.
Peverati
,
V. A.
Rassolov
,
Y.
Shao
,
L. V.
Slipchenko
,
T.
Stauch
,
R. P.
Steele
,
J. E.
Subotnik
,
A. J.
Thom
,
A.
Tkatchenko
,
D. G.
Truhlar
,
T.
Van Voorhis
,
T. A.
Wesolowski
,
K. B.
Whaley
,
H. L.
Woodcock
,
P. M.
Zimmerman
,
S.
Faraji
,
P. M.
Gill
,
M.
Head-Gordon
,
J. M.
Herbert
, and
A. I.
Krylov
, “
Software for the frontiers of quantum chemistry: An overview of developments in the Q-Chem 5 package
,”
J. Chem. Phys.
155
,
084801
(
2021
).
76.
DIRAC
,
A relativistic ab initio electronic structure program, Release DIRAC23
,
2023
, written by
R.
numerov
,
A. S. P.
Gomes
,
T.
Saue
and
L.
Visscher
and
H. J. Aa.
Jensen
, with contributions from
I.
A. Aucar
,
V.
Bakken
,
C.
Chibueze
,
J.
Creutzberg
,
K. G.
Dyall
,
S.
Dubillard
,
U.
Ekström
,
E.
Eliav
,
T.
Enevoldsen
,
E.
Faßhauer
,
T.
Fleig
,
O.
Fossgaard
,
L.
Halbert
,
E. D.
Hedegård
,
T.
Helgaker
,
B.
Helmich–Paris
,
J.
Henriksson
,
M.
van Horn
,
M.
Iliaš
,
Ch. R.
Jacob
,
S.
Knecht
,
S.
Komorovský
,
O.
Kullie
,
J. K.
Lærdahl
,
C. V.
Larsen
,
Y. S.
Lee
,
N. H.
List
,
H. S.
Nataraj
,
M. K.
Nayak
,
P.
Norman
,
A.
Nyvang
,
G.
Olejniczak
,
J.
Olsen
,
J. M. H.
Olsen
,
A.
Papadopoulos
,
Y. C.
Park
,
J. K.
Pedersen
,
M.
Pernpointner
,
J. V.
Pototschnig
,
R.
di Remigio
,
M.
Repisky
,
K.
Ruud
,
P.
Sałek
,
B.
Schimmelpfennig
,
B.
Senjean
,
A.
Shee
,
J.
Sikkema
,
A.
Sunaga
,
A. J.
Thorvaldsen
,
J.
Thyssen
,
J.
van Stralen
,
M. L.
Vidal
,
S.
Villaume
,
O.
Visser
,
T.
Winther
,
S.
Yamamoto
, and
X.
Yuan
, available at https://doi.org/10.5281/zenodo.7670749, see also https://www.diracprogram.org.
77.
B.
Schimmelpfennig
, “
AMFI, an atomic mean-field spin-orbit integral program
,” 1996 and 1999, Bernd Schimmelpfennig, University of Stockholm.
78.
C.
Thierfelder
,
G.
Rauhut
, and
P.
Schwerdtfeger
, “
Relativistic coupled-cluster study of the parity-violation energy shift of ChFClBr
,”
Phys. Rev. A
81
,
032513
(
2010
).
79.
K. G.
Dyall
, “
Relativistic quadruple-zeta and revised triple-zeta and double-zeta basis sets for the 4p, 5p, and 6p elements
,”
Theor. Chem. Acc.
115
,
441
447
(
2006
).
80.
B. V.
Noumerov
, “
A method of extrapolation of perturbations
,”
Mon. Not. R. Astron. Soc.
84
,
592
602
(
1924
).
81.
J. W.
Cooley
, “
An improved eigenvalue corrector formula for solving the schrodinger equation for central fields
,”
Math. Comput.
15
,
363
(
1961
).
82.
R.
Bast
, bast/numerov v0.5.0, https://doi.org/10.5281/zenodo.1000406,
2017
.
83.
C. R.
Brazier
and
P. F.
Bernath
, “
Observation of gas phase organometallic free radicals: Monomethyl derivatives of calcium and strontium
,”
J. Chem. Phys.
86
,
5918
5922
(
1987
).
84.
Y.
Chamorro
,
A.
Borschevsky
,
E.
Eliav
,
N. R.
Hutzler
,
S.
Hoekstra
, and
L. F.
Pašteka
, “
Molecular enhancement factors for the P, T-violating electric dipole moment of the electron in BaCh3 and YbCh3 symmetric top molecules
,”
Phys. Rev. A
106
,
052811
(
2022
).

Supplementary Material