The odd isotopologues of ytterbium monohydroxide, 171,173YbOH, have been identified as promising molecules to measure parity (P) and time reversal (T) violating physics. Here, we characterize the Ã2Π1/2(0,0,0)X̃2Σ+(0,0,0) band near 577 nm for these odd isotopologues. Both laser-induced fluorescence excitation spectra of a supersonic molecular beam sample and absorption spectra of a cryogenic buffer-gas cooled sample were recorded. In addition, a novel spectroscopic technique based on laser-enhanced chemical reactions is demonstrated and used in absorption measurements. This technique is especially powerful for disentangling congested spectra. An effective Hamiltonian model is used to extract the fine and hyperfine parameters for the Ã2Π1/2(0,0,0) and X̃2Σ+(0,0,0) states. A comparison of the determined X̃2Σ+(0,0,0) hyperfine parameters with recently predicted values [Denis et al., J. Chem. Phys. 152, 084303 (2020); K. Gaul and R. Berger, Phys. Rev. A 101, 012508 (2020); and Liu et al., J. Chem. Phys. 154,064110 (2021)] is made. The measured hyperfine parameters provide experimental confirmation of the computational methods used to compute the P,T-violating coupling constants Wd and WM, which correlate P,T-violating physics to P,T-violating energy shifts in the molecule. The dependence of the fine and hyperfine parameters of the Ã2Π1/2(0,0,0) and X̃2Σ+(0,0,0) states for all isotopologues of YbOH are discussed, and a comparison to isoelectronic YbF is made.

Polar molecules have emerged as ideal systems to measure symmetry-violating physics in low energy, table top experiments.1–4 The large internal fields and closely spaced opposite parity states present in molecules provide extreme sensitivity to both time reversal (T) violating and parity (P) violating physics.5,6 With the enhanced coherence times resulting from laser cooling, precision measurements with heavy polar molecules have the potential to probe beyond standard model (BSM) physics at PeV scales.3,7,8 In addition, P-violating effects are amplified in molecules when opposite parity states, such as neighboring rotational states, are tuned to near degeneracy via an external field.9–11 Searches for P-violation in molecules can provide both precision tests of the standard model (SM) via measurement of nuclear spin-dependent parity violation (NSD-PV),12,13 and sensitive probes of P-violating BSM physics.14 The odd isotopes of the linear triatomic molecule ytterbium monohydroxide, 171YbOH and 173YbOH, have been identified as promising candidates for next generation precision measurements of P- and T-violating physics. The 171YbOH isotopologue provides sensitivity to NSD-PV,15 while the 173YbOH isotopologue is sensitive to new T,P-violating hadronic BSM physics via a measurement of the 173Yb nuclear magnetic quadrupole moment (NMQM).7,16,17 These molecules combine advantageous parity doublets, which are absent in their diatomic analogs, with the ability to be laser cooled.7,18

NSD-PV arises from three major sources: vector electron-axial nucleon electroweak coupling (VeAn), the nuclear anapole moment, and the combination of nuclear electroweak charge and normal hyperfine structure. The P-odd, effective NSD-PV Hamiltonian for a single molecular electronic state is10,15

(1)

where κ is the measurable NSD-PV parameter encapsulating the effects of all NSD-PV sources, WP is an effective parameter quantifying the overlap of the valence electron’s wavefunction with the nucleus, S is the valence electron spin, Î=II, where I is the nuclear spin of the 171Yb nucleus, and n̂ is a unit vector along the molecular symmetry axis. Currently, the only non-zero measurement of NSD-PV is in atomic Cs.19 

The coupling of new T, P-violating BSM physics to standard model particles will result in T, P-violating energy shifts in atoms and molecules.1,3,4 The T, P-odd effective molecular Hamiltonian describing these energy shifts is20 

(2)

where kS is the T, P-odd scalar-pseudoscalar nucleon-electron current interaction constant,21,de is the electron electric dipole moment (eEDM), Q is the nuclear Schiff moment (NSM), M is the NMQM, WS, Wd, WQ, and WM are coupling constants parameterizing the sensitivity of the molecule to the different T,P-violating sources, and T̂ is a rank 2 tensor that relates the NMQM shift to the nuclear spin orientation. Note the NMQM term is only nonzero for nuclei with I>12. A measurement of a non-zero eEDM, NSM, or NMQM within the projected sensitivity of contemporary experiments would be a confirmation of new T,P-violating BSM physics, while a null measurement places bounds that constrain potential BSM theories. The current limit on the eEDM5 is de<1.1×1029ecm, and several upcoming molecular experiments aim to improve this limit by an order of magnitude or more.8,18,22–24 The most stringent limit on an NSM results from measurements involving 199Hg.25 Several other NSM experiments in other atomic26–29 and molecular30,31 systems are also under way. The best bounds on an NMQM are obtained from measurements of atomic Cs.32 NMQMs are enhanced by collective effects in deformed nuclei,20,33 and the large quadrupole deformation of the 173Yb nucleus indicates that it should have an enhanced NMQM. A measurement of an NMQM in 173YbOH would complement eEDM searches as the former probes BSM physics in the hadronic sector, as opposed to the leptonic sector.20,34,35 Furthermore, an NMQM measurement would complement NSM searches since a measurement of both an NMQM and an NSM in several different systems will allow the exact source of the hadronic BSM physics to be pinpointed.

Polyatomic molecules offer several distinct advantages over diatomic molecules for precision measurements.4,7 First, linear triatomic molecules generically have parity doublets, known as l-doublets, in their excited bending modes. In YbOH and other similar molecules, the l-doublet splitting is ∼10 MHz.7 In the case of T,P-violation measurements, l-doublets enable full polarization in modest electric fields, <1 kV/cm7, and act as internal comagnetometer states, allowing for robust control over systematic errors via reversal of the T,P-violating energy shift without switching external laboratory fields. In the case of an NSD-PV measurement, the opposite parity l-doublet states can be tuned to near degeneracy with modest magnetic fields, ∼1 to 10 mT,15 both increasing sensitivity and reducing the systematics associated with large B fields and field reversals. The electronic structure of the molecule can then be leveraged for laser cooling (if the molecule has an electronic structure amendable to laser cooling, such as YbOH) without conflicting with the internal structure needed for parity doublets. Precision measurements utilizing laser cooled and trapped molecules could increase measurement sensitivity by orders of magnitude compared to beam experiments.7 Laser cooling of the 174YbOH isotopologue has already been demonstrated18 and could be extended to the odd isotopologues by taking hyperfine structure into consideration.

Here, we report on the characterization of the Ã2Π1/2(0,0,0)X̃2Σ1/2(0,0,0) transition of 171YbOH and 173YbOH, which for convenience will be designated as 000Ã2Π1/2-X̃2Σ+. Our study includes both excitation spectroscopy using laser induced fluorescence (LIF) detection of a cold supersonic molecular beam (MB) sample and laser absorption measurements of a cryogenic buffer-gas cooled (CBGC) sample. The derived molecular parameters are necessary for the implementation of both NSD-PV and NMQM measurements as well as for laser cooling. In order to extract the NSD-PV and NMQM parameters, κ and M, from P- or T-violating molecular energy shifts, the values of the coupling constants WP or WM must be known. These coupling constants cannot be measured experimentally and instead must be calculated via ab initio or semi-empirical methods. Comparison of measured hyperfine parameters with calculated ones provides a rigorous test of the computational methods used to calculate the T,P-violating coupling constants as both sets of parameters probe the nature of the valence electron in the vicinity of the nucleus. Finally, we demonstrate a novel spectroscopic technique utilizing laser-induced chemical reactions36 to both amplify the signal from the desired isotopologue and disentangle the respective isotopologue’s spectrum from that of other, more abundant, overlapping isotopologues.

A review of experimental measurements and theoretical predictions for YbOH can be found in the article describing high-resolution optical analysis performed on the even isotopologues.37 More recent experimental work includes the aforementioned demonstration of Sisyphus and Doppler laser cooling of 174YbOH18 as well as the determination of fluorescence branching ratios, radiative lifetimes, and transition moments.38 Recently reported theoretical studies include calculations of the molecular NMQM sensitivity coefficient, WM, of Eq. (2).16,17 As part of these studies, the ground state magnetic hyperfine parameters, A|| for 171YbOH and 173YbOH, and the nuclear electric quadrupole coupling parameter, e2q0Q for 173YbOH, were predicted.

Initial experiments were performed at ASU using a molecular beam LIF spectrometer similar to that used in previous high-resolution optical studies of 172,174YbOH.37,39 In brief, YbOH is produced by laser ablating (532 nm, ∼10 mJ/pulse, 20 Hz) a Yb rod in the presence of a methanol/argon supersonic expansion. The resulting beam is skimmed to produce a well collimated beam with a temporal pulse width of ∼40 µs in the detection region. The molecular beam is probed by an unfocused (∼5 mm), low power (∼5 mW), single frequency cw-dye laser ∼0.5 m downstream. It is estimated that the laser probes approximately 1 × 109 YbOH molecules in each molecular beam pulse. The resulting on-resonance LIF signal was viewed through a 580 ± 10 nm bandpass filter, detected by a photomultiplier tube (PMT), and processed using gated photon counting. Typically, the photon counts from 35 ablation pulses at each excitation laser frequency are summed. The absolute excitation wavelength is determined by co-recording a sub-Doppler I2 spectrum,40 and the relative wavelength is measured by co-recording the transmission of an actively stabilized etalon (free spectral range of 751.393 MHz).

Subsequent high-resolution spectra were recorded at Caltech by absorption spectroscopy. The apparatus was nearly identical to that used in the initial demonstration of laser-induced chemical enhancement of YbOH production.36 In summary, cold YbOH molecules are produced via cryogenic buffer-gas cooling, which is described in detail elsewhere.41–45 The molecules are produced inside a copper, cryogenic buffer-gas cell cooled to ∼4 K, with an internal cylindrical bore of 12.7 mm and a length of ∼83 mm. Helium buffer-gas is introduced into the cell by a 3.2 mm diameter copper tube. The helium then passes through a diffuser 3.2 mm from the gas inlet and exits the cell through a 5 mm aperture on the other end of the cell. All measurements presented here were performed inside the buffer-gas cell, as opposed to in the extracted beam. YbOH molecules are created by ablating (532 nm, ∼30 mJ/pulse, ∼5 Hz) solid pressed targets of Yb powder mixed with either Yb(OH)3 or Te(OH)6 powders. To make the Yb + Yb(OH)3 targets, the powders were mixed to obtain a 1:1 stoichiometric ratio of Yb to OH, then ground in a mortar and pestle, passed through a 230 mesh sieve, mixed with a 4% polyethylene glycol binder (PEG8000) by weight, and pressed in an 8 mm die at ∼10 MPa for ∼15 min. A similar procedure was used to produce the Yb + Te(OH)6 targets.

To measure the absorption spectra, three cw-laser beams are passed through the spectroscopy window: the primary tunable absorption spectroscopy beam (1 mm diameter, ∼30 μW), the normalization laser (1 mm diameter, ∼40 μW) used to monitor the shot-to-shot fluctuations in YbOH production, and the chemical enhancement laser (3 mm diameter, ∼300 mW), which increases the molecular yield by exciting atomic Yb to the metastable 3P1 state, thereby boosting chemical reactions rates between Yb atoms and OH-containing ablation products.36 

The normalization laser is fixed to either the OP12(2) or the RR11(2) line of the 000Ã2Π1/2-X̃2Σ+ transition of 172YbOH at 17 322.1732 cm−1 and 17 327.0747 cm−1, respectively.37 The frequency of the enhancement laser is fixed to the 3P11S0 transition of the desired Yb isotope. Specifically, the 174Yb(17 992.0003 cm−1), 176Yb(17 991.9685 cm−1), F″ = 1/2 → F′ = 1/2 171Yb (17 991.9292 cm−1), and the F″ = 5/2 → F′ = 7/2 173Yb (17 991.9207 cm−1) transitions46 were used. The fixed lasers are locked to a stabilized HeNe laser via a scanning transfer cavity and active feedback (∼5 MHz resolution). The primary absorption laser is continuously scanned in frequency, and the resulting absorption is detected with a photodiode. The absolute frequency of the primary absorption laser is monitored using a digital wavemeter, and the relative frequency is tracked in a separate transfer cavity with respect to a stabilized HeNe laser (∼7 MHz resolution). The absolute transition frequencies are calibrated using known 172YbOH and 174YbOH spectral lines.37 The light of the enhancement laser is switched on and off using a mechanical shutter so that both enhanced and unenhanced spectra of the desired isotopologue can be measured in successive shots.

The chemical enhancement is utilized as a spectroscopic tool to disentangle the complicated isotopologue structure. This technique is illustrated for 171YbOH in Fig. 1 and described below. The measured optical depth (OD) of the primary probe beam is integrated over the duration of the molecular pulse, typically ∼3 ms. This integrated OD is then normalized by the integrated OD of the normalization probe. The normalized signal at a single laser frequency without enhancement light is SUE = S171 + SB, where S171 is the measured integrated OD for a specific transition of 171YbOH and SB is the background integrated OD from all other overlapping YbOH isotopologues or other molecules. The measured normalized and integrated OD with the enhancement light on and tuned to the 171Yb atomic resonance is SE = ES171 + SB. Here, E is the enhancement factor, which is defined as NE/N0, the ratio of the number of molecules produced with the enhancement laser on, NE, to the number produced with the enhancement laser off, N0. Taking the difference between the enhanced and unenhanced signals SESUE=E1S171 results in the spectrum from only the 171YbOH isotopologue. It is observed that E in the cell is typically ∼4 to 8. This technique generalizes to the other isotopologues of YbOH and other molecules as well.

FIG. 1.

Extraction of the 171YbOH spectrum in the region of the even isotopologue RR11(2) lines using chemical enhancement. These data are taken from a cryogenic buffer-gas cooled sample. (a) The spectrum with no chemical enhancement. The lines of each isotopologue are indicated. The 174YbOH RR11(2) line is overlapped with a much weaker 171YbOH line. (b) The spectrum with 171YbOH chemical enhancement. (c) Difference in the enhanced and non-enhanced spectrum (a-b) this spectrum is purely from the 171YbOH isotopologue.

FIG. 1.

Extraction of the 171YbOH spectrum in the region of the even isotopologue RR11(2) lines using chemical enhancement. These data are taken from a cryogenic buffer-gas cooled sample. (a) The spectrum with no chemical enhancement. The lines of each isotopologue are indicated. The 174YbOH RR11(2) line is overlapped with a much weaker 171YbOH line. (b) The spectrum with 171YbOH chemical enhancement. (c) Difference in the enhanced and non-enhanced spectrum (a-b) this spectrum is purely from the 171YbOH isotopologue.

Close modal

A general description of the 000Ã2Π1/2-X̃2Σ+ band can be found in the previous report on the analysis of the field-free, Stark and Zeeman spectroscopy of 172,174YbOH.35 Here, we focus on the odd isotopologues. There are seven naturally occurring isotopes of Yb in moderate abundance: 168Yb (0.1%), 170Yb (3.0%), 171Yb (14.3%), 172Yb (21.8%), 173Yb (16.1%), 174Yb (31.8%), and 176Yb (12.8%). The 171YbOH and 173YbOH spectra are more complex than those for 172,174YbOH, primarily due to the large 171Yb(I = 1/2, μ = +0.493 67 μN) and 173Yb(I = 5/2, μ = −0.679 89 μN) magnetic hyperfine interaction and, in the case of 173YbOH, the large nuclear electric quadrupole hyperfine interaction (Q = 280.0 ± 4.0 fm2).47 The previously described branch designation used for the even and odd isotopologues of the (0,0)A2Π1/2-X2Σ+ band of YbF48 will also be used for the 000Ã2Π1/2-X̃2Σ+ band YbOH. The even isotopologues exhibit six branches, labeled PP11, QQ11, RR11, PQ12, OP12, and QR12, following the ΔNΔJFi'Fi″(N″) designation expected for a 2Π1/2(Hund’s case (a))- 2Σ(Hund’s case (b)) band. The low rotational levels of the X̃2Σ+ state for the odd isotopologues exhibit a Hund’s case (bβS) energy level pattern due to the large 171Yb(I = 1/2) and 173Yb(I = 5/2) magnetic hyperfine interaction, where the electron spin angular momentum, S, is coupled to the 171Yb(I = 1/2) or 173Yb(I = 5/2) nuclear spin angular momentum, I, to give an approximately good intermediate quantum number G. The angular momentum G is coupled to the rotational angular momentum, N, to give the intermediate angular momentum F1, which is then coupled to the proton nuclear spin angular momentum I2(=1/2) to give the total angular momentum F. The corresponding coupling limit wavefunction, SI(Yb)G(GN)F1F1I(H)F, is useful for describing the low-rotational energy levels of the X̃2Σ+(0,0,0) state. The Yb and H hyperfine interactions in the Ã2Π1/2 state are very small compared to the rotational and Λ-doubling spacing, and the energy level pattern is that of a molecule near a sequentially coupled Hund’s case (aβJ) limit. The corresponding coupling limit wavefunction, ηΛSΣJΩJI(Yb)F1F1I(H)F, is useful in describing the low-rotational energy levels of the Ã2Π1/2(0,0,0) state. With the exception of broadening of the lowest rotational branch features of the molecular beam spectra, there was no evidence of proton hyperfine splitting. The six 2Π1/2(Hund’s case (a)) - 2Σ(Hund’s case (b)) branches of the even isotopologues split and regroup into eight branches labeled OP1G, PP1G + PQ1G, QQ1G + QR1G, and RR1G, appropriate for a2Π1/2(case (aβJ))-2Σ(case (bβS)) band with G = 0 and 1 for 171YbOH and G = 2 and 3 for 173YbOH. The lines of the OP1G and RR1G branches (odd isotopologues) and the OP12 and RR11 branches (even isotopologues) form progressions in N″, with adjacent members separated by ∼4B″ extending to red and blue, respectively, and are relatively unblended. The 000Ã2Π1/2-X̃2Σ+ band exhibits a blue degraded head formed by low-N″ features of the PQ12 and PP11 (even isotopologues) and PP1G + PQ1G (odd isotopologues) branches. Isotopic spectral shifts are very small because the potential energy surfaces for the X̃2Σ+(0,0,0) and Ã2Π1/2(0,0,0) states are very similar, causing the PQ12, PP11, PP11, QQ11, and QR12 branches of the even isotopologues and PP1G + PQ1G and QQ1G + QR1G branches of the odd isotopes to be severely overlapped.

The observed and calculated LIF molecular beam spectra in the region of the OP12(3) even isotopologues branch features and the OP1G(3) odd isotopologues branch features are presented in Fig. 2. The observed spectrum in Fig. 2 is similar to that of the (0,0)A2Π1/2-X2Σ+ band of YbF (Ref. 48) with the exception that a small splitting (∼50 MHz) due to 19F(I = 1/2) was observed in YbF whereas the smaller H(I = 1/2) doubling in YbOH is not fully resolved. The predicted spectrum was obtained, assuming a 15 K rotational temperature, 30 MHz full width at half maximum (FWHM) Lorentzian lineshape, and optimized parameters (see below). The two components of the OP1G(3) transition are widely spaced and of the opposite order because the spacing between the X2Σ+(0,0,0) G = 3 and G = 2 groups of levels of 173YbOH is ∼−5660 MHz (∼3bF) and that between the G = 1 and G = 0 levels of 171YbOH is ∼+6750 MHz (∼bF). The spectral patterns of the G = 3 and G = 2 groups of 173YbOH are irregular because of the large nuclear electric quadrupole interaction (e2q0Q) in both the X̃2Σ+(0,0,0) and Ã2Π1/2(0,0,0) states.

FIG. 2.

Observed and calculated LIF spectra in the region of the OP12 (3) even isotopologues branch features and the OP1G (3) odd isotopologues of the 000A2Π1/2-X2Σ+ band of YbOH. The predicted spectra for the even isotopologues were obtained using the optimized parameters of Ref. 35, and those for 171YbOH and 173YbOH were obtained using optimized parameters given in Table I. A rotational temperature of 15 K and a Lorentzian full width at half maximum (FWHM) linewidth of 30 MHz were used. The features marked “*” are unidentified.

FIG. 2.

Observed and calculated LIF spectra in the region of the OP12 (3) even isotopologues branch features and the OP1G (3) odd isotopologues of the 000A2Π1/2-X2Σ+ band of YbOH. The predicted spectra for the even isotopologues were obtained using the optimized parameters of Ref. 35, and those for 171YbOH and 173YbOH were obtained using optimized parameters given in Table I. A rotational temperature of 15 K and a Lorentzian full width at half maximum (FWHM) linewidth of 30 MHz were used. The features marked “*” are unidentified.

Close modal

The laser-induced fluorescence spectrum of a molecular beam sample in the highly congested bandhead region (i.e., between 17323.50 and 17323.85 cm−1) is presented in Fig. 1 of Ref. 37. The bandhead region is dominated by low-rotational PP11, QQ11, PQ12, and QR12 branch features of the more abundant even isotopologues, which makes assignment of the PP1G + PQ1G and QQ1G + QR1G branch features of the less abundant odd isotopologues of the molecular beam sample extremely difficult. Recording the chemically enhanced absorption spectrum of a CBGC sample in the bandhead region was critical to the assignment and subsequent analysis. The observed and predicted high-resolution absorption spectra in the bandhead region of a CBGC sample are presented in Fig. 3, both with [panel (b)] and without [panel (a)] 173YbOH chemical enhancement induced by the 173Yb atomic excitation. There is approximately a factor of four increase in the signal upon atomic excitation. The difference with and without chemical enhancement [panels (a) and (b)] is presented in panel (c) of Fig. 3. The predicted spectrum was obtained assuming a 5 K rotational temperature, 90 MHz FWHM Lorentzian lineshape, and optimized parameters (see below). There is no evidence of proton hyperfine splitting at this spectral resolution. The CBGC spectra are greatly simplified relative to the supersonic molecular beam spectra due to both the isotopic selectivity, which is evident from Fig. 3 [panel (a) vs (c)], and the lower rotational temperature (∼5 vs ∼20 K). Even so, the spectral features shown in Fig. 3 [panel (c)] are a blend of many transitions (see below). For example, the feature marked “A” in Fig. 3 is predominantly a blend of the F1=3F1=3 component of the PP12 + PQ12(1) branches and the F1=4F1=4 component of the PP12 + PQ12(2) branch features of 173YbOH. The feature marked “B” is predominantly a blend of the PP12 + PQ12(3) (F1=5F1=5), PP12 + PQ12(1) (F1=2F1=3), and PP12 + PQ12(3) (F1=4F1=4) transitions.

FIG. 3.

High-resolution absorption spectra in the bandhead region of the 000A2Π1/2-X2Σ+ band of YbOH of a cryogenic buffer-gas cooled sample: (a) the spectrum with no chemical enhancement; (b) the spectrum recorded when the F″ = 5/2 → F′ = 7/2 component of the 3P11S0 transition of 173Yb (17 991.9207 cm−1) is excited; (c) difference in enhanced and unenhanced spectra (a) and (b); (d) prediction of the 173YbOH absorption spectrum using the optimized parameters given in Table I, a rotational temperature of 5 K and FWHM linewidth of 90 MHz. Units of d y-axis are arbitrary.

FIG. 3.

High-resolution absorption spectra in the bandhead region of the 000A2Π1/2-X2Σ+ band of YbOH of a cryogenic buffer-gas cooled sample: (a) the spectrum with no chemical enhancement; (b) the spectrum recorded when the F″ = 5/2 → F′ = 7/2 component of the 3P11S0 transition of 173Yb (17 991.9207 cm−1) is excited; (c) difference in enhanced and unenhanced spectra (a) and (b); (d) prediction of the 173YbOH absorption spectrum using the optimized parameters given in Table I, a rotational temperature of 5 K and FWHM linewidth of 90 MHz. Units of d y-axis are arbitrary.

Close modal

The observed and predicted high-resolution absorption spectra in the bandhead region of the CBGC sample are presented in Fig. 4, both with [panel (b)] and without [panel (a)] 171YbOH chemical enhancement induced by the 171Yb atomic excitation. The spectrum obtained by subtracting the signals with and without chemical enhancement induced by the 171Yb atomic excitation is presented in panel (c). The predicted spectrum was obtained assuming a 5 K rotational temperature, 90 MHz FWHM Lorentzian lineshape, and optimized parameters (see below). The head at 17 323.55 cm−1 is an unresolved blend of the PP12 + PQ12(1), PP12 + PQ12(2), and PP12 + PQ12(3) branch features (see below). Unlike the 173YbOH spectrum shown in Fig. 3, there are numerous 171YbOH spectral features that are unblended. For example, the feature marked “A” in Fig. 4 is the F1=1F1=1 component of QQ11 + QR11(1), “B” is the F1=5F1=5 component of PP11 + PQ11(5), and “C” is the F1=2F1=2 component of QQ11 + QR11(2) (see below).

FIG. 4.

High-resolution absorption spectra in the bandhead region of the 000A2Π1/2-X2Σ+ band of YbOH of a cryogenic buffer-gas cooled sample: (a) the spectrum with no chemical enhancement; (b) the spectrum recorded when the F″ = 1/2 → F′ = 1/2 component of the 3P11S0 transition of 171Yb (17 991.9292 cm−1) is excited; (c) difference in enhanced and unenhanced spectra (a) and (b); (d) prediction of the 171YbOH absorption spectrum using the optimized parameters given in Table I, a rotational temperature of 5 K and FWHM linewidth of 90 MHz. Units of d y-axis are arbitrary.

FIG. 4.

High-resolution absorption spectra in the bandhead region of the 000A2Π1/2-X2Σ+ band of YbOH of a cryogenic buffer-gas cooled sample: (a) the spectrum with no chemical enhancement; (b) the spectrum recorded when the F″ = 1/2 → F′ = 1/2 component of the 3P11S0 transition of 171Yb (17 991.9292 cm−1) is excited; (c) difference in enhanced and unenhanced spectra (a) and (b); (d) prediction of the 171YbOH absorption spectrum using the optimized parameters given in Table I, a rotational temperature of 5 K and FWHM linewidth of 90 MHz. Units of d y-axis are arbitrary.

Close modal

Although the RR1G branch features of the odd isotopologues are less blended than the PP1G + PQ1G and QQ1G + QR1G branch features, recording the absorption spectrum of the CBGC sample was still critical for spectral disentanglement and assignment of the molecular beam LIF spectrum. This is illustrated in Fig. 5 where the observed and predicted molecular beam LIF (left side) and CBGC absorption (right side) spectra in the region of the RR11(2) (even isotopologues) and RR1G(2) (odd isotopologues) lines are presented. Predicted molecular beam spectra assumed a rotational temperature of 20 K and an FWHM linewidth of 30 MHz, while the CBGC spectra assumed those to be 5 K and 90 MHz, respectively, and both used optimized parameters (see below). The CBGC absorption spectrum recorded without the atomic excitation is presented in panel (a). The CBGC absorption spectra recorded with the atomic excitation laser tuned to the 3P11S0 transitions of 176Yb, 174Yb, and 171Yb are presented panels (b)–(d), respectively. The 171YbOH absorption spectrum in panel (e) was obtained by taking the difference of “a” and “d,” revealing a spectral feature that is obscured in the higher resolution molecular beam LIF spectrum by the much more intense RR11(2) line of 174YbOH.

FIG. 5.

Observed and predicted molecular beam (MB) LIF (left side) and cryogenic buffer-gas cooled (CBGC) absorption (right side) spectra in the region of the RR11(2) line of the 000A2Π1/2-X2Σ+ band of YbOH. Predicted MB spectra assumed a rotational temperature of 15 K and FWHM linewidth of 30 MHz, while the CBGC spectrum used T = 5 K and FWHM = 90 MHz, and both used the optimized parameters of Table I (a) recorded with the atomic excitation laser blocked, (b) recorded with the atomic excitation laser tuned to the 3P11S0 transition of 176Yb (17 991.9685 cm−1), (c) recorded with the atomic excitation laser tuned to the 3P11S0 transition of 174Yb (17 992.0003 cm−1), and (d) recorded with the atomic excitation laser tuned to the F″ = 1/2 → F′ = 1/2 component of the 3P11S0 transition of 171Yb (17 991.9292 cm−1); (e) difference in “a” and “d,” showing direct, model-free isolation of the 171YbOH signal from other isotopologues.

FIG. 5.

Observed and predicted molecular beam (MB) LIF (left side) and cryogenic buffer-gas cooled (CBGC) absorption (right side) spectra in the region of the RR11(2) line of the 000A2Π1/2-X2Σ+ band of YbOH. Predicted MB spectra assumed a rotational temperature of 15 K and FWHM linewidth of 30 MHz, while the CBGC spectrum used T = 5 K and FWHM = 90 MHz, and both used the optimized parameters of Table I (a) recorded with the atomic excitation laser blocked, (b) recorded with the atomic excitation laser tuned to the 3P11S0 transition of 176Yb (17 991.9685 cm−1), (c) recorded with the atomic excitation laser tuned to the 3P11S0 transition of 174Yb (17 992.0003 cm−1), and (d) recorded with the atomic excitation laser tuned to the F″ = 1/2 → F′ = 1/2 component of the 3P11S0 transition of 171Yb (17 991.9292 cm−1); (e) difference in “a” and “d,” showing direct, model-free isolation of the 171YbOH signal from other isotopologues.

Close modal

The precisely measured transition wavenumber for 94 spectral features of the 000Ã2Π1/2-X̃2Σ+ band of 173YbOH, which are assigned to 124 transitions, is given in Table S1 of the supplementary material, along with the difference between the observed and calculated wavenumbers. The precisely measured transition wavenumber for 63 spectral features of the 000Ã2Π1/2-X̃2Σ+ band of 171YbOH, which are assigned to 68 transitions, is given in Table S2 of the supplementary material, along with the difference between the observed and calculated wavenumbers. Also presented are the quantum number assignments.

The analysis of the 171YbOH and 173YbOH 000Ã2Π1/2-X̃2Σ+ spectra was nearly identical to that of the (0,0)A2Π1/2-X2Σ+ band of 171YbF (Ref. 49) and 173YbF (Ref. 48) and will not be described in detail here. In brief, the observed transition wavenumbers listed in Tables S1 and S2 of the supplementary material were used as inputs into a weighted least-squares optimization routine. The molecular beam data were given twice the weight of the more extensive CBGC data due to the high spectral resolution (30 MHz FWHM vs 90 MHz) and the fact that it was co-recorded with the I2 wavelength calibration spectrum. The X̃2Σ+(0,0,0) and Ã2Π1/2(0,0,0) energies were obtained by diagonalizing 24 × 24 (=(2S + 1) (2I1 + 1)(2I2 + 1)) and 48 × 48 (=2(2S + 1)(2I1 + 1) (2I2 + 1)) matrices for 173YbOH and 8 × 8 and 16 × 16 matrices for 171YbOH constructed in a sequentially coupled Hund’s case (aβJ) basis set, ηΛSΣJΩJI(Yb)F1F1I(H)F. Note that a Hund’s case (aβJ) basis is used for the X̃2Σ+(0,0,0) state even though the energy level pattern is that of a molecule close to a Hund’s case (bβS). The X̃2Σ+(0,0,0) states were modeled using an effective Hamiltonian48,49 that included rotation (B and D), spin-rotation (γ), Yb and H magnetic hyperfine (Fermi contact, bF, and dipolar, c), and 173Yb axial nuclear electric quadrupole (e2q0Q) parameters. The energy levels for the Ã2Π1/2(0,0,0) states were modeled using an effective Hamiltonian that included the spin–orbit (A), rotation (B and D), Λ-doubling (p + 2q), Yb magnetic hyperfine (Λ-doubling type, d, and electronic orbital hyperfine, a) and 173Yb axial nuclear electric quadrupole (e2q0Q).50 In the Hund’s case (aβJ) limit, the determinable combination of magnetic hyperfine parameters is h1/2abF2c3, and the levels can be accurately modeled using aÎzL̂z and by fixing the Fermi contact, bF, and dipolar, c, terms to zero. The dataset, which is restricted to the Ω=1/2 levels of the Ã2Π(0,0,0) state, is insensitive to the perpendicular component of the nuclear electric quadrupole interaction (e2q2Q). The effective Hamiltonians are described in more detail in the supplementary material.

Analyses that floated various combinations of parameters were attempted. In all cases, the small proton magnetic hyperfine parameters for the X̃2Σ+(0,0,0) state were held fixed to those determined from the analysis of the 174YbOH microwave spectrum (bF = 4.80 MHz and c = 2.46 MHz).39 The spin–orbit parameter, A, of the Ã2Π(0,0,0) state was constrained to the previously determined52 value of 1350 cm−1. The analysis is relatively insensitive to the value of A, other than the obvious linear displacement, because of the large separation between the Ã2Π1/2(0,0,0) and Ã2Π3/2(0,0,0) states and the relatively small rotational parameter, B (i.e., the rotationally induced spin-uncoupling effect is negligible because B/A ≅ 2 × 10−4). The centrifugal correction parameters, D, for the X̃2Σ+(0,0,0) and Ã2Π(0,0,0) states and the small spin-rotation parameter, γ, for the X̃2Σ+(0,0,0) state were constrained to the values predicted from extrapolation of the values for 174YbOH using the expected isotopic mass dependence. In the end, the 173YbOH dataset was satisfactorily fit by optimizing 10 parameters, and the 171YbOH dataset, by optimizing 8 parameters. The standard deviation of the 171YbOH and 173YbOH fits was 27 and 25 MHz, respectively, which is commensurate with estimated weighted measurement uncertainty. The optimized parameters and associated errors are presented in Table I, along with those for the A2Π1/2(v = 0) and X2Σ+(v = 0) states of 171YbF (Ref. 49) and 173YbF (Ref. 48). The spectroscopic parameters are independently well determined with the largest correlation coefficient being between B″ and B′, which for the 171YbOH and 173YbOH fits are 0.854 and 0.960, respectively.

TABLE I.

Parameters in wavenumbers (cm−1) for the X̃2Σ+(0,0,0) and Ã2Π1/2(0,0,0) states of 171YbOH and173YbOH and the A2Π1/2(v = 0) and X2Σ+(v = 0) states of 171YbF and 173YbF.

Par.171YbOH171YbFa173YbOH173YbFb
X̃2Σ+ B 0.245 497(22) 0.241710 98(6) 0.245 211(18) 0.241 434 8 (12) 
D × 106 0.252 4(Fix) 0.219 8(17) 0.219 0(Fix) 0.227(Fix) 
γ −0.002 697 (fix) −0.000 448(1) −0.002 704(Fix) −0.000 446 4 (24) 
bF(Yb) 0.227 61(33) 0.242 60(37) −0.062 817(67) −0.067 04 (8) 
c(Yb) 0.007 8(14) 0.009 117(12) −0.002 73(45) −0.002 510 (12) 
e2Qq0(Yb) N/A N/A −0.110 7(16) −0.109 96 (6) 
bF(H or F) 0.000 160(fix) 0.005 679(Fix) 0.000 160(fix) 0.005 679(Fix) 
c(H or F) 0.000 082(fix) 0.002 849(Fix) 0.000 082(fix) 0.002 849(Fix) 
Ã2Π1/2 A 1 350 (fix) 1 365.3(Fix) 1 350.0(Fix) 1 365.294 (fix) 
B 0.253 435(24) 0.248 056 8(35) 0.253 185(16) 0.247 79 (6) 
D × 106 0.260 8 (fix) 0.203 2 (fix) 0.240 5(Fix) 0.203 2 (fix) 
p+2q −0.438 667(82) −0.397 62(Fix) −0.438 457(64) −0.397 20 (fix) 
a(Yb) 0.014 8(15) 0.012 8(61) −0.004 22(20) −0.005 07(18) 
d(Yb) 0.031 99(58) 0.033 1(16) −0.008 73(13) −0.008 85 (18) 
e2Qq0(Yb) N/A N/A −0.064 2(17) −0.064 7(12) 
T00 17 998.636 19(24) 18 788.650 2(4) 17 998.602 68(13) 18 788.859 39 
Par.171YbOH171YbFa173YbOH173YbFb
X̃2Σ+ B 0.245 497(22) 0.241710 98(6) 0.245 211(18) 0.241 434 8 (12) 
D × 106 0.252 4(Fix) 0.219 8(17) 0.219 0(Fix) 0.227(Fix) 
γ −0.002 697 (fix) −0.000 448(1) −0.002 704(Fix) −0.000 446 4 (24) 
bF(Yb) 0.227 61(33) 0.242 60(37) −0.062 817(67) −0.067 04 (8) 
c(Yb) 0.007 8(14) 0.009 117(12) −0.002 73(45) −0.002 510 (12) 
e2Qq0(Yb) N/A N/A −0.110 7(16) −0.109 96 (6) 
bF(H or F) 0.000 160(fix) 0.005 679(Fix) 0.000 160(fix) 0.005 679(Fix) 
c(H or F) 0.000 082(fix) 0.002 849(Fix) 0.000 082(fix) 0.002 849(Fix) 
Ã2Π1/2 A 1 350 (fix) 1 365.3(Fix) 1 350.0(Fix) 1 365.294 (fix) 
B 0.253 435(24) 0.248 056 8(35) 0.253 185(16) 0.247 79 (6) 
D × 106 0.260 8 (fix) 0.203 2 (fix) 0.240 5(Fix) 0.203 2 (fix) 
p+2q −0.438 667(82) −0.397 62(Fix) −0.438 457(64) −0.397 20 (fix) 
a(Yb) 0.014 8(15) 0.012 8(61) −0.004 22(20) −0.005 07(18) 
d(Yb) 0.031 99(58) 0.033 1(16) −0.008 73(13) −0.008 85 (18) 
e2Qq0(Yb) N/A N/A −0.064 2(17) −0.064 7(12) 
T00 17 998.636 19(24) 18 788.650 2(4) 17 998.602 68(13) 18 788.859 39 
a

X2Σ+(v = 0) values from Ref. 47; A2Π1/2(v = 0) values from Ref. 56.

b

From the combined fit of optical and microwave data (Ref. 46).

Spectral predictions, such as those presented in Figs. 25, were essential for the assignment and analysis. The approach was identical to that used in modeling the (0,0)A2Π1/2-X2Σ+ band of 171YbF (Ref. 49) and 173YbF (Ref. 48). The electric dipole transition moment matrix was constructed in a sequentially coupled Hund’s case (aβJ) basis set ηΛSΣJΩJI(Yb)F1F1I(H)F and cross multiplied by the eigenvectors to produce the transition moments. The relative intensities were taken as the product of the square of the transition moment, a Boltzmann factor, and the relative isotope abundance. The relative intensities and a Lorentzian lineshape were then used to predict the spectra. The complete set of spectroscopic parameters for the six most abundant isotopologues used in these predictions is presented in Table S3 of the supplementary material. The parameter values for 171YbOH and 173YbOH were taken from the present study, while those for 174YbOH [X̃2Σ+(0,0,0)] were taken from the analysis of the microwave spectrum.39 The parameters for 174YbOH [Ã2Π1/2(0,0,0)], 172YbOH [X̃2Σ+(0,0,0)], and 172YbOH [Ã2Π1/2(0,0,0)] were taken from the analysis of the optical spectra.37 The parameters for the 170YbOH and 176YbOH isotopologues were obtained by extrapolation of those for 171YbOH, 172YbOH, 173YbOH, and 174YbOH, using the expected mass dependence. In all cases, the X̃2Σ+(0,0,0) proton magnetic hyperfine parameters were constrained to the values of 174YbOH, and those for the Ã2Π1/2(0,0,0) were constrained to zero.

The major objectives of this study are to analyze the 000Ã2Π1/2-X̃2Σ+ rotationally resolved spectra of the odd isotopologues, precisely determine the Ã2Π1/2(0,0,0) and X̃2Σ+(0,0,0) energy levels, and demonstrate the utility of chemical enhancement as a spectroscopic tool. The obtained spectral information is needed for the design and implementation of experiments under way to characterize NSD-PV and search for new T-violating BSM physics. The calculated energies for the N = 0–4 levels of the X̃2Σ+(0,0,0) state and J = 0.5–5.5 levels of the Ã2Π1/2(0,0,0) state for 171YbOH are presented in Tables S4 and S5, respectively, of the supplementary material. The calculated energies for the N = 0–4 levels of the X̃2Σ+(0,0,0) state and J = 0.5–5.5 levels of the Ã2Π1/2(0,0,0) state for 173YbOH are presented in Tables S6 and S7, respectively, of the supplementary material. Spectroscopic parameters given in Table S3 can be used to accurately predict the energies of higher rotational levels of the even and odd isotopologues. A perusal of the energy level patterns for the odd isotopologues Ã2Π1/2(0,0,0) states reveal that the large Λ-doubling, relative to the rotational spacing, makes the patterns more similar to that of a2Σ state than the a2Π1/2 state. Specifically, the pattern is that of rotationally spaced J = N ± 1/2 pairs of levels (i.e., ρ-doublets), with the exception of the N = 0, J = 1/2 lowest energy level, which has negative parity. Although the energy levels could be modeled as the a2Σ state, a2Σ+2Σ transition is electric dipole forbidden and not consistent with the observed intensities.

The predicted stick spectra for 171YbOH and 173YbOH generated using the optimized spectroscopic parameters and a rotational temperature of 15 K are presented in Figs. 6 and 7, respectively. Of particular interest are the transitions that terminate at the J = 0.5, +parity level of the Ã2Π1/2(0,0,0) state, to be used for photon cycling and laser cooling.18 For the even isotopologues, these are the PQ12 (1) and PP12 (1) transitions, and for the odd isotopologues, they are the PQ1G + PP1G (1) transitions. These transitions are in a highly congested region of the spectrum (see Figs. 6 and 7). It is noteworthy that the PQ12 (1) and PP12 (1) lines of the even isotopologues are “rotationally closed” (i.e., the excitation and fluorescence spectra only involve the N″ = 1 levels), whereas in the odd isotopologues, the branching ratio for the (J = 0.5, +) Ã2Π1/2(0,0,0) → (N = 3, −) X̃2Σ+(0,0,0) transition is ∼1% relative to the (J = 0.5, +) Ã2Π1/2(0,0,0) → (N = 1, −) X̃2Σ+(0,0,0) transitions. This is primarily caused by the mixing of the J = 0.5, + parity and the close lying (ΔE ≅ 3 GHz) J = 1.5, + parity levels of the Ã2Π1/2(0,0,0) state by hyperfine terms in the effective Hamiltonian. There is a similar, yet much smaller in magnitude, hyperfine induced mixing of the N = 1 and N = 3 levels of the X̃2Σ+(0,0,0) state.

FIG. 6.

Predicted stick spectra with associated assignments for 171YbOH generated using the optimized spectroscopic parameters shown in Table I and a rotational temperature of 15 K.

FIG. 6.

Predicted stick spectra with associated assignments for 171YbOH generated using the optimized spectroscopic parameters shown in Table I and a rotational temperature of 15 K.

Close modal
FIG. 7.

Predicted stick spectra with associated assignments for 173YbOH generated using the optimized spectroscopic parameters shown in Table I and a rotational temperature of 15 K.

FIG. 7.

Predicted stick spectra with associated assignments for 173YbOH generated using the optimized spectroscopic parameters shown in Table I and a rotational temperature of 15 K.

Close modal

The hyperfine parameters are of particular interest because they are sensitive probes of the electronic wavefunction in the region of the nuclei. The ability to accurately predict hyperfine parameters is the most direct gauge of the computational methodology used to predict Ws, Wd, WQ, and WM coupling constants of Eq. (2). As part of the recent calculation,16 the nuclear magnetic quadrupole interaction constants WM for YbOH, the ground state A|| (= bF +23c) magnetic hyperfine constant for 173YbOH and 171YbOH, and the axial electric quadrupole coupling constant e2Qq0 for 173YbOH were predicted. These predictions were performed using numerical gradients of a four-component Dirac Coulomb Hamiltonian, with electronic correlation treated using the multireference Fock-space coupled cluster method (FSCC). Similarly, a recent calculation53 of Wd and Ws includes predictions for both the ground state A|| and A(= bF23c) magnetic hyperfine constants for 173YbOH. In this case, quasi-relativistic two-component calculations with many-body interactions treated at the level of complex generalized Hartree–Fock (cGHF) and complex generalized Kohn–Sham (cGKS) density functional theory methods were implemented. Most recently, the axial electric quadrupole coupling constant e2Qq0 for 173YbOH was predicted using analytic gradients formulated from spin–orbit CCSD(T) theory, computed against an atomic mean-field scalar relativistic (SFX2C-AMF) Hamiltonian.54 The predicted hyperfine parameters for the X̃2Σ+(0,0,0) state are compared with the measured values in Table II. There are no theoretical predictions for the hyperfine interactions in the Ã2Π1/2(0,0,0) state. The calculated hyperfine parameters obtained using relativistic coupled cluster approaches16,54 are in excellent agreement with the observed values, with the calculated magnetic hyperfine parameters being 2.5% (Ref. 16) larger in magnitude and the e2Qq0 parameter being 5.5% (Ref. 16) and 5.2% (Ref. 54) larger in magnitude. Evidently, calculating the core polarization is more difficult than calculating the valence electron properties. Nonetheless, the excellent agreement suggests that the calculated WM coupling constants for ground state 173YbOH of Ref. 16 and the effective electric field, Eeff, of Ref. 55, which was obtained using the same computational methodology, are quantitatively accurate. The predicted Eeff is relevant to the electron electric dipole moment (eEDM) measurements for all the isotopologues of YbOH. The quasirelativistic two component calculations in Ref. 53 give magnetic hyperfine constants that are smaller in magnitude by ∼15% and 31% for the cGHF and cGKS methods, respectively.

TABLE II.

Comparison of the measured hyperfine parameters of the X̃2Σ+0,0,0 state of 171,173YbOH to calculated values.

IsotopologueParameterMeasured (MHz)Theory Reference 16 (MHz)Theory Reference 51 cGHF(MHz)Theory Reference 51 cGKS(MHz)Theory Reference 52 (MHz)
171YbOH A||a 6979 (35) 7174.9    
171YbOH Ab 6745(15)     
173YbOH A|| −1929(11) −1976.3 −1600 −1300  
173YbOH A −1856 (5)  −1600   
173YbOH e2Qq0 −3319 (48) −3502   −3492 
IsotopologueParameterMeasured (MHz)Theory Reference 16 (MHz)Theory Reference 51 cGHF(MHz)Theory Reference 51 cGKS(MHz)Theory Reference 52 (MHz)
171YbOH A||a 6979 (35) 7174.9    
171YbOH Ab 6745(15)     
173YbOH A|| −1929(11) −1976.3 −1600 −1300  
173YbOH A −1856 (5)  −1600   
173YbOH e2Qq0 −3319 (48) −3502   −3492 
a

For a σ orbital: A|| = b+c=bF+23c.

b

For a σ orbital: A=b=bF13c.

Although the high-level relativistic electronic structure calculation16 accurately predicts the hyperfine interactions for the X̃2Σ+(0,0,0) state, a less quantitative, molecular-orbital based, model using atomic information is also useful. Such a model provides chemical insight, can be used to rationalize trends among similar molecules (e.g., YbF and YbOH), and is able to predict hyperfine interactions for excited electronic states that are not readily addressed by relativistic electronic structure calculations. The X̃2Σ+(0,0,0) hyperfine fitting parameters are related to various averages over the spatial coordinates of the electrons by51,56

(3)
(4)

and

(5)

where ŝzi is the spin angular momentum operators for the ith electron, δi(r) is a Dirac delta function, and r and θ are polar coordinates. In the case of bF and c, the sum runs only over unpaired electrons, whereas for e2Qq0, the sum is over all electrons. The observed ratio bF(171YbOH)/bF(173YbOH) for the X̃2Σ+(0,0,0) state (= −3.62 ± 0.04) is in excellent agreement with that of the nuclear g-factors, gN (171Yb)/gN (173Yb) (= −3.630). Meanwhile, the ratio for the less well determined dipolar ratio c(171YbOH)/c(173YbOH) (=−2.86 ± 0.57) is within two standard deviations. To a first approximation, the dominant electronic configuration for the X̃2Σ+(0,0,0) state has one unpaired electron in a hybridized 6s/6p/5d σ-type, Yb+-centered orbital, which is polarized away from the Yb–OH bond. The hybridization is driven by the stabilization achieved from shifting the center of the charge for the unpaired electron away from the electrophilic end of the Yb+OH molecule. Based upon a comparison with bF (171Yb+) (= 0.4217 cm−1)57 for Yb+(f146s2S1/2), the σ-type orbital is ∼54% 6s character. The bF(171YbOH) and bF(173YbOH) values are ∼7% smaller in magnitude than the corresponding values for YbF. Evidently, the unpaired 6s electron is more effectively polarized by OH than F (i.e., larger 6s/6p/5d hybridization). This is consistent with the point charge electrostatic model prediction that successfully models the permanent electric dipole moments for ground states of YbOH and YbF.37 The Yb+-centered, unpaired electron was predicted to be shifted away from the Yb–F and Yb–O bond by 0.859 and 0.969 Å, respectively, reflecting the increased 6s/6p/5d hybridization in YbOH relative to YbF.

Unlike the noted change in bF, the 173Yb axial nuclear electric quadrupole parameter, e2q0Q, for the ground states of 173YbOH and 173YbF is very nearly identical, indicating that core polarization makes a substantial contribution relative to that of the unpaired valence electron. As evident from a comparison of Eqs. (4) and (5), the contribution from the unpaired electron to e2q0Q is

(6)

Substitution of Q = 280 fm2 and gN of −0.271 95 gives e2Qq0(unpaired) = −0.0456 cm-1, which is ∼40% (−0.0456/−0.1107) of the determined e2Qq0 value for the X̃2Σ+(0,0,0) state. The remaining −0.0651 cm−1 portion is due to core polarization.

An interpretation of the excited state hyperfine parameters is more difficult because bF and c were constrained to zero. The Λ-doubling type magnetic hyperfine parameter, d, on the other hand, is well determined and is given by51,56

(7)

where ŝi+ is the one electron spin angular momentum raising operator and ϕi is the azimuthal angle of the electron. The ratio of d for 171YbOH and 173YbOH (=0.031 99/−0.008 73 ≅ 3.66) is in excellent agreement with the expected ratio of gN values. The same rationale presented for the observed d values of YbF58 is applicable to YbOH, and an estimate gives values for 171YbOH and 173YbOH of 0.0657 and −0.0182 cm−1, respectively, which are the correct signs but approximately a factor of two too large. These values were obtained by assuming that the orbital of the sole unpaired electron for the Ã2Π1/2(0,0,0) state is pure Yb+ 6p. The more realistic assumption that the orbital of the sole unpaired electron is an admixture of Yb+ 6p, Yb 6p, and Yb 5d and would give values in closer agreement with the experimental values.

The magnitude of the e2q0Q parameter decreases from 0.1107(16) to 0.0642(17) cm−1 upon X̃2Σ+(0,0,0) to Ã2Π1/2(0,0,0) excitation. The angular expectation value for the 6p orbital contribution to the open shell σ-orbital of the X̃2Σ+(0,0,0) state is p03cos2θi1p0=4/5 while that for the open shell π-orbital of the Ã2Π1/2(0,0,0) state is p±13cos2θi1p±1=2/5. Accordingly, if the σ- and π-orbitals of the two states had identical 6p contribution, then e2Qq0(unpaired) for Ã2Π1/2(0,0,0) would be ∼0.0228 cm−1 [i.e., half the magnitude and opposite sign of the X̃2Σ+(0,0,0) state value]. Assuming that the core polarization contributions in the Ã2Π1/2(0,0,0) state are the same as the X̃2Σ+(0,0,0) state, then e2Qq0 for the Ã2Π1/2(0,0,0) state is predicted to be −0.0423 cm−1, in qualitative agreement with observed values given the assumptions being made.

A novel spectroscopic technique utilizing laser-induced chemical reactions to distinguish features in congested and overlapped spectra was demonstrated. This technique was utilized to determine the fine and hyperfine structure of 171,173YbOH for the first time. The derived molecular parameters are consistent with previous measurements of the even isotopologues 172,174YbOH as well as with isoelectronic 171,173YbF. The determined hyperfine parameters also provide an experimental comparison to check the quality of the calculated P- and T, P-violating coupling constants Ws, WP, Wd, and WM. Finally, the characterization of the X̃2Σ+0,0,0Ã2Π1/2(0,0,0) transition of 171,173YbOH performed in this work will aid in the implementation of NSD-PV and NMQM measurements, laser cooling and photon cycling, and future spectroscopic investigations of these promising molecular isotopologues.

See the supplementary material for the description of the effective Hamiltonian and for the transition wavenumber for the A2Π1/2X2Σ+ (0,0,0) electronic transition of 171YbOH (Table S1), the transition wavenumber for the A2Π1/2X2Σ+ (0,0,0) electronic transition of 173YbOH (Table S2), spectroscopic parameters in wavenumbers (cm−1) used to model the X̃2Σ+(0,0,0) and Ã2Π1/2(0,0,0) states of YbOH (Table S3), the calculated low-rotational (N = 0–4) X̃2Σ+(0,0,0) state energies of 171YbOH (Table S4), the calculated low-rotational (J = 0.5–5.5) Ã2Π1/2(0,0,0) state energies of 171YbOH (Table S5), the calculated low-rotational (N = 0–4) X̃2Σ+(0,0,0) state energies of 173YbOH (Table S6), and the calculated low-rotational (J = 0.5–5.5) Ã2Π1/2(0,0,0) state energies of 173YbOH (Table S7).

The research at Arizona State University and Caltech was supported by grants from the Heising–Simons Foundation (ASU: Grant No. 2018-0681; Caltech: Grant No. 2019-1193). N.R.H. acknowledges support from the NIST Precision Measurement (Grant No. 60NANB18D253) and the NSF CAREER Award (Grant No. PHY-1847550). We thank Graceson Aufderheide and Richard Mawhorter (Physics and Astronomy Department, Pomona College, Pomona, CA 91711, USA) for assistance with Yb + Te(OH)6 target fabrication, Phelan Yu for help fabricating the Yb + Yb(OH)3 targets, and Dr. Anh Le (Chemistry Dept., Georgia Institute of Technology, Atlanta, GA 30318, USA) for assistance in recording the molecular beam spectra. We thank Ben Augenbraun and Phelan Yu for feedback on the article.

The data that support the findings of this study are available within the article and its supplementary material.

1.
D.
DeMille
,
J. M.
Doyle
, and
A. O.
Sushkov
,
Science
357
,
990
(
2017
).
2.
M. S.
Safronova
,
D.
Budker
,
D.
DeMille
,
D. F. J.
Kimball
,
A.
Derevianko
, and
C. W.
Clark
,
Rev. Mod. Phys.
90
,
025008
(
2018
).
3.
W. B.
Cairncross
and
J.
Ye
,
Nat. Rev. Phys.
1
,
510
(
2019
).
4.
N. R.
Hutzler
,
Quantum Sci. Tech.
5
,
044011
(
2020
).
5.
V.
Andreev
,
D. G.
Ang
,
D.
DeMille
,
J. M.
Doyle
,
G.
Gabrielse
,
J.
Haefner
,
N. R.
Hutzler
,
Z.
Lasner
,
C.
Meisenhelder
,
B. R.
O’Leary
,
C. D.
Panda
,
A. D.
West
,
E. P.
West
, and
X.
Wu
,
Nature
562
,
355
(
2018
).
6.
W. B.
Cairncross
,
D. N.
Gresh
,
M.
Grau
,
K. C.
Cossel
,
T. S.
Roussy
,
Y.
Ni
,
Y.
Zhou
,
J.
Ye
, and
E. A.
Cornell
,
Phys. Rev. Lett.
119
,
153001
(
2017
).
7.
I.
Kozyryev
and
N. R.
Hutzler
,
Phys. Rev. Lett.
119
,
133002
(
2017
).
8.
N. J.
Fitch
,
J.
Lim
,
E. A.
Hinds
,
B. E.
Sauer
, and
M. R.
Tarbutt
,
Quantum Sci. Technol.
6
,
014006
(
2021
).
9.
D.
DeMille
,
S. B.
Cahn
,
D.
Murphree
,
D. A.
Rahmlow
, and
M. G.
Kozlov
,
Phys. Rev. Lett.
100
,
023003
(
2008
).
10.
V. V.
Flambaum
and
I. B.
Khriplovich
,
Phys. Lett. A
110
,
121
(
1985
).
11.
M. G.
Kozlov
,
L. N.
Labzovskii
, and
A. O.
Mitrushchenkov
,
Zh. Eksp. Teor. Fiz.
100
,
749
(
1991
).
12.
E.
Altuntas
,
J.
Ammon
,
S. B.
Cahn
, and
D.
DeMille
,
Phys. Rev. Lett.
120
,
142501
(
2018
).
13.
E.
Altuntas
,
J.
Ammon
,
S. B.
Cahn
, and
D.
DeMille
,
Phys. Rev. A
97
,
042101
(
2018
).
14.
V. A.
Dzuba
,
V. V.
Flambaum
, and
Y. V.
Stadnik
,
Phys. Rev. Lett.
119
,
223201
(
2017
).
15.
E. B.
Norrgard
,
D. S.
Barker
,
S.
Eckel
,
J. A.
Fedchak
,
N. N.
Klimov
, and
J.
Scherschligt
,
Commun. Phys.
2
,
77
(
2019
).
16.
M.
Denis
,
Y.
Hao
,
E.
Eliav
,
N. R.
Hutzler
,
M. K.
Nayak
,
R. G. E.
Timmermans
, and
A.
Borschesvky
,
J. Chem. Phys.
152
,
084303
(
2020
).
17.
D. E.
Maison
,
L. V.
Skripnikov
, and
V. V.
Flambaum
,
Phys. Rev. A
100
,
032514
(
2019
).
18.
B. L.
Augenbraun
,
Z. D.
Lasner
,
A.
Frenett
,
H.
Sawaoka
,
C.
Miller
,
T. C.
Steimle
, and
J. M.
Doyle
,
New J. Phys.
22
,
022003
(
2020
).
19.
C. S.
Wood
,
S. C.
Bennett
,
D.
Cho
,
B. P.
Masterson
,
J. L.
Roberts
,
C. E.
Tanner
, and
C. E.
Wieman
,
Science
275
,
1759
(
1997
).
20.
V. V.
Flambaum
,
D.
DeMille
, and
M. G.
Kozlov
,
Phys. Rev. Lett.
113
,
103003
(
2014
).
21.
K.
Gaul
,
S.
Marquardt
,
T.
Isaev
, and
R.
Berger
,
Phys. Rev. A
99
,
032509
(
2019
).
22.
X.
Wu
,
Z.
Han
,
J.
Chow
,
D. G.
Ang
,
C.
Meisenhelder
,
C. D.
Panda
,
E. P.
West
,
G.
Gabrielse
,
J. M.
Doyle
, and
D.
DeMille
,
New J. Phys.
22
,
023013
(
2020
).
23.
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
,
Phys. Rev. Lett.
124
,
053201
(
2020
).
24.
P.
Aggarwal
,
H. L.
Bethlem
,
A.
Borschevsky
,
M.
Denis
,
K.
Esajas
,
P. A. B.
Haase
,
Y.
Hao
,
S.
Hoekstra
,
K.
Jungmann
,
T. B.
Meijknecht
,
M. C.
Mooij
,
R. G. E.
Timmermans
,
W.
Ubachs
,
L.
Willmann
, and
A.
Zapara
,
Eur. Phys. J. D
72
,
197
(
2018
).
25.
B.
Graner
,
Y.
Chen
,
E. G.
Lindahl
, and
B. R.
Heckel
,
Phys. Rev. Lett.
116
,
161601
(
2016
).
26.
M.
Bishof
,
R. H.
Parker
,
K. G.
Bailey
,
J. P.
Greene
,
R. J.
Holt
,
M. R.
Kalita
,
W.
Korsch
,
N. D.
Lemke
,
Z. T.
Lu
,
P.
Mueller
,
T. P.
O’Connor
,
J. T.
Singh
, and
M. R.
Dietrich
,
Phys. Rev. C
94
,
025501
(
2016
).
27.
N.
Sachdeva
,
I.
Fan
,
E.
Babcock
,
M.
Burghoff
,
T. E.
Chupp
,
S.
Degenkolb
,
P.
Fierlinger
,
S.
Haude
,
E.
Kraegeloh
,
W.
Kilian
,
S.
Knappe-Grüneberg
,
F.
Kuchler
,
T.
Liu
,
M.
Marino
,
J.
Meinel
,
K.
Rolfs
,
Z.
Salhi
,
A.
Schnabel
,
J. T.
Singh
,
S.
Stuiber
,
W. A.
Terrano
,
L.
Trahms
, and
J.
Voigt
,
Phys. Rev. Lett.
123
,
143003
(
2019
).
28.
F.
Allmendinger
,
I.
Engin
,
W.
Heil
,
S.
Karpuk
,
H.-J.
Krause
,
B.
Niederländer
,
A.
Offenhäusser
,
M.
Repetto
,
U.
Schmidt
, and
S.
Zimmer
,
Phys. Rev. A
100
,
022505
(
2019
).
29.
E. R.
Tardiff
,
E. T.
Rand
,
G. C.
Ball
,
T. E.
Chupp
,
A. B.
Garnsworthy
,
P.
Garrett
,
M. E.
Hayden
,
C. A.
Kierans
,
W.
Lorenzon
,
M. R.
Pearson
,
C.
Schaub
, and
C. E.
Svensson
,
Hyperfine Interact.
225
,
197
(
2014
).
30.
O.
Grasdijk
,
O.
Timgren
,
J.
Kastelic
,
T.
Wright
,
S.
Lamoreaux
,
D.
DeMille
,
K.
Wenz
,
M.
Aitken
,
T.
Zelevinsky
,
T.
Winick
, and
D.
Kawall
, “
CeNTREX: A new search for time-reversal symmetry violation in the 205Tl nucleus
,”
Quantum Sci. Technol.
(to be published) (
2021
).
31.
R. F.
Garcia Ruiz
,
R.
Berger
,
J.
Billowes
,
C. L.
Binnersley
,
M. L.
Bissell
,
A. A.
Breier
,
A. J.
Brinson
,
K.
Chrysalidis
,
T. E.
Cocolios
,
B. S.
Cooper
,
K. T.
Flanagan
,
T. F.
Giesen
,
R. P.
de Groote
,
S.
Franchoo
,
F. P.
Gustafsson
,
T. A.
Isaev
,
Á.
Koszorús
,
G.
Neyens
,
H. A.
Perrett
,
C. M.
Ricketts
,
S.
Rothe
,
L.
Schweikhard
,
A. R.
Vernon
,
K. D. A.
Wendt
,
F.
Wienholtz
,
S. G.
Wilkins
, and
X. F.
Yang
,
Nature
581
,
396
(
2020
).
32.
S. A.
Murthy
,
D.
Krause
, Jr.
,
Z. L.
Li
, and
L. R.
Hunter
,
Phys. Rev. Lett.
63
,
965
(
1989
).
33.
V. V.
Flambaum
,
Phys. Lett. B
320
,
211
(
1994
).
34.
J.
Engel
,
M. J.
Ramsey-Musolf
, and
U.
Van Kolck
,
Prog. Part. Nucl. Phys.
71
,
21
(
2013
).
35.
N.
Yamanaka
,
B. K.
Sahoo
,
N.
Yoshinaga
,
T.
Sato
,
K.
Asahi
, and
B. P.
Das
,
Eur. Phys. J. A
53
,
54
(
2017
).
36.
A.
Jadbabaie
,
N. H.
Pilgram
,
J.
Kłos
,
S.
Kotochigova
, and
N. R.
Hutzler
,
New J. Phys.
22
,
022002
(
2020
).
37.
T. C.
Steimle
,
C.
Linton
,
E. T.
Mengesha
,
X.
Bai
, and
A. T.
Le
,
Phys. Rev. A
100
,
052509
(
2019
).
38.
E. T.
Mengesha
,
A. T.
Le
,
T. C.
Steimle
,
L.
Cheng
,
C.
Zhang
,
B. L.
Augenbraun
,
Z.
Lasner
, and
J.
Doyle
,
J. Phys. Chem. A
124
,
3135
(
2020
).
39.
S.
Nakhate
,
T. C.
Steimle
,
N. H.
Pilgram
, and
N. R.
Hutzler
,
Chem. Phys. Lett.
715
,
105
(
2019
).
40.
E. J.
Salumbides
,
K. S. E.
Eikema
,
W.
Ubachs
,
U.
Hollenstein
,
H.
Knöckel
, and
E.
Tiemann
,
Mol. Phys.
104
,
2641
(
2006
).
41.
N. R.
Hutzler
,
H.-I.
Lu
, and
J. M.
Doyle
,
Chem. Rev.
112
,
4803
(
2012
).
42.
H.-I.
Lu
,
J.
Rasmussen
,
M. J.
Wright
,
D.
Patterson
, and
J. M.
Doyle
,
Phys. Chem. Chem. Phys.
13
,
18986
(
2011
).
43.
N. R.
Hutzler
,
M. F.
Parsons
,
Y. V.
Gurevich
,
P. W.
Hess
,
E.
Petrik
,
B.
Spaun
,
A. C.
Vutha
,
D.
Demille
,
G.
Gabrielse
, and
J. M.
Doyle
,
Phys. Chem. Chem. Phys.
13
,
18976
(
2011
).
44.
J. F.
Barry
,
E. S.
Shuman
, and
D.
Demille
,
Phys. Chem. Chem. Phys.
13
,
18936
(
2011
).
45.
S.
Truppe
,
M.
Hambach
,
S. M.
Skoff
,
N. E.
Bulleid
,
J. S.
Bumby
,
R. J.
Hendricks
,
E. A.
Hinds
,
B. E.
Sauer
, and
M. R.
Tarbutt
,
J. Mod. Opt.
65
,
246
(
2018
).
46.
P. E.
Atkinson
,
J. S.
Schelfhout
, and
J. J.
McFerran
,
Phys. Rev. A
100
,
042505
(
2019
).
47.
N. J.
Stone
,
At. Data Nucl. Data Tables
90
,
75
(
2005
).
48.
H.
Wang
,
A. T.
Le
,
T. C.
Steimle
,
E. A. C.
Koskelo
,
G.
Aufderheide
,
R.
Mawhorter
, and
J.-U.
Grabow
,
Phys. Rev. A
100
,
022516
(
2019
).
49.
Z.
Glassman
,
R.
Mawhorter
,
J.-U.
Grabow
,
A.
Le
, and
T. C.
Steimle
,
J. Mol. Spectrosc.
300
,
7
(
2014
).
50.
M. C. L.
Gerry
,
A. J.
Merer
,
U.
Sassenberg
, and
T. C.
Steimle
,
J. Chem. Phys.
86
,
4754
(
1987
).
51.
J. M.
Brown
and
A.
Carrington
,
Rotational Spectroscopy of Diatomic Molecules
(
Cambridge University Press
,
2003
).
52.
T. C.
Melville
and
J. A.
Coxon
,
J. Chem. Phys.
115
,
6974
(
2001
).
53.
K.
Gaul
and
R.
Berger
,
Phys. Rev. A
101
,
012508
(
2020
).
54.
J.
Liu
,
X.
Zheng
,
A.
Asthana
,
C.
Zhang
, and
L.
Cheng
,
J. Chem. Phys.
154
,
064110
(
2021
).
55.
M.
Denis
,
P. A. B.
Haase
,
R. G. E.
Timmermans
,
E.
Eliav
,
N. R.
Hutzler
, and
A.
Borschevsky
,
Phys. Rev. A
99
,
042512
(
2019
).
56.
J. A. J.
Fitzpatrick
,
F. R.
Manby
, and
C. M.
Western
,
J. Chem. Phys.
122
,
084312
(
2005
).
57.
R.
Blatt
,
H.
Schnatz
, and
G.
Werth
,
Phys. Rev. Lett.
48
,
1601
(
1982
).
58.
T. C.
Steimle
,
T.
Ma
, and
C.
Linton
,
J. Chem. Phys.
127
,
234316
(
2007
).

Supplementary Material