Publication Cover
Molecular Physics
An International Journal at the Interface Between Chemistry and Physics
Volume 119, 2021 - Issue 1-2: David Parker Festschrift
1,558
Views
11
CrossRef citations to date
0
Altmetric
Research Articles

Characterisation of the b3Σ+, v = 0 state and its interaction with the A1Π state in aluminium monofluoride

ORCID Icon, ORCID Icon, ORCID Icon, ORCID Icon, ORCID Icon, ORCID Icon, ORCID Icon, ORCID Icon, ORCID Icon & ORCID Icon show all
Article: e1810351 | Received 30 Jun 2020, Accepted 11 Aug 2020, Published online: 04 Sep 2020

Abstract

Recently, we determined the detailed energy level structure of the X1Σ+, A1Π and a3Π states of AlF that are relevant to laser cooling and trapping experiments [Truppe et al., Phys. Rev. A. 100 (5), 052513 (2019)]. Here, we investigate the b3Σ+,v=0 state of the AlF molecule. A rotationally resolved (1 + 2)-REMPI spectrum of the b3Σ+,v=0a3Π,v=0 band is presented and the lifetime of the b3Σ+,v=0 state is measured to be 190(2) ns. Hyperfine-resolved, laser-induced fluorescence spectra of the b3Σ+,v=0X1Σ+,v=1 and the b3Σ+,v=0a3Π,v=0 bands are recorded to determine fine- and hyperfine structure parameters. The interaction between the b3Σ+,v=0 and the nearby A1Π state is studied and the magnitude of the spin–orbit coupling between the two electronic states is derived using three independent methods to give a consistent value of 10(1) cm1. The triplet character of the A state causes an Aa loss from the main AX laser cooling cycle below the 106 level.

GRAPHICAL ABSTRACT

View correction statement:
Correction

1. Introduction

Polar molecules, cooled to low temperatures, have wide-ranging applications in physics and chemistry [Citation1, Citation2]. Recently, we identified the AlF molecule as an excellent candidate for direct laser cooling to low temperatures and with a high density [Citation3]. AlF has one of the strongest chemical bonds known (6.9 eV), can be produced efficiently, and captured and cooled in a magneto-optical trap using any Q-line of its strong A1Π,v=0X1Σ+,v=0 band near 227.5 nm. Subsequent to trapping and cooling on the strong AX transition, the molecules may be cooled to a few µ K on any (narrow) Q-line of the spin-forbidden a3Π1,v=0X1Σ+,v=0 band. To laser-cool AlF successfully, it is essential to study the detailed energy level structure of the molecule, to measure the radiative lifetime of the states involved and to investigate and quantify the potential decay channels to states that are not coupled by the cooling laser, i.e. losses from the optical cycle.

For nearly a century, spectroscopists have been interested in the AlF molecule, in part due to its similarity to the much-studied CO and N2 molecules. AlF is also important to the astrophysics community and it has been detected in sunspots, stellar atmospheres, circumstellar envelopes and proto-planetary nebulae [Citation4, Citation5]. Moreover, 26AlF was the first radioactive molecule to be discovered in space [Citation6]. AlF has been the subject of theoretical studies using ab initio quantum chemistry to determine radiative lifetimes, dipole moments and potential energy curves for its electronic states [Citation7–10]. Precise spectroscopic parameters for AlF are useful for future astrophysical observations and new spectroscopic studies of electronic states can serve as a benchmark for quantum chemistry calculations.

In this paper, we present rotationally resolved optical spectra of the b3Σ+,v=0a3Π,v=0 band and demonstrate state-selective ionisation of a3Π molecules via (1 + 2)-resonance-enhanced multi-photon ionisation (REMPI). Pulsed excitation followed by delayed ionisation with a KrF excimer laser is used to determine the radiative lifetime of the b3Σ+,v=0 state. Next, high-resolution, cw laser-induced fluorescence (LIF) spectra of the b3Σ+,v=0X1Σ+,v=1 band are presented. The hyperfine structure in the b3Σ+,v=0 state is resolved and precise spectroscopic constants for the b state are determined. Laser-induced fluorescence spectra, recorded using a cw laser to drive rotational lines of the b3Σ+,v=0a3Π,v=0 band, allow us to reduce the uncertainty in the spin–orbit constant A and the spin–spin interaction constant λ of the a3Π state by nearly two orders of magnitude. A fraction of the excited state molecules decay on the b3Σ+,v=0X1Σ+,v bands to the ground state. This is caused by the spin–orbit coupling of the b3Σ+,v=0 state with the nearby A1Π state. This perturbation of the A1Π state and the b3Σ+ state is analysed in detail as it also induces a small loss channel for the strong AX cooling transition.

2. Previous work

In 1953, Rowlinson and Barrow reported the first observation of transitions between triplet states in AlF by recording emission spectra from a hollow-cathode discharge [Citation11]. In the following decades, new triplet states were identified and characterised in more detail [Citation12–16]. Kopp and Barrow analysed the interaction of the A1Π state with the nearby b3Σ+ state and thereby determined the term energy of the triplet states relative to the singlet states [Citation16]. A comprehensive study of the electronic states and the interaction between the singlet and triplet states followed [Citation17]. In 1976, Rosenwaks et al. observed the a3ΠX1Σ+ transition directly in emission [Citation18] at the same time as Kopp et al. did in absorption [Citation19]. Both studies confirmed the singlet-triplet separation determined from the earlier perturbation analysis.

The hyperfine structure of the rotational states in b3Σ+ was partly resolved and analysed by Barrow et al. [Citation17]. Their low-resolution spectra (typical linewidth of 0.05 cm1) showed a triplet structure, which they ascribed to magnetic hyperfine effects of the 27Al nucleus, whose nuclear spin is IAl=5/2. A more detailed analysis of both the fine and hyperfine structure in the triplet states was presented by Brown et al. in 1978 [Citation20]. The hyperfine structure was partially resolved and they determined values for the Fermi contact parameter bF(Al), the electron spin–spin interaction parameter λ, and the spin–rotation parameter γ.

We have recently reported on the detailed energy level structure of the X1Σ+, A1Π and a3Π states in AlF [Citation3]. That study resolves the complete hyperfine structure, including the contribution due to fluorine nuclear spin, in all three states. The energy levels in X1Σ+ and in the three Ω manifolds of the metastable a3Π state were measured with kHz resolution; this allowed us to determine the spectroscopic constants precisely and study subtle effects, such as a spin–orbit correction to the nuclear quadrupole interaction of the Al nucleus.

3. Rotational structure of the b3Σ+,v=0a3Π1,v=0 band and the radiative lifetime of the b3Σ+,v=0 state

To study the ba transition, the molecules must first be prepared in the a3Π state. Previously, we described this state and the aX transition in detail [Citation3]. The hyperfine structure in X1Σ+ is small and for the purpose of this study, the ground state is comprised of single rotational levels with energies given by BJ(J+1).

The angular momentum coupling in the a3Π state of AlF is well-described by Hund's case (a). The energy levels are labelled by the total angular momentum (excluding hyperfine interaction) J=R+L+S, with R the rotational angular momentum of the rigid nuclear framework, and L and S the total orbital and spin angular momenta of the electrons, respectively. Both L and S are not well defined. However, their projection onto the internuclear axis Λ and Σ and the total electronic angular momentum along the internuclear axis Ω=Λ+Σ are well defined. The spin–orbit interaction leads to three fine structure states with Ω=|Λ+Σ|, labelled with F1, F2 and F3, in order of increasing energy, corresponding to the states a3Π0, a3Π1 and a3Π2. The splitting between the Ω manifolds is determined by A, the electron spin–orbit coupling constant and the spin–spin interaction coupling constant λ. Each Ω has its own rotational manifold with energies BJ(J+1), where the rotational constant B is slightly different for each Ω manifold.

The b3Σ+ state is well-described by Hund's case (b), for which L couples to the rotation R to form N=R+L. The rotational energy levels follow BN(N+1). For Σ electronic states N=R and the total angular momentum (without hyperfine interaction) is J=N+S. The combined effect of the spin–rotation interaction γ(NS) and the spin–spin interaction 23λ(3Sz2S2) splits each rotational level N>0 of b3Σ+ into three J-levels; these components are labelled F1, F2 and F3 with quantum numbers J = N + 1, J = N and J = N−1, respectively.

Following the convention of Brown et al. [Citation21] the parity states can be labelled by e and f, where e levels have parity +(1)J and f labels have parity (1)J. All rotational levels of the X1Σ+ state are e-levels, while in the b3Σ+ state, all F2 levels are e-levels, while F1 and F3 are f-levels. In the a3Π state, each J-level has an e and an f component due to Λ-doubling.

Figure (a) shows the energy level diagram of the electronic states relevant to this study, together with a rotationally resolved spectrum of the b3Σ+,v=0a3Π1,v=0 band in (b), and a sketch of the experimental setup in (c), which is similar to the one reported previously [Citation3]. The molecules are produced by laser-ablating an aluminium rod in a supersonic expansion of 2% SF6 seeded in Ne. After passing through a skimmer, the ground-state molecules are optically pumped to the metastable a3Π1,v=0 state by a frequency-doubled pulsed dye laser using the Q-branch of the a3Π1,v=0X1Σ+,v=0 band. For this, 367 nm radiation with a bandwidth of 0.1 cm1 and pulse energy of 6 mJ in a beam with a e2 waist radius of about 2 mm is used. The Q-branch of this transition falls within this bandwidth. Therefore, many rotational levels in the metastable a3Π1,v=0 state are populated simultaneously. Via the Q-branch, only the f-levels in a3Π1 are populated. Alternatively, if the molecules are optically pumped to the metastable state using spectrally isolated R or P lines, only the e-levels are populated. Further downstream, at z = 55 cm from the source, the molecular beam is intersected with light from a second pulsed dye laser tuned to the b3Σ+,v=0a3Π1,v=0 transition near 569 nm. For pulse energies exceeding 6 mJ (unfocused, with an e2 waist radius of about 5 mm), this laser transfers population to the b3Σ+,v=0 state and subsequently ionises the molecules by having them absorb two more photons from the same laser. Such a one-colour (1 + 2)-REMPI scheme using an unfocused laser beam is very uncommon. However, AlF has numerous electronically excited states that lie one photon-energy above the b state energy, strongly enhancing the non-resonant two-photon ionisation probability. The ions are mass-selected in a short time-of-flight mass spectrometer (TOF-MS) and detected using micro-channel plates. The TOF-MS voltages are switched on shortly after the ionisation laser fires; this way ionisation occurs under field-free conditions and the states have a well-defined parity. The (1 + 2)-REMPI scheme uses low-energy photons and an unfocused laser beam, which has the benefit of producing a mass spectrum with only a single peak, corresponding to AlF. The spectrum, displayed in Figure (b), shows the ion signal as a function of the REMPI laser frequency. The spectral lines are labelled by Δ(NJ)F(J), where Δ(NJ)=NJ, as the quantum number J is not well-defined in the b state, vide infra. Since only the f-levels of the a3Π state are populated, the spectrum consists of Δ(NJ)=2,0,+2, i.e. O, Q and S branches.

Figure 1. (a) Electronic energy level scheme of the relevant states of AlF. The transitions used for laser excitation are shown as solid arrows. The laser-induced fluorescence used for detecting the molecules is indicated by downward wavy arrows. The indicated energies are the gravity centres of the respective states in the absence of hyperfine structure. (b) (1 + 2)-REMPI spectrum of the b3Σ+, v=0a3Π1, v=0 band. The a3Π1, v=0 state is populated via the Q-branch of the a3Π1, v=0X1Σ+, v=0 band using a frequency-doubled pulsed dye laser. (c) Schematic of the experimental setup used for the determination of the lifetime of the b3Σ+, v=0 state.

Figure 1. (a) Electronic energy level scheme of the relevant states of AlF. The transitions used for laser excitation are shown as solid arrows. The laser-induced fluorescence used for detecting the molecules is indicated by downward wavy arrows. The indicated energies are the gravity centres of the respective states in the absence of hyperfine structure. (b) (1 + 2)-REMPI spectrum of the b3Σ+, v′=0←a3Π1, v″=0 band. The a3Π1, v=0 state is populated via the Q-branch of the a3Π1, v′=0←X1Σ+, v″=0 band using a frequency-doubled pulsed dye laser. (c) Schematic of the experimental setup used for the determination of the lifetime of the b3Σ+, v=0 state.

To determine the radiative lifetime of the b3Σ+,v=0 state, we reduce the pulse energy of the ba excitation laser to about 1 mJ. At pulse energies below 2 mJ, the a3Π1 molecules are excited to the b state, but are not ionised via (1 + 2)-REMPI. Instead, a KrF excimer laser is used to ionise the b state molecules with a single 248 nm photon and pulse energy of about 3 mJ. The radiative lifetime of the b3Σ+, v=0 state is measured by varying the time-delay between excitation and ionisation [Citation22]. The determination of radiative lifetimes in the range of 20 ns to 1 μs is relatively straightforward, because it is longer than the laser pulse duration, but short enough so that the molecules do not leave the detection region. Figure  shows the ion signal (black dots) as a function of the time-delay with 1σ standard error bars. The blue line is a fit to the data using the model S(t)=Cet/τberfc[(t0t)/(2σ)], where erfc is the complementary error function, C and t0 are fit parameters, τb is the lifetime of the b3Σ+,v=0 state and σ is the measured, combined pulse-width of the excitation and ionisation laser. In this model, the lifetime τb=190(2) ns is fixed to the value determined from a linear fit to the semi-log plot shown in the inset. In 1988, Langhoff et al. calculated an approximate radiative lifetime of the b3Σ+ state of 135 ns [Citation8], considerably shorter than the measured lifetime.

Figure 2. Measurement of the radiative lifetime of the b3Σ+, v=0 state. First, the ground-state molecules are optically pumped to the b3Σ+,v=0 state via the two-colour excitation scheme described in the text. The population in the b3Σ+, v=0 state is probed by single-photon ionisation using a KrF excimer laser, followed by TOF-MS detection. The black dots show the AlF+ ion signal as a function of the time-delay between excitation and ionisation. The blue line is a fit to the data using the model described in the text. The inset shows a semi-log plot of a second measurement for eight specific time delays. A linear fit to the data gives a b state lifetime of 190(2) ns (colour online only).

Figure 2. Measurement of the radiative lifetime of the b3Σ+, v=0 state. First, the ground-state molecules are optically pumped to the b3Σ+,v=0 state via the two-colour excitation scheme described in the text. The population in the b3Σ+, v=0 state is probed by single-photon ionisation using a KrF excimer laser, followed by TOF-MS detection. The black dots show the AlF+ ion signal as a function of the time-delay between excitation and ionisation. The blue line is a fit to the data using the model described in the text. The inset shows a semi-log plot of a second measurement for eight specific time delays. A linear fit to the data gives a b state lifetime of 190(2) ns (colour online only).

4. The fine and hyperfine structure of the b3Σ+ state

Aluminium and fluorine have a nuclear spin of IAl=5/2 and IF=1/2, respectively. Following the description of Brown et al. [Citation20] and including the magnetic interaction of the fluorine nuclear spin, the effective Hamiltonian reads (1) Heff=23λ3Sz2S2+γNS+bF(Al)IAlS+13c(Al)(3IAl,zSzIAlS)+bF(F)IFS+13c(F)(3IF,zSzIFS)+eq0Q4IAl(2IAl1)(3IAl,z2IAl2),(1) where λ is the spin–spin interaction constant and γ the spin–rotation interaction constant. The parameters bF(Al) and bF(F) describe the Fermi contact interaction for the aluminium and fluorine nucleus, respectively, c(Al) and c(F) the dipolar interaction, and eq0Q the electric quadrupole interaction of the Al nucleus.

In the b3Σ+ state of AlF, the Fermi contact interaction bF(Al)(IAlS) between the nuclear spin of aluminium and the electronic spin angular momentum is strong compared to the spin–rotation interaction [Citation20]. The coupling case approximates (bβS). The N = 0 level has only one spin-component for which J = N + S = 1; this J = 1 level is split into three components due to the aluminium nuclear spin, each of which is again split by the nuclear spin of fluorine. This results in a total of six F levels. For N>0, J is not well defined and it is useful to introduce an intermediate angular momentum G=IAl+S (see Figure ). G then couples to N which results in three sets of sub-levels, with quantum numbers G=3/2,5/2 and 7/2, each of which contains the closely spaced stacks of N+3/2,,N3/2, N+5/2,,N5/2 and N+7/2,N+5/2,,N7/2 levels. Each of these levels is again split by the interaction with the fluorine nuclear spin to give the total angular momentum F=N+G+IF. For N = 1, 2 and 3 this results in a total of 18, 28 and 34 F levels, respectively. For N>3, the number of F levels reaches its limit of 36.

Figure 3. Relative weights WG and WJ of the |N,G,F and |N,J,F basis wave functions in the eigenfunctions |E of the energy levels in the three lowest rotational states of the b3Σ+, v=0 state. The weights are obtained by projecting the eigenstate of the Hamiltonian onto the basis wave function |N,G,F and |N,J,F, respectively. G can take the values 3/2, 5/2 and 7/2 and J can be N1, N or N + 1. The energy levels are described uniquely by the quantum numbers F, p and the index n. This analysis shows that G is a good quantum number for the description of any rotational level of the b3Σ+, v=0 state, while J only characterises the levels well for N = 0.

Figure 3. Relative weights WG and WJ of the |N,G,F⟩ and |N,J,F⟩ basis wave functions in the eigenfunctions |E⟩ of the energy levels in the three lowest rotational states of the b3Σ+, v=0 state. The weights are obtained by projecting the eigenstate of the Hamiltonian onto the basis wave function |N,G,F⟩ and |N,J,F⟩, respectively. G can take the values 3/2, 5/2 and 7/2 and J can be N−1, N or N + 1. The energy levels are described uniquely by the quantum numbers F, p and the index n. This analysis shows that G is a good quantum number for the description of any rotational level of the b3Σ+, v=0 state, while J only characterises the levels well for N = 0.

To determine the hyperfine energy levels of the b3Σ+ state, we drive the b3Σ+, v=0X1Σ+,v=1 band near 227.2 nm with a cw laser and detect the ba fluorescence at 569 nm. The spectrum of this transition directly reflects the energy level structure in the b3Σ+ state, because the hyperfine structure of the X1Σ+ state is smaller than the residual Doppler broadening in the molecular beam. The wavelength required to drive this transition is close to that of the A1Π,v=0X1Σ,v=0 band near 227.5 nm, for which we have a powerful UV laser system installed.

The b3Σ+,v=0X1Σ+,v=1 transition is spin-forbidden and the overlap of the vibrational wave functions is poor, with a calculated Franck–Condon factor of 0.02. However, the transition becomes weakly allowed due to the spin–orbit interaction between the b3Σ+ and the nearby A1Π state. In Section 7, this interaction will be discussed in more detail. To compensate for the weak transition dipole moment, we use a high excitation laser intensity of up to 75 mW in a laser beam with an e2 waist radius of 0.6 mm. The resulting laser-induced fluorescence occurs mainly on the dipole-allowed ba transition near 569 nm. The far off-resonant fluorescence allows us to block scattered laser light with a bandpass filter and to record background-free spectra. To increase the number of molecules in X1Σ+,v=1, we use a cryogenic helium buffer gas source, instead of the supersonic molecular beam introduced in the previous section. Figure  shows a sketch of the experimental setup that is used for this measurement.

Figure 4. Schematic of the experimental setup used to measure the fine and hyperfine structure of the b3Σ+,v=0 state. AlF molecules are produced in a cryogenic helium buffer gas source. The molecular beam is intersected with UV laser light from a frequency-quadrupled cw titanium sapphire laser. The molecules are excited on the weak, spin-forbidden b3Σ+,v=0X1Σ+,v=1 transition. The laser-induced fluorescence occurs mainly on the b3Σ+,v=0a3Π,v=0 transition near 569 nm and is imaged onto a PMT.

Figure 4. Schematic of the experimental setup used to measure the fine and hyperfine structure of the b3Σ+,v=0 state. AlF molecules are produced in a cryogenic helium buffer gas source. The molecular beam is intersected with UV laser light from a frequency-quadrupled cw titanium sapphire laser. The molecules are excited on the weak, spin-forbidden b3Σ+,v′=0←X1Σ+,v″=1 transition. The laser-induced fluorescence occurs mainly on the b3Σ+,v′=0→a3Π,v″=0 transition near 569 nm and is imaged onto a PMT.

The design of this source is similar to the one described in [Citation23–25]. A pulsed Nd:YAG laser ablates a solid aluminium target in the presence of a continuous flow of 0.01 sccm room-temperature SF6 gas, which is mixed with 1 sccm of cryogenic helium gas (2.7 K) inside a buffer gas cell. The hot Al atoms react with the SF6 and form hot AlF molecules, which are subsequently cooled through collisions with the cold helium atoms. Compared to the supersonic molecular beam, this source delivers over 100 times more AlF molecules per pulse in the rovibronic ground-state to the detection region. The forward velocity of the molecules is four times lower. To increase the number of molecules in X1Σ+,v=1, we increase the pulse energy and repetition rate of the ablation laser, increase the temperature of the SF6 gas to 350 K and increase its mass flow rate to 0.1 sccm. At z = 35 cm the molecules interact with cw UV laser light to drive the b3Σ+,v=0X1Σ+,v=1 transition near 227.2 nm. The laser light is produced by frequency-doubling the output of a cw titanium sapphire laser twice. The laser-induced fluorescence passes through an optical filter to block scattered light from the excitation laser and is imaged onto a photomultiplier tube (PMT). The photo-current is amplified and acquired by a computer. The wavelength of the excitation laser is recorded with an absolute accuracy of 120 MHz using a calibrated wavemeter (HighFinesse WS8-10).

Figure  shows the recorded spectra reaching the three lowest N levels in the b state. The panels demonstrate the increasing complexity of the hyperfine structure with increasing N. The three spectra allow us to measure the rotational constant of the b state. Gaussian lineshapes are fitted to the experimental spectra to determine the line-centres. We then fit the eigenvalues of the hyperfine Hamiltonian to the measured energy levels with the spectroscopic parameters as fit parameters. We assign a total of 48 lines and the standard deviation of the fit is 11 MHz. The best-fit parameters together with their standard deviations are summarised in Table . E0 is the pure vibronic energy of b3Σ+,v=0, i.e. the energy of the N = 0 level in the absence of spin, fine and hyperfine splitting. This is referenced to the J = 0 level of the X1Σ+,v=0 state by using the precise infrared emission lines of [Citation26] to determine the energy difference between the v = 0 and v = 1 level in the X state. The inverted spectra in Figure  are simulated spectra using the spectroscopic parameters presented in Table  and reproduce the measured spectra well. The Fermi contact parameter bF(F) for fluorine and the two hyperfine parameters c(Al) and c(F) for the aluminium and fluorine nucleus, respectively, are determined for the first time. In previous, low-resolution studies, the interaction of the fluorine nuclear spin has been neglected. However, we conclude that the magnitude of the interaction parameter for the two nuclei is comparable. This indicates that there is a significant electron density at both nuclei, which is in stark contrast to the situation in the a3Π state. In the latter state, the Fermi contact term for the 19F nucleus is about seven times smaller than for the 27Al nucleus [Citation3]. The Fermi contact parameter for the aluminium nucleus bF(Al) and the spin–spin interaction parameter λ presented here are consistent with the previously determined values, but their uncertainty is reduced significantly. Table lists the energies of the relevant hyperfine levels in b3Σ+,v=0 with their respective magnetic g-factors. The weights of the |N,G,F and |N,J,F basis wave functions in the eigenfunctions of these energy levels are shown in Figure .

5. The b3Σ+,v=0a3Π,v=0 transition

The ba bands have a diagonal Franck–Condon matrix and the b3Σ+,v=0a3Π,v=0 band has a Franck–Condon factor of 0.994. Its natural linewidth is 100 times smaller than the natural linewidth of the strong A1Π,v=0X1Σ+,v=0 transition near 227.5 nm. Laser cooling AlF molecules on the ba transition could therefore reach temperatures far below the Doppler limit of the strong AX transition. The vibrational branching to a3Π,v=1 is small and, if addressed with a repump laser, a molecule could scatter about 1000 photons before being pumped into a3Π,v=2. However, the radiative decay from the b state to multiple J levels in all three spin–orbit manifolds of the a state is allowed. This results in a large number of rotational branches that must be addressed to close the optical cycle [Citation28]. In addition, the hyperfine structure in a is large compared to the linewidth of the transition. Both effects make laser cooling of AlF, using an optical transition in the triplet manifold very challenging. This is in stark contrast to the strong AX transition, for which all Q-lines are rotationally closed and for which all hyperfine levels of a given rotational level in the X state lie within the natural linewidth.

Here, we demonstrate that laser-induced fluorescence spectroscopy of the b3Σ+,v=0a3Π,v=0 transition can be used to efficiently detect a3Π molecules with hyperfine resolution. In this section, we show that this method works well to detect molecules in all three Ω manifolds to improve two important spectroscopic constants of the a3Π state: the spin–orbit (A) and spin–spin (λ) interaction parameter, which determines the relative spacing of the three Ω manifolds in a3Π. The effective fine structure Hamiltonian is (2) H=ALzSz+23λ(3Sz2S2).(2) The interval between the Ω=0 and Ω=1 manifolds is approximately A2λ, while the Ω=1 and Ω=2 manifolds are about A+2λ apart.

For this experiment, we use the supersonic molecular beam setup introduced in Section 3, with an additional LIF detector installed between the excitation region and the TOF-MS. After passing through the skimmer, the molecules are excited on a specific rotational line of the a3Π,v=0X1Σ+,v=0 band to one of the three spin–orbit manifolds, using a frequency-doubled pulsed dye amplifier (PDA), which is seeded by a cw titanium sapphire laser. About 30 cm further downstream, a cw ring dye laser intersects the molecular beam orthogonally and is scanned over a rotational line of the b3Σ+,v=0a3Π,v=0 band. The laser frequency is stabilised and scanned with respect to a frequency-stabilised and calibrated HeNe reference laser (SIOS SL 03), using a scanning transfer cavity. The wavelength is recorded with an absolute accuracy of 10 MHz using the wavemeter, calibrated by the same HeNe laser.

Figure  shows hyperfine-resolved LIF excitation spectra of the b3Σ+,v=0a3Π,v=0 band. Figure (a) shows the hyperfine spectrum of the b3Σ+,v=0,N=1a3Π0,v=0,J=0 lines. The positive parity Λ-doublet level of a3Π0,J=0 has only two hyperfine components with total angular momentum quantum numbers F = 2 and F = 3 that are only split by about 3 MHz, much smaller than the residual Doppler broadening in the molecular beam. Therefore, this ba spectrum directly reflects the energy level structure in the b state and is identical to the one of the bX spectrum shown in Figure (b).

Figure 5. Direct measurement of the fine and hyperfine structure of the three lowest rotational levels in the b state, i.e. for N=0 in panel (c), N=1, in panel (b) and N=2 in panel (a). The experimental data are shown in black, pointing up and the simulated spectra are shown in blue, pointing down. The spectra are centred at the gravity centre, i.e. the line position in the absence of hyperfine, spin--spin and spin--rotation interaction (colour online only).

Figure 5. Direct measurement of the fine and hyperfine structure of the three lowest rotational levels in the b state, i.e. for N′=0 in panel (c), N′=1, in panel (b) and N′=2 in panel (a). The experimental data are shown in black, pointing up and the simulated spectra are shown in blue, pointing down. The spectra are centred at the gravity centre, i.e. the line position in the absence of hyperfine, spin--spin and spin--rotation interaction (colour online only).

Figure 6. LIF excitation spectra of the ba transition. Panels (a) and (b) show spectra of the b3Σ+,v=0,N=1a3Π0,v=0,J=0 and b3Σ+,v=0,N=2a3Π1,v=0,J=1 transitions, respectively. The spectrum in the inset labelled UV is detected by recording the emission on the b3Σ+,v=0,N=2X1Σ+,v bands as a function of the excitation frequency. (c) Hyperfine-resolved spectrum of the R3(1) line of the a3Π,v=0X1Σ+,v=0 band, showing the large hyperfine splitting in a3Π2. (d) Part of the hyperfine resolved spectrum of the b3Σ+,v=0,N=1a3Π2,v=0,J=2 transition. The low (high) frequency part of this spectrum was recorded after excitation to the F = 4 (F = 5) component of the R3(1) line of the a3ΠX1Σ+ transition. (a)–(d) The relevant energy level scheme is displayed next to each spectrum. Since J is not well-defined in the b state, the b3Σ+a3Π transitions are labelled with Δ(NJ)F(J). The measured spectra are printed in black, while the simulated spectra are shown in blue and inverted (colour online only).

Figure 6. LIF excitation spectra of the b−a transition. Panels (a) and (b) show spectra of the b3Σ+,v′=0,N′=1←a3Π0,v″=0,J″=0 and b3Σ+,v′=0,N′=2←a3Π1,v″=0,J″=1 transitions, respectively. The spectrum in the inset labelled UV is detected by recording the emission on the b3Σ+,v=0,N=2→X1Σ+,v″ bands as a function of the excitation frequency. (c) Hyperfine-resolved spectrum of the R3(1) line of the a3Π,v′=0←X1Σ+,v″=0 band, showing the large hyperfine splitting in a3Π2. (d) Part of the hyperfine resolved spectrum of the b3Σ+,v′=0,N′=1←a3Π2,v″=0,J″=2 transition. The low (high) frequency part of this spectrum was recorded after excitation to the F = 4 (F = 5) component of the R3(1) line of the a3Π←X1Σ+ transition. (a)–(d) The relevant energy level scheme is displayed next to each spectrum. Since J is not well-defined in the b state, the b3Σ+←a3Π transitions are labelled with Δ(NJ)F′(J′). The measured spectra are printed in black, while the simulated spectra are shown in blue and inverted (colour online only).

The spectrum of the b3Σ+,v=0,N=2a3Π1,v=0,J=1 transition, presented in Figure (b), shows a richer structure, and no longer directly reflects the energy level structure in the b state. The total span of the hyperfine splitting in the a3Π1,v=0,J=1 level is about 500 MHz [Citation3], and contributes to the complexity of this spectrum.

Optical pumping to the a3Π2,v=0 state is challenging because the R3(1) transition of the a3Π2,v=0X1Σ+,v=0 band is about 1000 times weaker than the corresponding R2(1) line to the a3Π1 state. Only a small fraction of the ground-state molecules produced in the source is transferred to the a3Π2,v=0 level, even with a PDA pulse energy of about 10 mJ in beam with a e2 waist radius of 1.5 mm. Figure (c) shows the R3(1) line of the a3Π2,v=0X1Σ+,v=0 band, recorded by scanning the seed laser of the frequency-doubled PDA, followed by (1 + 2)-REMPI and TOF-MS detection. The hyperfine structure in a3Π2,v=0 is very large and it is possible to resolve six of the ten hyperfine levels using a narrow-band pulsed laser. This enables us to populate a specific hyperfine component of the a3Π2,v=0,J=2 level, which is then followed by cw excitation to the b state and LIF detection. Figure (d) shows two LIF spectra that originate from two different hyperfine states in a3Π2, indicated by the two colours. The green spectrum originates from the a3Π2,v=0,J=2,F=4 level and the grey spectrum from the a3Π2,v=0,J=2,F=5 level. The line-centres are determined by fitting Gaussians to the spectral lines.

The hyperfine structure and Λ-doubling of the a3Π state has been investigated extensively in our previous study and is known to kHz precision [Citation3]. However, the constants A and λ of the a3Π state could only be determined with an accuracy of about 200 MHz, due to the finite bandwidth of the pulsed laser and the lower accuracy of the wavemeter that was used in that study to measure the relative energy of the Ω manifolds. Here, we take advantage of the reduced linewidth in the cw LIF spectra of the ba transition, in combination with the increased accuracy of our new wavemeter, to improve this measurement and therefore the spectroscopic constants A and λ. By using the hyperfine constants for the a3Π state from reference [Citation3] in combination with the parameters for the b state, listed in Table , the ba spectra can be simulated, and new values for A and λ are derived. The simulated spectra, using the full Hamiltonian, including the hyperfine interactions in both states are shown in Figure  as blue, inverted curves. The uncertainty of the improved spectroscopic constants, listed in , is reduced by nearly two orders of magnitude compared to the ones presented previously [Citation3]. E0 is the term energy of the a state in the absence of rotation, fine and hyperfine structure, calculated by using the term energy of the b state determined in the previous section. This means that the gravity centre, i.e. the position of the a3Π1, v=0, J=1 level in the absence of hyperfine structure, relative to the X1Σ+, J=0 level is at 27255.1737(5) cm1.

Table 1. Experimentally determined spectroscopic constants of b3Σ+,v=0.

Table 2. Spectroscopic constants for the a3Π,v=0 state of AlF, determined from a fit to the spectra.

It is also possible to record the LIF excitation spectra by detecting the weak UV fluorescence, on the b3Σ+,v=0X1Σ+,v bands which occurs predominantly at 223.1 nm. Part of the spectrum displayed in Figure  (b) has also been recorded this way and is shown by the inset labelled ‘UV’. To measure the ratio of the emission occurring in the UV, relative to the emission occurring in the visible, we lock the excitation laser to the resonance indicated by the arrow in Figure  (b) and average the signal over 1000 shots. To distinguish the two wavelengths, we use two different PMTs: a UV sensitive one with a specified quantum efficiency of 0.35±0.05 at 223.1 nm and negligible sensitivity for wavelengths >350  nm and a second PMT with a specified quantum efficiency of 0.1±0.015 at 569 nm that is sensitive in the range of 200–800  nm. The quantum efficiencies of the PMTs are taken from the datasheet and are not calibrated further. However, three identical UV PMTs give the same photon count rate to within 15% which we take as the systematic uncertainty. Both PMTs are operated in photon-counting mode, counting the number of UV and visible photons, nuv and nvis, respectively. The visible PMT is combined with a bandpass interference filter to block the UV fluorescence and a small amount of phosphorescence on the a3Π,v=0X1Σ+,v=0 transition near 367 nm. The measured transmission of the filter at 569 nm is 0.57. We combine the slightly different transmission through the imaging optics, the different detector efficiencies and the filter transmission into the total detection efficiencies ηuv and ηvis. Including these values, we measure a ratio of (3) Rb=nuvnvisηvisηuv=(4.3±1.3)×103.(3) This is the result of multiple experimental runs with a statistical error that is significantly smaller than the uncertainty in the quantum efficiencies of the PMTs. An additional 20% systematic uncertainty is added to the total error budget because we do not correct for the spherical and chromatic aberration of the fluorescence detector. However, simulations using ray-tracing software show a typical difference in imaging efficiency for the two wavelengths of about 10%. The value for Rb is identical to the ratio of the Einstein A-coefficients for bX and ba emission, i.e. Rb = Ab,X/Ab,a. The specific level in the b3Σ+,v=0 state for which this ratio of Einstein A-coefficients is determined, is the positive parity level with N = 2, F = 4, n = 4 which is almost a pure f-level, with 42.3% F3 and 57.4% F1 character, and only 0.3% F2 character. The level is highlighted in bold in and Figure .

Table 3. Energies, E, of the hyperfine levels in b3Σ+,v=0, relative to the X1Σ+,v=0,J=0 level, magnetic g-factors gF, rotational quantum number N, total angular momentum F and parity p.

6. The A1Πa3Π bands

To probe the amount of triplet wave function that is mixed into the A1Π state directly, we measure the ratio RA of the number of fluorescence photons emitted on the A1Πa3Π and on the A1ΠX1Σ+ transition. The value for RA is identical to the ratio of the Einstein A-coefficients for Aa and AX emission, i.e. RA = AA,a/AA,X. The value for RA gives the loss from the main laser cooling cycle due to electronic branching to the a3Π state. Previously, we measured this electronic branching ratio on the Aa bands indirectly to be at the 107 level, by comparing the absorption-strength of the A1Π,v=0a3Π0,v=0 transition relative to the absorption-strength of the A1Π,v=0X1Σ+,v=0 transition [Citation3]. In that analysis, we only took into account the amount of singlet character in the wave function of the a state to deduce the strength of the Aa transition. We did not account for the (then unknown) amount of triplet character in the wave function of the A state due to the interaction of the A and b states, which, as we will see here, turns out to be the dominant effect.

To measure such a small branching ratio directly, we use the buffer gas molecular beam source in combination with optical cycling on the Q(1) line of the AX transition. The small amount of visible fluorescence is isolated from the strong UV fluorescence by a high-reflectivity UV mirror, which is transparent in the visible, in combination with a long-pass and two bandpass filters in front of the PMT. The transmission band of the filters is chosen such that only wavelengths that cover the A1Π,v=0a3Π,v=0 transition are detected by the PMT. The mirror reduces the UV fluorescence that is incident on the spectral filters to the 106 level, suppressing their broadband phosphorescence when irradiated with UV light. The transmission of each optical element is measured individually at 227.5 nm and 599 nm, using laser light and a calibrated photodiode. The total detection efficiency, accounting for the transmittances and PMT responses becomes ηuv=0.2±0.05 for UV photons and ηvis=0.044±0.01 for visible photons in the range of 596–604  nm. The ratio of Einstein coefficients becomes (4) RA=nvisnuvηuvηvis=(6±2)×107,(4) where nvis and nuv are the number of photons detected in the visible and UV, respectively. A typical measurement is presented in Figure . The molecules are optically pumped on the Q(1) line of the 0-0 band of the AX transition and the LIF is imaged and detected by two different PMTs. The majority of the fluorescence is emitted in the UV and imaged onto the UV sensitive PMT. The PMT is operated in current mode, which is converted into a voltage, amplified and read into the computer. We calibrate this PMT output voltage against the output of the PMT in photon-counting mode for low incident light intensities. The small fraction of the LIF that is emitted in the visible is shown as red dots. The two time-of-flight profiles are very similar, with the detected signal in the visible being (1.3±0.05)×107 of the emission in the UV. By accounting for the different detection efficiencies for the two wavelengths the measured ratio is as given in Equation (Equation4). The total uncertainty in this measurement is dominated by the systematic uncertainty in the quantum efficiency of the two PMTs and by the uncertainty in the imaging efficiency for the two wavelengths as described in Section 5.

Figure 7. The LIF emitted in the UV (227.5 nm, black) and VIS (599 nm, red) as a function of time when the UV laser frequency is locked to the Q(1) line of the A1Π,v=0X1Σ+,v=0 transition. The inset shows the configuration of the fluorescence detector used for the VIS experiment (colour online only).

Figure 7. The LIF emitted in the UV (227.5 nm, black) and VIS (599 nm, red) as a function of time when the UV laser frequency is locked to the Q(1) line of the A1Π,v′=0←X1Σ+,v″=0 transition. The inset shows the configuration of the fluorescence detector used for the VIS experiment (colour online only).

The measurement of RA is significantly more challenging than the measurement of Rb. This is because nuv is 107 times larger than nvis and the phosphorescence of the optical elements typically occurs red-shifted, i.e. in the wavelength range of the weak fluorescence in the visible. This is in contrast to the measurement of Rb, for which nuv is about 200 times smaller than nvis, and for which any background caused by the phosphorescence of the optical elements is absent.

We also measure the ratio of the visible to the UV fluorescence subsequent to excitation on the Q(1) line of the 1−1 band of the AX transition. Since the A1Π,v=1 level is energetically much closer to the b3Σ+,v=0 level, one might expect a significantly larger fraction of visible fluorescence. However, we find that the two ratios RA are equal to within the 15% uncertainty of the measurement.

7. Spin–orbit interaction between the A1Π and b3Σ+ states

The observation of the b3Σ+,v=0X1Σ+,v and of the A1Π,v=0a3Π,v intersystem bands reported here indicates that the wave function of the b state of AlF contains a fraction of singlet character and that the A state contains a fraction of triplet character. This is due to the spin–orbit interaction of the b state with the nearby A1Π state. Figure  shows the potential energy curves for the lowest singlet and triplet electronic states of AlF. The inset shows a more detailed view of the A1Π and b3Σ+ states with their vibrational levels indicated. For low vibrational quantum numbers, the vibrational level vb in the b state lies slightly above the vibrational level vA=vb+1 in the A state. The energy difference between vb and vA=vb+1 decreases with increasing vibrational quantum numbers until the levels have just crossed and are nearly degenerate at vA=6 (vb=5), leading to a large perturbation of the rotational energy levels. This interaction has already been analysed by Barrow et al. in 1974, who introduced the parameter H0(vA) to describe the effect of the spin–orbit interaction of specific pairs of vibrational levels, i.e. of vA=4,5 and 6 with vb=3,4 and 5, respectively [Citation17]. They used the observed perturbation to determine the energy of the triplet manifold relative to the singlet manifold with an accuracy of 0.05 cm1.

Figure 8. Potential energy curves for the lowest singlet and triplet electronic states of AlF, using precise Expanded Morse Oscillator (EMO) functions. We obtain these EMO potentials by fitting to the point-wise RKR potentials generated by LeRoy's program [Citation29] and adjust the parameters to predict the vibrational levels with a high accuracy of 0.05 cm1 (using the non-perturbed or de-perturbed values from the appendix from [Citation17] and from [Citation30]). These potentials are much more precise than the simple Morse model we use in Section 7, but do not allow us to treat the two electronic states independently to extract the spin–orbit interaction. Ab initio calculations indicate that the A1Π state has a barrier in the region marked with a flash, which cannot be reproduced with our EMO potentials. The inset shows a more detailed view of the A1Π and b3Σ+ potentials, with the vibrational levels indicated.

Figure 8. Potential energy curves for the lowest singlet and triplet electronic states of AlF, using precise Expanded Morse Oscillator (EMO) functions. We obtain these EMO potentials by fitting to the point-wise RKR potentials generated by LeRoy's program [Citation29] and adjust the parameters to predict the vibrational levels with a high accuracy of 0.05 cm−1 (using the non-perturbed or de-perturbed values from the appendix from [Citation17] and from [Citation30]). These potentials are much more precise than the simple Morse model we use in Section 7, but do not allow us to treat the two electronic states independently to extract the spin–orbit interaction. Ab initio calculations indicate that the A1Π state has a barrier in the region marked with a flash, which cannot be reproduced with our EMO potentials. The inset shows a more detailed view of the A1Π and b3Σ+ potentials, with the vibrational levels indicated.

In this section, we give a more general description of the spin–orbit interaction between the A1Π,vA state and the b3Σ+,vb state. We deduce the expressions for Rb and RA in terms of the spin–orbit coupling constant Aso between the A and the b state, to experimentally determine the value for Aso. This is then compared to the value of Aso that can be deduced from the H0(vA) parameters as given by Barrow and co-workers [Citation17]. The effect of the interaction on the rotational energy levels in the A1Π,v=0 and b3Σ+,v=0 states is also discussed.

For the spin–orbit interaction of a 1Π state with a 3Σ+ state, the interaction terms between the e- and f-levels belonging to a given J are given by [Citation31]: (5) H(1Π,f|3Σ+,F1)=J+12J+1ξ2,(5) (6) H(1Π,e|3Σ+,F2)=ξ2,(6) (7) H(1Π,f|3Σ+,F3)=J2J+1ξ2.(7) In the Born–Oppenheimer approximation, the parameter ξ only depends on the vibrational wave functions of the coupled states and can be written in terms of the spin–orbit operator Hso as (8) ξ(vA,vb)=ΨA,vA|Hso|Ψb,vb=Asoφb,vb(ρ)φA,vA(ρ)dρ=AsoqvAvb.(8) The expressions φb,vb(ρ) and φA,vA(ρ) are the vibrational wave functions for the b and A state, respectively, and ρ is the inter-nuclear distance between the Al and F atoms. The square of the expression for the integral is the Franck–Condon factor qvAvb between the A,vA and b,vb levels. The Franck–Condon matrix between the A and b states is very diagonal. The value of q00 is very close to one and even though the v = 0 levels of both states are about (Eb,0EA,0) = 855 cm1 apart, the interaction between these levels dominates. The value of q10 is much smaller than one, but as the vb=0 and vA=1 levels are only (Eb,0EA,1) = 63 cm1 apart, this interaction also has to be taken into account. The interaction with all the other vibrational levels can be neglected.

7.1. Intensities of the intersystem bands

The spin–orbit interaction between the A state and the b state mixes their wave functions and causes a shift of the rotational levels. If the interacting levels are much further apart than the magnitude of the interaction terms, we can calculate the effect of the interaction using first-order perturbation theory. In this case, the contribution to the wave function of the A1Π,v=0 (b3Σ+,v=0) state due to the interaction with the b (A) state is given by the interaction term divided by the energy separation of the interacting levels. The wave function of the A state can be written as (9) |ΨA,0=CA,0s|ΨA,0+CA,0;b,0t|Ψb,0,(9) where |CA,0s|1 is the total amount of singlet character and where |CA,0;b,0t||CA,0t| is the total amount of triplet character in the wave function of the A state. It is readily seen from the interaction terms given above that the triplet contribution to the wave functions of the e- and f-levels is equal, does not depend on J, and is given by (10) |CA,0t|=Asoq002(Eb,0EA,0).(10) The wave function of the b state can be written as (11) |Ψb,0=Cb,0t|Ψb,0+Cb,0;A,0s(J,Fi)|ΨA,0+Cb,0;A,1s(J,Fi)|ΨA,1,(11) where |Cb,0t|1 is the total amount of triplet character and where |Cb,0;A,0s(J,Fi)|2+|Cb,0;A,1s(J,Fi)|2|Cb,0s(J,Fi)| is the total amount of singlet character in the wave function of the b state. The singlet contribution to the wave function of the b state depends on the (J,Fi) (i = 1, 2, 3) level. For the positive parity, N = 2, F = 4, n = 4 level that we used for the measurement of Rb we find, using the weights WJ as indicated in Figure  (W1J=42.3%, W2J=0.3% and W3J=57.4%), (12) |Cb,0s|=0.69Aso2q00(Eb,0EA,0)2+q10(Eb,0EA,1)2.(12) The expression for RA can now be rewritten as (13) RA=AA,aAA,X=|ΨA,0|μ(r)|Ψa,0|2λA,X3|ΨA,0|μ(r)|ΨX,0|2λA,a3=|CA,0t|2Ab,aλb,a3AA,XλA,a3=|CA,0t|2116,(13) where μ(r) is the electronic transition dipole moment, λb,a=569 nm and λA,a=599 nm are the wavelengths of the ba and Aa transitions, respectively. The ratio of the Einstein A-coefficients is calculated from the experimentally known lifetimes of the A state (1.90 ns, [Citation3]) and of the b state (190 ns, Section 3). The expression for Rb can now be rewritten as (14) Rb=Ab,XAb,a=i=01|Ψb,0|μ(r)|ΨX,i|2λb,a3|Ψb,0|μ(r)|Ψa,0|2λb,X3=|Cb,0s|2AA,XλA,X3Ab,aλb,X3=106|Cb,0s|2,(14) where λA,X = 227.5 nm and λb,X = 223 nm. The sum in the numerator results in the two terms given in Equation (Equation12), because the Franck–Condon matrix between the A and X states is highly diagonal.

The parameter H0(vA), introduced by Barrow and co-workers, describes the effect of the spin–orbit interaction between two specific vibrational levels and is equivalent to (15) H0(vA)=ξ(vA,vb=vA1)2=AsoqvAvA12.(15) Barrow and co-workers did not discuss the dependence of the values for H0(vA) on the square root of the Franck–Condon factors, and they therefore did not extract a single value for the spin–orbit interaction parameter Aso from the three values of H0(vA) that they reported [Citation17].

In the following, we determine the Franck–Condon factors qvAvb and Aso from the measured H0(vA) values [Citation17], and then use these Franck–Condon factors to extract a value for Aso from our measurements of RA and Rb. For this, a Morse potential is fitted to the term-values of the vibrational levels in the A and b state, as listed in [Citation17]. The parameters of the Morse potential are optimised to reproduce the measured vibrational levels to better than 0.3 cm1. This optimisation is independent of the equilibrium distance re. Next, the difference between the equilibrium distances of the A and b state, Δre=re(A)re(b), is optimised such that the vibrational level dependence of qvAvA1 agrees with the observed vibrational level dependence of H0(vA). We find the best agreement for Δre=0.0046Å, about half the value for Δre extracted from the reported values for Be in the A and b state [Citation17] of 0.0094 Å. Finally, we fit to the experimentally determined rotational constants reported in [Citation17] using only re(A) as a fit parameter. The data is reproduced to within 1% for re(A) = 1.63098 Å. Considering the simple model for the potentials, this is an excellent agreement. The value for Aso determined from this fitting procedure is Aso=8.9 cm1 and the corresponding Franck–Condon matrix is shown in Table . The uncertainty in the value of Aso is difficult to determine, as the main contribution to the total uncertainty stems from the assumption that the potentials can be approximated by Morse potentials. We estimate this uncertainty to be at least 2.5 cm1.

Table 4. Calculated Franck–Condon factors between the A1Π state and the b3Σ+ state.

Using the information from the Franck–Condon matrix shown in Table , the experimental value of RA=(6±2)×107 implies a value for Aso=(10±2) cm1. The value of Rb=(4.3±1.3)×103 implies a value for Aso=(10.4±1.5) cm1. These two values for Aso, as well as the value for Aso extracted from the H0(vA) values, all overlap within their error bars, yielding a final experimental value for Aso=(10.0±1.1) cm1.

As mentioned in the previous section, we observe that the ratio of the visible to the UV fluorescence that is emitted from the A1Π,v=1 level is equal to that from the A1Π,v=0 level. Based on the derivation presented in this section, we expect that the fractional emission in the visible from the A1Π,v=1 level is larger than the emission from the A1Π,v=0 level by a factor of (16) Eb,0EA,0Eb,1EA,12q11q00+Eb,0EA,0Eb,0EA,12q10q00=1.23,(16) where the energy difference (Eb,1EA,1)=834 cm1, and where the values for the Franck–Condon factors are taken from Table . The predicted 23% increase in visible fluorescence is consistent with the experimental observation, within the experimental uncertainty of 15%. This highlights that the spin–orbit mixing is dominated by the interaction between vibrational levels that have the same vibrational quantum number.

7.2. Effect on the fine structure in the b3Σ+,v=0 state

In the first-order perturbation theory, the shift of the energy levels is given by the square of the interaction matrix elements, divided by the energy separation of the interacting levels. The interaction is repulsive, and for the A1Π,v=0 state all rotational levels shift downwards by Aso2q00/[2(Eb,0EA,0)]. Such an overall shift is difficult to determine and is absorbed in the value for the term-energy. In the b3Σ+,v=0 state, the e-levels will be shifted upward by about the same amount, but the f-levels will have a lower, J-dependent shift. ‘Curiously’, Hebb wrote originally in 1936, the shift of the levels in a 3Σ+ state due to the spin–orbit interaction with a 1Π state has the same J-dependence as the shift due to the spin–spin and spin–rotation interaction, and both effects cannot be distinguished [Citation32]. The origin of the spin–spin and spin–rotation interaction in a 3Σ state has first been described in a classic paper by Kramers [Citation33], and more general expressions have been given soon after that by Schlapp [Citation34]. Normally, the spin–spin and spin–rotation interactions are expected to be the dominant effects and the spin–orbit interaction with a nearby 1Π state is expected to be only a second-order correction. Both effects add up, and this means that the values for λ and γ as found from fitting the energy levels in the b3Σ+,v=0 state should actually be interpreted as (17) λ=λss+Aso2q004(Eb,0EA,0),(17) (18) γ=γsr+BAso2q002(Eb,0EA,0)2,(18) where λss and γsr describe the contribution due to the spin–spin and spin–rotation interaction in the b3Σ+ state, respectively, and where the additional terms describe the contribution due to the spin–orbit interaction with the nearby A1Π state.

When we take the final experimental value for Aso, then the spin–orbit contribution to λ amounts to about +900  MHz. We conclude therefore that the value for λss in the b3Σ+,v=0 state is about −1800  MHz, and that about half of this value is cancelled by the spin–orbit interaction with the nearby A1Π state. The spin–orbit contribution to γ amounts only to +1.1  MHz, and this contribution can be neglected. For higher vibrational levels in the b3Σ+ state a slightly different behaviour is expected. The spin–orbit contribution to λ remains about +900 MHz for vb = 0–3, but increases to about +1200 MHz for vb=4 due to the near-resonant interaction with the vA=5 level. For the vb=6 level, the near-resonant contribution to λ due to spin–orbit interaction with the vA=7 level is negative, reducing the total spin–orbit contribution to λ to about +600 MHz. The vb=5 level is special, as the F3 levels of the b-state and the f-levels of A,vA=6 cross, between J = 1 and J = 2, making an interpretation in terms of a contribution to λ less meaningful. It is interesting to note that this crossing causes the J-levels that belong to low N-values in the b-state to be split considerably further than for the b3Σ+,v=0 state, making J a good quantum number. This strong interaction opens a ‘doorway’ to efficiently drive transitions between the singlet and triplet manifolds [Citation35, Citation36] and when J is a good quantum number, this can be done highly rotationally selective.

8. Conclusion

We investigated the b3Σ+ state of AlF and the spin–orbit interaction between this lowest electronically excited state in the triplet system and the first electronically excited singlet state, the A1Π state. First, we presented a low-resolution rotational spectrum of the b3Σ+,v=0a3Π1,v=0 transition and determined the radiative lifetime of the b3Σ+,v=0 state to be 190(2) ns. Molecules in the a3Π,v=0 state can be efficiently detected using a (1 + 2)-REMPI scheme via the b3Σ+,v=0 state even at relatively low laser intensity. Then, we used cw laser-induced fluorescence excitation spectroscopy of the b3Σ+,v=0X1Σ+,v=1 transition to determine the fine and hyperfine structure of the b3Σ+,v=0 state with a precision of about 10 MHz. The eigenvalues of the hyperfine Hamiltonian have been fitted to the experimentally determined line positions and all relevant spectroscopic constants have been determined. Hyperfine-resolved LIF spectra of the b3Σ+,v=0a3Π,v=0 band, originating from all three spin–orbit manifolds in a3Π, were used to improve the spin–orbit (A) and spin–spin (λ) interaction parameters that determine the relative spacing of the three Ω manifolds in the a state. Despite the highly diagonal Franck–Condon matrix of the ba transition, laser cooling of AlF in the triplet manifold is challenging, due to the large number of rotational branches. This is further complicated by the spin–orbit interaction between the A1Π state and the b3Σ+ state, which is concluded to be governed by an interaction parameter Aso of about 10 cm1. The spin–orbit interaction mixes up to about 1% of the wave function of the A1Π state into the wave function of the hyperfine levels in the v = 0 level of the triplet state; the exact amount of mixing depends on the J,Fi (i = 1, 2, 3) character of the hyperfine levels in the b3Σ+ state. By the same mechanism, about 1% of the wave function of the b3Σ+ state is mixed into the wave functions of the A1Π,v=0 state; the amount of triplet character is the same for all levels in the A state. The triplet character of the A1Π,v=0 state, causes an A1Π,v=0a3Π,v=0 loss below the 106 level from the main A1ΠX1Σ+ laser cooling transition.

Disclosure statement

No potential conflict of interest was reported by the author(s).

References