Publication Cover
Molecular Physics
An International Journal at the Interface Between Chemistry and Physics
Latest Articles
124
Views
0
CrossRef citations to date
0
Altmetric
Research Article

Efficient transfer of population between rotational levels in deuterated ammonia using THz transitions

ORCID Icon, , ORCID Icon, ORCID Icon &
Article: e2353333 | Received 11 Mar 2024, Accepted 03 May 2024, Published online: 14 May 2024

Abstract

We demonstrate the efficient transfer of population between rotational levels of ND3 using a commercially available table-top THz source. We use a Stark decelerator to prepare a packet of para-ND3 molecules in the 11 upper inversion component of the rotational ground state, and induce the 21+11 electric dipole allowed transition both in the presence of an electric field and under field free conditions. Using THz chirps to address all hyperfine components, transfer efficiencies up to 87% are reached, limited by the power of the THz source. Population transfer from the 11 to the 21 level is demonstrated either by driving directly the 2111 transition in a field, or by a 1111+21 double-resonance excitation scheme. In addition, we present a novel method to produce packets of ND3 (11) with pure MJ=0 or |MJ|=1 character using the Stark decelerator, and a method to probe the MJ composition of a packet of molecules using THz excitation.

GRAPHICAL ABSTRACT

1. Introduction

The method of using coherent radiation to efficiently transfer population by employing the principle of adiabatic rapid passage (ARP) is well established. Some recent examples include the use of sub-THz radiation to pump significant amounts of population between different rotational levels of CO molecules on a chip [Citation1], the transfer between vibrational levels of CH4 in molecule-surface scattering experiments [Citation2], the transfer between different electronic states of trapped barium ions using a laser at 1.76μm [Citation3], and the use of microwave radiation in electron spin resonance experiments [Citation4]. ARP is a fundamental tool in quantum simulations to manipulate the state of qubits (see e.g. [Citation5, Citation6]) and theoretical investigations are still carried out to determine best performance schemes under various conditions [Citation7].

In a previous article we demonstrated the efficient transfer of population from the upper () to the lower (+) inversion level within the rotational ground state JKp=11± of para-ND3 using ARP employing microwave chirps at frequencies near 1.6 GHz [Citation8]. Despite the presence of a complicated 144 state hyperfine structure, population transfer with near unit efficiencies were achieved.

Here, we demonstrate the efficient transfer of population between rotational levels of ND3 using ARP and direct excitation with a commercially available THz source, similar to earlier work on driving pure rotational transitions in CO(a3Π) [Citation1]. We produce a sample of ND3 molecules in the 11 upper inversion doublet component using a Stark decelerator, and induce the 21+11 transition in ND3 at frequencies near 615 GHz. Despite the complicated hyperfine structure of ND3, up to 87% of the initial population can be transferred into the 21+ lower inversion component of the rotationally excited state, limited only by the power of our present THz source. In the presence of an electric field of a few kV/cm, we populate the 21+ level in specific MJ levels by driving MJ-resolved Stark-split transitions. We furthermore investigate routes to transfer population from the 11 state to specific MJ levels within the 21 upper inversion component, either by direct excitation using the 2111 transition in an electric field, or by a 1111+21 double resonance microwave-THz excitation scheme. Finally, we present novel methods to produce samples of ND3 (11) molecules with pure MJ=0 or |MJ|=1 character using the Stark decelerator, and describe methods to directly probe the MJ purity of a sample of ND3 molecules using THz excitations.

Our motivation in investigating these possibilities is mainly given by future low-energy crossed beam collision experiments using ND3 (or other molecules with a similar energy level structure). The ND3 molecule has been important in experiments using cold molecules for more than two decades [Citation9–13]. Interest in this species mainly stems from its ease in production and detection, its relatively large electric dipole moment, and its favourable energy level scheme featuring near-resonant rotational levels of opposite parity. The ability to prepare samples of molecules in (specific MJ components of) the 21± rotationally excited level before the collision, in combination with the ability to probe collision products MJ selectively, offers exciting prospects for detailed collision studies at low collision energies. First, molecular beams with large populations in 21± can be used in collisional de-excitation experiments. The pre-collision added quantum of angular momentum causes the partial wave(s) of the entrance channels to change during the collision. This partial wave dynamics is encoded in the differential cross-section (DCS), which can be probed using velocity map imaging (VMI) [Citation14]. Second, the 20 cm1 recoil energy of the 2111 de-excitation process increases the diameter of the scattering image. In current high-resolution experiments, this 20 cm1 recoil energy results in a scattering image with a sufficiently large radius to discern structure in velocity-mapped scattering images, even in the limit of near-zero collision energies [Citation14]. Third, the THz tagging pulse can be used to set a ‘time zero’ for the collisions, which is particularly important in merged beam configurations in which beams can overlap for extended periods of time [Citation15]. The ability to prepare either the 21+ or the 21 state as the initial level gives access to probing both parity-changing and parity-conserving rotational de-excitation collisions, which are governed by different parts of the interaction potential. Last but not least, there currently is large interest to study low-energy collisions in the presence of electric fields [Citation16]. At sufficiently low collision energies, the interaction energy with an external field can become larger than the collision energy itself, offering distinctive possibilities to actively manipulate interaction Hamiltonians and to steer collision cross sections. In an electric field, cross sections depend on the MJ quantum numbers of the initial and final rotational level. The ability to record fully state-resolved MJMJ cross sections is essential to study field effects in low-energy collisions in the required detail, and can reveal new aspects relating to the stereodynamics of the collision [Citation17]. The study and understanding of fully MJ-resolved collisions is also urgently needed to advance experiments that aim to confine (and further cool) ultracold molecules in traps, as inelastic MJ-changing collisions could lead to trap loss.

This article is organised as follows: after describing the experimental methods in Section 2, we will describe in Section 3.1 investigations on the general performance of the THz source with respect to resolution and ARP transfer using the dipole-allowed 21+11 transition in zero electric field. For measurements inside an electric field, we probe the homogeneity of the field of a capacitor by driving the 1111+ inversion doublet transition using microwaves at various field strengths (see Section 3.2). In Section 3.3, we will then add a bias electric field and investigate the MJ-resolved 21+11 transition. In Sections 3.4 and 3.5 we will discuss the possibilities to transfer ND3 molecules from the 11 to the 21 state, either directly inside an electric field or by a double resonance 1111+21 scheme that combines microwaves and THz capabilities. Finally, in Section 3.6 we will discuss how beams with pure MJ composition can be made using the Stark decelerator, and how rotational transitions using THz radiation can be used as an easy-to-implement tool to probe the MJ-resolved beam purity.

Throughout the different sections, an understanding of the Stark shift of the involved levels of ND3 is required. The Stark effect can be calculated from the total Hamiltonian matrix representation (1) Htotal=HvibrotμE,(1) which was obtained from the rovibrational Hamiltonian Hvibrot and the transition dipole matrix μ as calculated using Pickett's SPCAT [Citation18]. The results for 14ND3 are shown in Figure . The required rovibrational Hamiltonian was derived from semi-rigid rotor parameters as provided on the Cologne database for molecular spectroscopy (CDMS) [Citation19–21] and spin-spin, spin-rotation and quadrupole coupling parameters as provided in [Citation22, Citation23]. Throughout the Figures in this article we will use a colour convention depending on which states are probed: Blue: 11+, Cyan: 21+, Orange: 11, Red: 21. We note that although the scope of our experiments is primarily focussed on the rotational structure and associated quantum numbers J and MJ, the hyperfine structure of ND3 is partially resolved in our experiments. A detailed account of the Stark effect including hyperfine structure is given by van Veldhoven et al. [Citation22].

Figure 1. Stark curves for the 11±,MJ (lower panel) and 21±,MJ (upper panel) states of 14ND3.

Figure 1. Stark curves for the 11±,MJ (lower panel) and 21±,MJ (upper panel) states of 14ND3.

2. Experimental

The experimental setup is schematically shown in Figure , and is operated at a repetition rate of 10Hz. Details on the Stark decelerator and detector are described elsewhere [Citation24, Citation25]. A gas mixture of 5% ND3 in Xenon was expanded through a Nijmegen Pulsed Valve [Citation26] at typical backing pressures of 1bar into a vacuum, creating a rotationally cold molecular beam. The molecular beam was skimmed (skimmer radius 1.5mm) and passed through a Stark decelerator, operating at ±15kV in the s = 3 guiding mode around 400m/s. The ND3 molecules that exit the decelerator almost exclusively populate the 11 upper inversion level of the rotational ground state of para-ND3 [Citation8]. The detection region was placed 448mm downstream from the decelerator and consists of a VMI spectrometer utilising a state-selective 2 + 1 REMPI scheme via the v = 4 and v = 5 levels of the excited B electronic state [Citation27]. Signal was recorded either as voltage change, picked up on the microchannel plate (MCP) and passed to a Pico-Scope 5444D oscilloscope or visually, using a camera recording the light flashes of a phosphor screen.

Figure 2. Scaled down representation of the experimental setup showing the last section of the Stark decelerator (S), capacitor plates (C) to generate an electric field, microwave wire antenna (M), THz horn antenna (T), and velocity map imaging electrodes with laser used for resonance-enhanced multiphoton ionisation (REMPI-VMI).

Figure 2. Scaled down representation of the experimental setup showing the last section of the Stark decelerator (S), capacitor plates (C) to generate an electric field, microwave wire antenna (M), THz horn antenna (T), and velocity map imaging electrodes with laser used for resonance-enhanced multiphoton ionisation (REMPI-VMI).

In the region between the end of the decelerator and laser detection point, transfer of ND3 molecules from the 11 to either the 11+, 21 or 21+ can be induced using microwave or THz radiation. To study these transitions in an electric field, a 85mm long capacitor is placed 27mm downstream from the decelerator. This capacitor consisted of two aluminum plates separated by 10mm, with one operated at voltages up to 5kV and the other grounded. Microwaves were provided using a Rohde & Schwarz SMA103B signal generator with chirp functionality, as described in our recent paper [Citation8]. The signal generator was connected to a custom made λ/4 monopole antenna, positioned near the exit of the Stark decelerator. The resulting microwave radiation was not directional, effectively creating microwave fields in the entire vacuum chamber.

THz radiation was generated according to the circuit shown in Figure . The 10GHz output of a National Instruments QuickSyn Signal generator at 15dBm, was passed as ‘local oscillator’ (LO) into a mixer (Mini-Circuits ZX05-153MH-S+). The LO signal was mixed with an ‘intermediate frequency’ (IF) provided by a second SMA103B, which was varied around 2.8GHz and passed through a 10dB attenuator (to limit the possible power input) before entering the mixer. The mixer creates a ‘radio frequency’ (RF) signal composed of the difference signal at 7.2GHz and the sum signal at 12.8GHz. The mixed signal was then passed through a bandpass filter for 11.714.5GHz (Mini-Circuits ZVBP-13R1G-S+) for signal purification and the elimination of one side band. Keeping the QuickSyn generator running at all time, the control over output and pulse characteristics – such as chirp parameters – in this setup was provided by the SMA103B. The signal – typically around 12.8GHz – was then passed into a VDI (Virginia Diodes) amplifier multiplier chain (VDI, Model WR1.5AMC-615-5mWp) providing a multiplication factor of 48 and a typical output power of 5mW in a range from 605 to 625GHz.

Figure 3. Circuit diagram for the generation of THz radiation around 615GHz. Signals from a QuickSyn and SMA103B generator are mixed, passed through a bandpass filter, multiplied by a factor of 48 using a VDI chain, and are transmitted using a horn antenna.

Figure 3. Circuit diagram for the generation of THz radiation around 615GHz. Signals from a QuickSyn and SMA103B generator are mixed, passed through a bandpass filter, multiplied by a factor of 48 using a VDI chain, and are transmitted using a horn antenna.

The THz radiation was highly directional and passed either through fused silica viewports (on CF100 or CF40 flanges with a window thickness of 6.35mm and 4.04mm and a transmission of 50% and 72%, respectively) or through a special polymer (ZEONEX, obtained from TYDEX) window (CF40 – window thickness 4.04mm, transmission of 88%). Depending on the experiment a polymer (TPX) collimator lens with a focal length of 10cm was used (THORLABS TPX100). The THz horn antenna was positioned such that the THz beam intercepted the ND3 molecules within the capacitor, see Figure .

3. Results and discussion

3.1. 21+11 THz excitation in zero field

To test the resolution of our THz source and its ability to drive pure rotational transitions in ND3, we positioned the THz source relatively far away from the chamber (about 70cm) and worked with unfocused radiation passing over a mirror and through a 100CF fused silica viewport into the machine (transmission 50%). Single frequency pulses of 30μs duration were used, while the population in the 21+ state was probed using the laser. The recorded experimental THz spectrum is shown in the left panel of Figure , together with a simulation of the spectrum based on a low power approximation. The intensity in the spectrum is proportional to the population transferred into the probed 21+ state. The simulations were performed using the codes as described in Ref. [Citation8]. Intensities were predicted using Pickett's SPCAT [Citation18] program, while the linewidth (Full Width at Half Maximum (FWHM) =21kHz) was determined from the pulse duration t=30μs of a rectangular pulse, using the relation FWHM =1/(1.568t). As our experiment suffers from inaccuracies in absolute frequency due to the absence of a good frequency standard, the experimental spectrum was shifted by 30kHz to match the simulated spectrum. It is seen that the experimental spectrum is well reproduced by the simulations, and the hyperfine structure pertaining to the F1 quantum number is fully resolved. Note that all hyperfine levels that correlate to the MJ=0 and |MJ|=1 quantum numbers are populated. This is a consequence of the scrambling of MJ states in the zero field of the free flight section between the exit of the decelerator and interaction region, as discussed in detail in Ref. [Citation8].

Figure 4. Left: Hyperfine resolved spectrum for the 21+11 rotational transition in ND3 using single frequency pulses of 30μs duration. The F1F1 quantum numbers referring to coupled rotational angular momentum J and nitrogen spin IN by F1=J+IN are indicated. Right: Measured (orange solid lines) and predicted (orange dashed lines) depletion in 11 population (orange) and corrsponding derived gain in 21+ population (cyan) throughout a 40μs THz chirp covering 3.6MHz from 614,969.418MHz to 614,965.818MHz.

Figure 4. Left: Hyperfine resolved spectrum for the 21+←11− rotational transition in ND3 using single frequency pulses of 30μs duration. The F1′←F1″ quantum numbers referring to coupled rotational angular momentum J and nitrogen spin IN by F1→=J→+IN→ are indicated. Right: Measured (orange solid lines) and predicted (orange dashed lines) depletion in 11− population (orange) and corrsponding derived gain in 21+ population (cyan) throughout a 40μs THz chirp covering 3.6MHz from 614,969.418MHz to 614,965.818MHz.

To investigate the maximum population transfer efficiency that can be achieved, we moved the THz source closer to the window, collimated the beam using a TPX lens and radiated through a low loss ZEONEX polymer window. To transfer population from all hyperfine components, a 40μs chirp of 3.6MHz bandwidth that covers the complete hyperfine structure was applied. The chirp was stopped at different timings to monitor the development of the 11 population (with the 21+ population derived from it) and the experimental results are plotted together with predictions in the right panel of Figure . These predictions follow from our model that includes the full hyperfine structure of ND3 and numerical integration of the von Neumann equation as detailed in Ref. [Citation8]. It is seen that a maximum population transfer of 87% was achieved. In the simulations, we chose the power of the THz radiation to best match our experimental observations (3W/m2). However, increasing the power in the simulation would still lead to higher expected transfers (up to 97% at 12.8W/m2; data not shown). This suggests that the maximum transfer observed in our current experiments is limited by the power of the THz source, but that near-unit transfer efficiencies are inherently possible.

3.2. Calibration of electric field by 1111+ microwave transfer

In our previous study our capabilities to carry out MJ specific transitions were limited by working in inhomogeneous fringe fields [Citation8]. In the present study we use a dedicated electrostatic capacitor consisting of two parallel aluminum plates to create homogeneous fields. For the long wavelengths of the microwave radiation, the capacitor plates have the additional benefit to act as a polariser, only allowing radiation polarised parallel to the direction of the static electric field. This enforces a ΔMJ=0 selection rule for microwave transitions, greatly facilitating the interpretation of the spectra.

To calibrate the strength and test the homogeneity of the electric field of the capacitor, we simulated and measured the hyperfine resolved spectrum of the 11,|MJ|=111+,|MJ|=1 transition of 14ND3 using a voltage difference of 1.2kV applied to the capacitor plates. The results are shown in the left panel of Figure . From the spectrum it is evident that the field of the capacitor must be fairly homogeneous since the components of the hyperfine structure arising from the electric quadrupole interaction of the 14N nucleus can be resolved. Details arising from the deuterium nuclei are completely blurred out. The measured spectrum is reproduced well by the simulations, although the observed linewidth was only reproduced by assuming a reduced pulse duration of 9μs (compared to the rectangle pulse of 20μs as applied in the experiment). Such deviations can reflect significant inhomogeneities in the microwave field, such as a frequency dependence of the amplitude or a variation of amplitude over the distance travelled by the molecular packet. Small scale electric field inhomogeneities on the order of a few V/cm can also cause broadening. A slightly inhomogeneous microwave field can be expected, since the antenna is placed relatively far from the capacitor, and not in line of sight of the molecules. Despite this broadening on the kHz scale, the spectrum is consistent with a homogeneous electric field of 1.2kV/cm.

Figure 5. Left: Experimental (blue) and simulated (black) hyperfine resolved microwave spectrum for the 1111+ transition in 14ND3 in an electric field of 1.2 kV/cm. Right: Measured (orange solid lines) and predicted (orange dashed lines) depletion in 11 population and corresponding derived gain in 11+ population (blue) throughout a 20μs microwave chirp covering 5MHz from 1820.1MHz to 1825.1MHz.

Figure 5. Left: Experimental (blue) and simulated (black) hyperfine resolved microwave spectrum for the 11−→11+ transition in 14ND3 in an electric field of 1.2 kV/cm. Right: Measured (orange solid lines) and predicted (orange dashed lines) depletion in 11− population and corresponding derived gain in 11+ population (blue) throughout a 20μs microwave chirp covering 5MHz from 1820.1MHz to 1825.1MHz.

Similar to our previous study [Citation8], we tested the maximum population transfer to specific MJ final states that can be achieved in the electric field. For this purpose we used a 5MHz bandwidth 20μs microwave pulse centred at 1822.6MHz increasing frequency with time (up direction). We stopped the microwave chirps at different timings to measure the population transfer throughout the pulse, as is illustrated in the right panel of Figure . The orange line represents the 11 signal depletion, the blue line represents the derived increase in 11+ population, and the dashed lines symbolise simulated results. We measured a maximum population transfer of 93%, consistent with our earlier findings [Citation8] for 14ND3. Similar results, also with 93% efficiency, were achieved for 15ND3. The high efficiency for the 15ND3 population transfer is in contrast to the 77% efficiency achieved previously in zero field [Citation8]. We conclude that the preservation of pure |MJ| states caused by the presence of an electrical bias field and driving transitions exclusively between the |MJ|=1 states is beneficial in the context of transfer efficiency.

3.3. 21+11 THz excitation in electric fields

To test the ability to also drive pure rotational MJ-resolved transitions in the presence of an electric field with our THz source, we measured the 21+11 transition as a function of the electric field strength. The THz source was operated using pulses of varying bandwidth (dependent on chosen stepsize), while the spectrum was scanned downwards from the zero field transition frequency (see Figure (a)). The THz source was oriented in a way that the electric field vector of the THz radiation was perpendicular to the capacitor field, resulting in a ΔMJ=±1 selection rule. The obtained spectra for fields up to 3 kV/cm are shown in Figure . Lines are labelled according to the MJ quantum numbers in the 11 and 21+ levels, that formally only become good quantum numbers in fields above 1kV/cm. This is directly visible in the oberved spectra. At fields of 2 and 3 kV/cm, only transitions with ΔMJ=±1 are observed, consistent with the selection rules imposed by the geometry of the experiment. At lower fields, where MJ loses its meaning and descriptions in terms of the hyperfine quantum numbers F and MF are more appropriate, transitions with ΔMJ=0 are visible that become weaker as the field is increased.

Figure 6. Stark THz spectrum of the 21+11 rotational transition in ND3 as a function of electric field strength. The detuning with respect to the most intense line in the zero field hyperfine spectrum (F1=3F1=2, see left panel of Figure ) is given. Lines are labelled according to quantum numbers |MJ|. Transitions that follow or violate the ΔMJ=±1 selection rule are indicated in green and red, respectively. The dashed coloured lines that connect spectral features are manually drawn to guide the eye.

Figure 6. Stark THz spectrum of the 21+←11− rotational transition in ND3 as a function of electric field strength. The detuning with respect to the most intense line in the zero field hyperfine spectrum (F1′=3←F1″=2, see left panel of Figure 4) is given. Lines are labelled according to quantum numbers |MJ|. Transitions that follow or violate the ΔMJ=±1 selection rule are indicated in green and red, respectively. The dashed coloured lines that connect spectral features are manually drawn to guide the eye.

From the measured line intensities in comparison to the zero field spectrum, we conclude that significant population transfer into MJ-selected levels of the 21+ can be achieved using our THz source. From simulations and degeneracies of the involved levels, we expect that maximum transfer efficiencies close to unity can in principal be achieved for the |MJ|=1|MJ|=1 and |MJ|=1|MJ|=2 transitions, whereas the maximum transfer efficiency is close to 50% for |MJ|=1|MJ|=0.

3.4. Direct 2111 THz excitation in electric fields

As a second objective, we studied the possibilities to transfer population from the 11 state into specific MJ levels of the 21 state. This is inherently more difficult than populating the 21+ state, as the direct 2111 transition is electric dipole forbidden in zero field. Yet, in the presence of an electric field, the inversion doublet components of both the 11 and the 21 levels mix, such that the direct transition becomes allowed and gains intensity.

We measured the 2111 transition of 14ND3 and 15ND3 in a field of 1.2kV/cm, focussing on the MJ=0|MJ|=1 and |MJ|=2|MJ|=1 transitions. The THz source was again oriented in a way that the electric field vector of the THz radiation was perpendicular to the capacitor field, resulting in a ΔMJ=±1 selection rule. A frequency chirp of 1.44 MHz bandwidth was used when scanning the frequency of the THz source. The resulting Stark shifted spectra are shown in Figure  for 14ND3 and 15ND3. The |MJ|=2|MJ|=1 transitions are observed at smaller detunings than the MJ=0|MJ|=1 transitions, as the Stark shifts of the low-field seeking 21,|MJ|=2 and 11,|MJ|=1 levels partially cancel. A small signature of the |MJ|=1|MJ|=1 transition is observed in the spectrum of 15ND3, just right of the MJ=0|MJ|=1 peak. The appearance of this peak is likely related to an imperfect orientation of the THz source, resulting in weak transitions with ΔMJ=0 character.

Figure 7. Stark shifted spectra of the zero field forbidden 2111 transition of (left) 14ND3 and (right) 15ND3, using 1.44MHz chirps. Peaks are labelled by MJ quantum numbers, and the maximum observed transfer efficiency is given for each transition. The region of the MJ=0MJ=1 transition in 14ND3 was recorded with additional samples to improve the signal-to-noise ratio.

Figure 7. Stark shifted spectra of the zero field forbidden 21−←11− transition of (left) 14ND3 and (right) 15ND3, using 1.44MHz chirps. Peaks are labelled by MJ quantum numbers, and the maximum observed transfer efficiency is given for each transition. The region of the MJ=0←MJ=1 transition in 14ND3 was recorded with additional samples to improve the signal-to-noise ratio.

Population transfer efficiencies were determined by probing the depletion of the 11 population. The maximum transfer efficiencies that could be achieved for the different transitions are indicated by numbers at the respective peaks in Figure . It is seen that in particular the MJ=0|MJ|=1 transitions are driven with poor efficiency, and transfer efficiencies of only 6% and 17% are achieved for 14ND3 and 15ND3, respectively. Much higher efficiencies exceeding 50% are achieved for the |MJ|=2|MJ|=1 transitions.

The dramatic difference in performance for MJ=0 and |MJ|=2 transfer is rationalised by the small transition dipole moments of these types of transitions when MJ=0 states are involved. For the observed transitions to become allowed, the states of different parity 11± and 21± must each mix with their counterparts. The MJ=0 states do not mix in first order, such that the overlap is determined only by the mixing of the 11±,|MJ|=1 states.

From Figure it is clear that higher transfer efficiencies were observed for 15ND3 than for 14ND3. This is related to the smaller inversion splitting of 15ND3 compared to 14ND3 resulting in a more pronounced parity mixing of the levels in an electric field, and to the slightly lower rotational spacing around 614,150MHz for 15ND3, which is closer to the optimal output frequency of the THz source near 615GHz. However, we speculate that mostly the more narrow hyperfine structure of 15ND3 is beneficial. The optimum chirps that were found empirically were of 1.44MHz bandwidth. Comparison with Figure shows that these chirps only cover a part of the hyperfine structure for 14ND3, while for 15ND3 the hyperfine structure is dominated by the quadrupole coupling of the deuterium nuclei and only extends over a few 100kHz.

3.5. 1111+21 microwave-THz double resonance excitation in electric fields

One of our objectives for future experiments is to populate the MJ=0 component of the 21 state, as this state is (almost) immune to electric fields, potentially affording advanced beam merging protocols. The low transfer efficiency observed when inducing the direct 2111 transition in a field is clearly not ideal to achieve this goal. We therefore designed an alternative double-resonance scheme to populate this level. In this scheme, first the microwave transition 1111+ is induced populating the 11+,|MJ|=1 level with 93% transfer efficiency (see Figure ). Subsequently, the zero-field dipole-allowed 21,MJ=011+,|MJ|=1 transition is induced using a THz pulse. The quantum mechanical limit for the transfer efficiency of this second transition, involving transfer from a doubly degenerate |MJ|=1 level into a single MJ=0 level, is 50%. Together, the double resonance scheme should allow for a total transfer efficiency of population from the 11 state into the MJ=0 level of the 21 state of about 46%.

We investigated this double resonance scheme in a field of 1.2kV/cm, using different velocities for the ND3 beam ranging between 360m/s and 450m/s. As before, we used chirps of 1.44MHz bandwidth to achieve optimal transfer. The observed Stark shifted 2111+ spectrum corresponding to the second step of the double-resonance scheme is shown for 14ND3 in the left panel of Figure . The 21,MJ=011+,|MJ|=1 and the 21,|MJ|=211+,|MJ|=1 transitions, that are both allowed by the ΔMJ=±1 selection rule, are clearly visible.

Figure 8. Left: Experimental Stark shifted THz spectrum of the 2111+ rotational transition in 14ND3 in an electric field of 1.2kV/cm, subsequent to a microwave pulse transferring 93% of the original 11 population into the 11+ level (as shown in the right panel of Figure ). Peaks are labelled with the MJ quantum numbers, and the transfer efficiencies of the chirp are given for each peak (estimates flagged with ‘≈’). THz chirps of 30μs duration covering 1.44MHz were used. Right: measured depletion in 11+ population (blue) and corresponding derived gain in 21 population (red) throughout the THz chirp from 618,242.62MHz to 618,241.18MHz.

Figure 8. Left: Experimental Stark shifted THz spectrum of the 21−←11+ rotational transition in 14ND3 in an electric field of 1.2kV/cm, subsequent to a microwave pulse transferring 93% of the original 11− population into the 11+ level (as shown in the right panel of Figure 5). Peaks are labelled with the MJ quantum numbers, and the transfer efficiencies of the chirp are given for each peak (estimates flagged with ‘≈’). THz chirps of 30μs duration covering 1.44MHz were used. Right: measured depletion in 11+ population (blue) and corresponding derived gain in 21− population (red) throughout the THz chirp from 618,242.62MHz to 618,241.18MHz.

From a comparison of Figures and , it is apparent that the 1111+21 double resonance scheme offers a (much) larger transfer efficiency to populate the 21,MJ=0 level than the direct zero-field dipole forbidden 2111 transition. This is quantified further by measuring the depletion of 11+ population while applying a 30μs chirp of decreasing frequency centred at 618,241.900MHz (14ND3) or 615,805.152MHz (15ND3), again using 1.44MHz bandwidth. The recorded evolution of 11+ population during the THz chirp is shown in the right panel of Figure .

Accurate values for the maximum population depletion were acquired by applying the full chirp, and repeating the experiment 1000 times (at 10Hz repetition) toggling between measurements with and without radiation applied. We went for this level of precision only for transitions populating the 21,MJ=0 level; transfer efficiencies for transitions that populate the 21,|MJ|=2 level were estimated from the relative peak intensities in the Stark spectrum of Figure , to be twice as high as the population in MJ=0.

We determined a transfer efficiency for the 21,MJ=011+,|MJ|=1 transition of 36% and 39% for 14ND3 and 15ND3, respectively. These transfer efficiencies were averaged for performance over the 360−450 m/s velocity range probed, varying only by a few % for the different velocities. For the 21,MJ=211+,|MJ|=1 transition, transfer efficiencies of 72% and 78% were estimated for 14ND3 and 15ND3, respectively.

To obtain the total population transfer from 11,|MJ|=1 to 21,MJ=0, the efficiencies of microwave and THz pulse have to be multiplied, yielding in total 93%36%=33.5% for 14ND3 and 36.3% for 15ND3. Clearly, the double resonance scheme is much more effective in populating the 21,MJ=0 level than the direct excitation. But the scheme is also beneficial for population transfer into the 21,|MJ|=2 level, in particular for 14ND3, for which the achieved direct transfer from the 11 to the 21 state was low as seen in Figure . We note that transitions to the 21,|MJ|=1 level were not recorded but could in principle be observed if the THz source was rotated by 90 to change the selection rules.

3.6. THz radiation as a tool to probe MJ-resolved rotational populations

Referring again to Figure , rotational THz transitions in the presence of an electric field provide the option to directly probe the MJ-resolved population in a rotational level. To fully exploit this possibility in future MJMJ resolved collision experiments, methods to prepare molecules in a specific MJ level prior to the collision are required as well. For population transfer into the 21± levels, this is possible by driving THz transitions in an electric field, as discussed in the previous sections. Here, we present a novel method that also allows for the production of samples of ND3 molecules in the 11 state with either pure MJ=0, pure MJ=1, or mixed MJ character. For this, some changes were made to the setup presented in Figure . No microwave radiation was used, the THz source was moved to radiate at the centre of the VMI region, the capacitor plates were removed, and the distance between the Stark decelerator and detection region changed to 530mm. To generate the THz radiation, the LO frequency was changed to 10.5GHz since an improved performance was observed for this setting. Furthermore, the gas mixture was changed to ∼5% ND3 in Neon, producing beams with a mean speed of 890m/s.

Let's first discuss the possibilities to produce packets of ND3 (11) molecules with well-defined MJ composition. In principle, the Stark decelerator only selects those ND3 molecules from the molecular beam pulse that reside in the 11,|MJ|=1 level, as this level is low-field seeking. Throughout the deceleration process, the |MJ|=1 projection is preserved, as the molecules experience a sufficiently large electric field strength at all times and at all positions. However, upon exiting the decelerator, preservation fields may not be present, and reprojections into the MJ=0 can occur. Indeed, in default operation with the Stark decelerator being switched off with subsequent free flight towards the detector, as described in [Citation8], the ND3 molecules are distributed over the MJ=0 (25%) and |MJ|=1 (75%) levels. This non-statistical population ratio can be understood from the hyperfine levels of ND3 that are only partially populated when ND3 redistributes over the MJ=0 and |MJ|=1 levels [Citation8].

It is relatively straightforward to keep the |MJ|=1 projection upon exiting the decelerator, by ensuring that the molecules keep experiencing a sufficiently strong bias field throughout their path towards the detector [Citation28, Citation29]. To allow for experiments with molecules in a pure |MJ|=1 state, a rectangular electrode was placed parallel to the beam axis, starting 57mm downstream from the decelerator and reaching to the entrance of the VMI detector. This electrode will be referred to as ‘preservation electrode’, and was operated at +1kV. The preservation electrode secured a bias field towards the machine ground over the full length of the electrode, as well as in the regions where the molecules exit the decelerator or enter the VMI detector. Fringe fields in the region between the exit of the decelerator and the preservation electrode were enhanced by leaving the last electrodes of the decelerator on for elongated times.

The generation of a pure sample of molecules in MJ=0 is unfortunately less straightforward. We found that the Stark decelerator itself can be used to first produce a sample of ND3 (11) molecules with mixed MJ=0 and |MJ|=1 composition by switching all electrodes off when the packet has reached a position close to the exit of, but still within, the decelerator. An off-time of approximately 10μs was found sufficient to fully reproject the molecules onto the MJ=0 and |MJ|=1 levels. Molecules in the |MJ|=1 level can then subsequently be deflected from the beam axis by operating the remaining stages of the decelerator assymmetrically. This so-called ‘kicker operation’ is achieved by applying 13 kV to only a single electrode from the horizontal pairs, while keeping all other electrodes (the other horizontal electrodes and both vertical ones) at ground. The effectiveness of this kicker operation depends on the forward velocity of the ND3 packet, but we found that typically 10 stages of kicker operation in our experiment was sufficient to fully eliminate molecules in |MJ|=1 from the beam for speeds up to 900 m/s.

To probe the MJ composition of the packet of ND3 (11) molecules, we recorded 21+11 THz spectra in the presence of an electric field, similar to the spectra presented in Figure . For this, we used the THz source in single frequency mode. We used the repeller and extractor plates of the VMI detector to produce a field of 4.6kV/cm, and crossed the THz radiation with the laser beam in the centre of the VMI detector. We then recorded Stark spectra with the REMPI laser tuned to detect population in the 21+ level. The voltages on the VMI electrodes were switched back to their default values a few μs prior to the laser ionisation in order to detect the ions.

Figure  shows the resulting THz spectra when packets of ND3 are produced with different MJ composition as outlined above. In default operation – switching off the Stark decelerator when the molecules exit the decelerator, while the preservation electrode is kept grounded – three peaks are observed. These three peaks correspond to all possible ΔMJ=±1 transitions and are a signature of a mixed MJ=0 and |MJ|=1 population (see also again Figure ). Keeping the Stark decelerator on for an extended period of time to produce a fringe field upon exiting the decelerator, and combining this with the preservation field, only the two peaks remain that originate from 11,|MJ|=1 states. This demonstrates that a pure |MJ|=1 population is kept throughout the experiment, provided that bias fields are present at all times. Switching the Stark decelerator off 10 stages early and then using the last electrodes in kicker operation, followed by the preservation field with the preservation electrode at 1kV, only a single peak in the spectrum remained. This peak originates from molecules in MJ=0, demonstrating the ability to produce samples of ND3 molecules that are exclusively in the 11,MJ=0 state.

Figure 9. THz spectrum of the 21+11 transition of 14ND3 in an electric field of 4.6kV/cm. Packets of ND3 (11) molecules are produced with different MJ compositions. Top: default operation of the decelerator resulting in a mixed composition (25% in MJ=0 and 75% in |MJ|=1). Center: preservation fields downstream from the decelerator enable a pure sample of |MJ|=1 molecules. Bottom: operation of the decelerator with kicker to deflect molecules in |MJ|=1 in combination with preservation fields results in a pure sample of MJ=0 molecules. Spectra are recorded with the THz source in single frequency mode, using a stepsize of 10 kHz while scanning the frequency.

Figure 9. THz spectrum of the 21+←11− transition of 14ND3 in an electric field of 4.6kV/cm. Packets of ND3 (11−) molecules are produced with different MJ compositions. Top: default operation of the decelerator resulting in a mixed composition (25% in MJ=0 and 75% in |MJ|=1). Center: preservation fields downstream from the decelerator enable a pure sample of |MJ|=1 molecules. Bottom: operation of the decelerator with kicker to deflect molecules in |MJ|=1 in combination with preservation fields results in a pure sample of MJ=0 molecules. Spectra are recorded with the THz source in single frequency mode, using a stepsize of 10 kHz while scanning the frequency.

4. Conclusions

We have demonstrated the direct transfer of population between rotational levels of ND3 molecules by driving the 21+11 transition using a commercially available THz radiation source at frequencies near 615 GHz. Using chirps to address all hyperfine components, transfer efficiencies up to 87% were achieved, limited by the maximum available power of the THz source. In the presence of an electric field, fully MJ-resolved transitions were measured for both the zero-field dipole allowed 21+11 and dipole forbidden 2111 transitions. Population transfer from the 11 to the 21 level was also achieved with a superior transfer efficiency using a zero-field dipole allowed 1111+21 microwave-THz double resonance excitation scheme.

We presented a novel method to produce packets of ND3 (11) molecules with well defined MJ composition. Beams with near-perfect MJ=0 or |MJ|=1 purity can be produced using novel operation modes of the Stark decelerator in combination with preservation fields, whereas beams with a MJ=0 (25%) and |MJ|=1 (75%) population distribution can be produced by propagating the beam in zero field downstream from the decelerator. The MJ purity is conveniently probed by inducing MJ-resolved 21+11 THz transitions in the presence of an electric field. The different routes to produce beams with pure MJ composition in either the rotational ground or 21± excited states, in combination with methods to record MJ resolved populations, allows for state-to-state inelastic (rotational de-excitation) collision experiments with full ΔMJ resolution. This offers interesting prospects in scattering experiments with cold and ultracold molecules in the presence of external electric fields.

Acknowledgments

This work is part of the research program of the Netherlands Organisation for Scientific Research (NWO). We thank Niek Janssen and André van Roij for expert technical support.

Disclosure statement

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

Additional information

Funding

S.Y.T.v.d.M. acknowledges support from the European Research Council (ERC) under the European Union's Horizon 2020 Research and Innovation Program (Grant Agreement No. 817947 FICOMOL) and Dutch Research Council (NWO).

Unknown widget #5d0ef076-e0a7-421c-8315-2b007028953f

of type scholix-links

References

  • G. Santambrogio, S.A. Meek, M.J. Abel, L.M. Duffy and G. Meijer, Chem. Phys. Chem. 12 (10), 1799–1807 (2011). doi:10.1002/cphc.v12.10
  • H. Chadwick and R.D. Beck, Annu. Rev. Phys. Chem. 68, 39–61 (2017). doi:10.1146/physchem.2017.68.issue-1
  • T. Noel, M.R. Dietrich, N. Kurz, G. Shu, J. Wright and B.B. Blinov, Phys. Rev. A 85 (2), 023401 (2012). doi:10.1103/PhysRevA.85.023401
  • K. Sato, R. Hirao, S. Yamamoto, K.L. Ivanov and T. Takui, Appl. Magn. Reson. 54 (1), 183–201 (2022). doi:10.1007/s00723-022-01513-8
  • M.H. Oliveira, G. Higgins, C. Zhang, A. Predojević, M. Hennrich, R. Bachelard and C.J. Villas-Boas, Phys. Rev. A 107 (2), 023706 (2023). doi:10.1103/PhysRevA.107.023706
  • G. Pelegr´ı, A.J. Daley and J.D. Pritchard, Quantum Sci. Technol. 7 (4), 045020 (2022). doi:10.1088/2058-9565/ac823a
  • N. Chanda, P. Patnaik and R. Bhattacharyya, Phys. Rev. A 107 (6), 063708 (2023). doi:10.1103/PhysRevA.107.063708
  • S. Herbers, Y.M. Caris, S.E.J. Kuijpers, J.U. Grabow and S.Y.T. van de Meerakker, Mol. Phys. 121, e2129105 (2023). doi:10.1080/00268976.2022.2129105
  • H.L. Bethlem, F.M.H. Crompvoets, R.T. Jongma, S.Y.T. van de Meerakker and G. Meijer, Phys. Rev. A 65, 053416 (2002). doi:10.1103/PhysRevA.65.053416
  • L.P. Parazzoli, N.J. Fitch, P.S. Żuchowski, J.M. Hutson and H.J. Lewandowski, Phys. Rev. Lett.106, 193201 (2011). doi:10.1103/PhysRevLett.106.193201
  • X. Wu, T. Gantner, M. Koller, M. Zeppenfeld, S. Chervenkov and G. Rempe, Science 358, 645–648 (2017). doi:10.1126/science.aan3029
  • M.T. Bell and T.P. Softley, Mol. Phys. 107, 99–132 (2009). doi:10.1080/00268970902724955
  • C. Cheng, A.P.P. van der Poel, P. Jansen, M. Quintero-Pérez, T.E. Wall, W. Ubachs and H.L. Bethlem, Phys. Rev. Lett. 117, 253201 (2016). doi:10.1103/PhysRevLett.117.253201
  • T. de Jongh, Q. Shuai, G.L. Abma, S. Kuijpers, M. Besemer, A. van der Avoird, G.C. Groenenboom and S.Y.T. van de Meerakker, Nat. Chem. 14, 538–544 (2022). doi:10.1038/s41557-022-00896-2
  • G. Tang, M. Besemer, S. Kuijpers, G.C. Groenenboom, A. van der Avoird, T. Karman and S.Y.T. van de Meerakker, Science 379, 1031–1036 (2023). doi:10.1126/science.adf9836
  • R.V. Krems, Int. Rev. Phys. Chem. 24, 99–118 (2005). doi:10.1080/01442350500167161
  • M.C. van Beek, G. Berden, H.L. Bethlem and J.J. ter Meulen, Phys. Rev. Lett. 86, 4001–4004 (2001). doi:10.1103/PhysRevLett.86.4001
  • H.M. Pickett, J. Mol. Spectrosc. 148 (2), 371–377 (1991). doi:10.1016/0022-2852(91)90393-O
  • C.P. Endres, S. Schlemmer, P. Schilke, J. Stutzki and H.S.P. Müller, J. Mol. Spectrosc. 327, 95–104 (2016). doi:10.1016/j.jms.2016.03.005
  • H.S.P. Müller, F. Schlöder, J. Stutzki and G. Winnewisser, J. Mol. Struct. 742 (1-3), 215–227 (2005). doi:10.1016/j.molstruc.2005.01.027
  • H.S.P. Müller, S. Thorwirth, D.A. Roth and G. Winnewisser, A&A 370 (3), L49–L52 (2001).
  • J. van Veldhoven, R.T. Jongma, B. Sartakov, W.A. Bongers and G. Meijer, Phys. Rev. A 66 (3), 32501 (2002). doi:10.1103/PhysRevA.66.032501
  • J. van Veldhoven, J. Küpper, H.L. Bethlem, B. Sartakov, A.J.A. van Roij and G. Meijer, Eur. Phys. J. D 31 (2), 337–349 (2004). doi:10.1140/epjd/e2004-00160-9
  • G. Tang, M. Besemer, T. de Jongh, Q. Shuai, A. van der Avoird, G.C. Groenenboom and S.Y.T. van de Meerakker, J. Chem. Phys. 153 (6), 064301 (2020). doi:10.1063/5.0019472
  • J. Onvlee, S.N. Vogels, A. von Zastrow, D.H. Parker and S.Y.T. van de Meerakker, Phys. Chem. Chem. Phys. 16 (30), 15768–15779 (2014). doi:10.1039/C4CP01519C
  • B. Yan, P.F.H. Claus, B.G.M. van Oorschot, L. Gerritsen, A.T.J.B. Eppink, S.Y.T. van de Meerakker and D.H. Parker, Rev. Sci. Instrum. 84 (2), 023102 (2013). doi:10.1063/1.4790176
  • M.N.R. Ashfold, R.N. Dixon, N. Little, R.J. Stickland and C.M. Western, J. Chem. Phys. 89, 1754–1761 (1988). doi:10.1063/1.455715
  • C.E. Heiner, H.L. Bethlem and G. Meijer, Phys. Chem. Chem. Phys. 8, 2666–2676 (2006). doi:10.1039/b602260j
  • E.W. Steer, L.S. Petralia, C.M. Western, B.R. Heazlewood and T.P. Softley, J. Mol. Spectrosc. 332, 94–102 (2017). doi:10.1016/j.jms.2016.11.003