Controlling few-body reaction pathways using a Feshbach resonance

Shinsuke Haze [email protected] Institut für Quantenmaterie and Center for Integrated Quantum Science and Technology IQST, Universität Ulm, D-89069 Ulm, Germany    Jinglun Li Institut für Quantenmaterie and Center for Integrated Quantum Science and Technology IQST, Universität Ulm, D-89069 Ulm, Germany    Dominik Dorer Institut für Quantenmaterie and Center for Integrated Quantum Science and Technology IQST, Universität Ulm, D-89069 Ulm, Germany    José P. D’Incao Institut für Quantenmaterie and Center for Integrated Quantum Science and Technology IQST, Universität Ulm, D-89069 Ulm, Germany JILA, NIST and Department of Physics, University of Colorado, Boulder, CO 80309-0440, USA    Paul S. Julienne Institut für Quantenmaterie and Center for Integrated Quantum Science and Technology IQST, Universität Ulm, D-89069 Ulm, Germany Joint Quantum Institute, University of Maryland and NIST, College Park, MD 20742, USA    Eberhard Tiemann Institut für Quantenoptik, Leibniz Universität Hannover, 30167 Hannover, Germany    Markus Deiß Institut für Quantenmaterie and Center for Integrated Quantum Science and Technology IQST, Universität Ulm, D-89069 Ulm, Germany    Johannes Hecker Denschlag [email protected] Institut für Quantenmaterie and Center for Integrated Quantum Science and Technology IQST, Universität Ulm, D-89069 Ulm, Germany
(August 27, 2024)
Abstract

Gaining control over chemical reactions on the quantum level is a central goal of the modern field of cold and ultracold chemistry. Here, we demonstrate a novel method to coherently steer reaction flux of a three-body recombination process across different product spin channels. For this, we employ a magnetically-tunable Feshbach resonance to admix, in a controlled way, a specific spin state to the reacting collision complex. This allows for the control of the reaction flux into the admixed spin channel, which can be used to significantly change the reaction products. Furthermore, we also investigate the influence of an Efimov resonance on the reaction dynamics. We find that while the Efimov resonance can be used to globally enhance three-body recombination, the relative flux between the reaction channels remains unchanged. Our control scheme is general and can be extended to other reaction processes. It also provides new opportunities in combination with other control schemes, such as quantum interference of reaction paths.

A chemical reaction in a low-density gas phase is typically well-described by a fully coherent quantum mechanical evolution. Therefore, such a gas is a promising testbed for quantum control of chemical processes. In fact, recent platforms based on ensembles of ultracold atoms or molecules have paved the way for extended quantum mechanical steering of reactions. Demonstrated control schemes include the use of photoassociation [1, 2, 3], Feshbach resonances [4, 5, 6, 7, 8, 9], microwave-engineered collisions [10, 11, 12, 13, 14], electric-field-controlled reactions [15], relative positioning of traps [16, 17, 18, 19], confinement-induced effects [20, 21, 22, 23, 24, 25], quantum interference [26, 27], or rely on propensity rules and conservation laws [28, 29]. This progress has been further promoted by emerging technologies that enable state-to-state measurements (e.g. [30, 31, 32, 33, 34]).

A prominent tool for controlling chemical reactions is a tunable Feshbach resonance. A Feshbach resonance in atomic gases occurs when the energy of the scattering state of two colliding atoms is tuned into degeneracy with that of a molecular state, leading to the mixing of two such states [8]. As they offer unique control over the interparticle interaction, tunable Feshbach resonances have been essential for the development of the field of ultracold quantum gases. An established application of Feshbach resonances for chemical reactions is the controlled production of ultracold molecules. By magnetically ramping over a Feshbach resonance, ultracold pairs of atoms can be converted into an extremely weakly-bound molecule, the Feshbach molecule [4, 5, 35, 36, 37, 8]. In three-body recombination where three free atoms collide to form a diatomic molecule, Feshbach resonances have been used to tune the total molecular production rate and specifically to suppress atom loss [38, 39, 40, 41], and to demonstrate the Efimov effect [42, 40, 43]. Feshbach resonances and resonant scattering have also been proposed for controlling complex few-body reactions, see e.g. [44, 45].

Here, we demonstrate the use of a Feshbach resonance in a three-body recombination process to steer reaction flux between families of molecular product channels with different spin states. More specifically, by tuning the magnetic field towards a Feshbach resonance we can gradually increase the initially negligible reaction rate into a specific spin channel, so that it becomes close to the total rate into all channels. The process is coherent and represents a novel tool for state-selective controlling of molecular production rates, using the applied magnetic field as a precisely tunable control knob.

The experiments are carried out with a 860nK860nK860\>\textrm{nK}860 nK-cold cloud of about 2.5×1052.5superscript1052.5\times 10^{5}2.5 × 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT 85Rb atoms where each atom i𝑖iitalic_i is in the hyperfine state (fi,mfi)=(2,2)subscript𝑓𝑖subscript𝑚subscript𝑓𝑖22(f_{i},m_{f_{i}})=(2,-2)( italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT ) = ( 2 , - 2 ) of the electronic ground state. The atoms are confined in a far-detuned crossed optical dipole trap, for more details see Methods and [46]. In the atom cloud, three-body recombination spontaneously occurs, predominantly producing weakly-bound molecules in states of the coupled molecular complex X1Σg+a3Σu+superscript𝑋1superscriptsubscriptΣ𝑔superscript𝑎3superscriptsubscriptΣ𝑢X^{1}\Sigma_{g}^{+}-a^{3}\Sigma_{u}^{+}italic_X start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT roman_Σ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT - italic_a start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT roman_Σ start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT. By tuning the magnetic field B𝐵Bitalic_B in the vicinity of the s𝑠sitalic_s-wave Feshbach resonance at B=155G𝐵155GB=155\>\textrm{G}italic_B = 155 G [47, 48] we can control the product distribution of the molecules. For the details of the Feshbach resonance, see also Supplemental Materials. The molecules are state-selectively detected via resonance-enhanced multiphoton ionization (REMPI), see Methods for details.

Refer to caption
Figure 1: Scheme for controlling the reaction flux into different spin channels using a two-body Feshbach resonance. (a) Atoms (a,b,c𝑎𝑏𝑐a,b,citalic_a , italic_b , italic_c) undergo three-body recombination, where (a,b𝑎𝑏a,bitalic_a , italic_b) form a molecule. During this process, atom (c)𝑐(c)( italic_c ) is outside the ranges (cyan areas) for spin-exchange interaction with the other atoms. (b) Schematic representation of the Born-Oppenheimer potential energy curves for the atom pair (a,b𝑎𝑏a,bitalic_a , italic_b). At close distance, the incoming scattering state (1) with spin |ket|\uparrow\rangle| ↑ ⟩ experiences admixing of the bound state (2) which has spin |ket|\downarrow\rangle| ↓ ⟩. Upon collision with the third atom (c𝑐citalic_c) (not shown here) the scattering state can then relax into molecular bound states (3) or (4), with their respective spin states |ket|\downarrow\rangle| ↓ ⟩ and |ket|\uparrow\rangle| ↑ ⟩.

Our scheme for controlling the reaction flux into different spin channels is illustrated in Fig. 1. In the three-body recombination process, the Rb atoms (a,b,c𝑎𝑏𝑐a,b,citalic_a , italic_b , italic_c) collide and (a,b𝑎𝑏a,bitalic_a , italic_b) form a molecule, see Fig. 1(a). In the particular reactions we study here, the third atom (c𝑐citalic_c) is far enough away, so that it interacts with the atoms (a,b𝑎𝑏a,bitalic_a , italic_b) merely mechanically and no spin flip between (c𝑐citalic_c) and the pair (a,b𝑎𝑏a,bitalic_a , italic_b) occurs [29, 49]. Therefore, spin physics aspects of the reaction can be understood to a large extent in a two-body picture, where atom (a)𝑎(a)( italic_a ) collides with atom (b)𝑏(b)( italic_b ). At large internuclear distances the (a,b𝑎𝑏a,bitalic_a , italic_b) scattering state has the hyperfine spin quantum numbers (F,fa,fb,mF)=(4,2,2,4)𝐹subscript𝑓𝑎subscript𝑓𝑏subscript𝑚𝐹4224(F,f_{a},f_{b},m_{F})=(4,2,2,-4)( italic_F , italic_f start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT , italic_f start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) = ( 4 , 2 , 2 , - 4 ), where F𝐹Fitalic_F denotes the total angular momentum of the molecule excluding rotation, and mF=mfa+mfbsubscript𝑚𝐹subscript𝑚subscript𝑓𝑎subscript𝑚subscript𝑓𝑏m_{F}=m_{f_{a}}+m_{f_{b}}italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = italic_m start_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT end_POSTSUBSCRIPT + italic_m start_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT end_POSTSUBSCRIPT represents its projection. We denote this spin state by |ket|\uparrow\rangle| ↑ ⟩. At short internuclear distances the scattering state couples to an energetically near-by molecular bound level, giving rise to the Feshbach resonance. This level has the spin state (F,fa,fb,mF)=(4,3,3,4)𝐹subscript𝑓𝑎subscript𝑓𝑏subscript𝑚𝐹4334(F,f_{a},f_{b},m_{F})=(4,3,3,-4)( italic_F , italic_f start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT , italic_f start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) = ( 4 , 3 , 3 , - 4 ) which we denote by |ket|\downarrow\rangle| ↓ ⟩ [50]. The coupling leads to an admixture of the |ket|\downarrow\rangle| ↓ ⟩ state to the initial scattering state with spin |ket|\uparrow\rangle| ↑ ⟩, and the strength of this admixture can be magnetically controlled. Next, in the mechanical collision with atom (c)𝑐(c)( italic_c ), the scattering state of (a,b𝑎𝑏a,bitalic_a , italic_b) can transition into a molecular bound state. Due to angular momentum conservation, the spin state of the newly formed molecule must, however, have overlap either with the spin state |ket|\uparrow\rangle| ↑ ⟩ or with |ket|\downarrow\rangle| ↓ ⟩ [51]. In fact, by tuning the |ket|\downarrow\rangle| ↓ ⟩ admixture of the (a,b𝑎𝑏a,bitalic_a , italic_b) scattering state we can control the reaction flux into molecular product channels with spin |ket|\downarrow\rangle| ↓ ⟩.

Refer to caption
Figure 2: Observation of |ket|\uparrow\rangle| ↑ ⟩ and |ket|\downarrow\rangle| ↓ ⟩ molecules. Shown are REMPI spectra as a function of the REMPI laser frequency ν𝜈\nuitalic_ν for various magnetic fields B𝐵Bitalic_B. Here, ν0=497603.591GHzsubscript𝜈0497603.591GHz\nu_{0}=497603.591\>\textrm{GHz}italic_ν start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 497603.591 GHz. Each dip in a trace corresponds to a signal from a distinct molecular level. The REMPI signals are normalized, ranging from 0 to 1, as indicated by the vertical bar. The bar is valid for all data traces. The diamonds mark the theoretical positions of possible molecular signals and the colors indicate the spin state as well as the vibrational and rotational level (v,LRvsubscript𝐿𝑅\textrm{v},L_{R}v , italic_L start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT). The faint color bands connecting the diamonds are guides to the eye. We note that the binding energy of the |ket|\uparrow\rangle| ↑ ⟩ level is smaller than that of the two |ket|\downarrow\rangle| ↓ ⟩ levels. In the shown spectra, however, the signal for |ket|\uparrow\rangle| ↑ ⟩ is at higher frequency ν𝜈\nuitalic_ν since the intermediate rotational level for the REMPI is different, see also Methods.

We now demonstrate this control scheme experimentally. Figure 2 presents REMPI spectra in a selected frequency range, showing signals from three different molecular product states. The spectra have been taken within a range of magnetic fields B𝐵Bitalic_B between 4.6G4.6G4.6\>\textrm{G}4.6 G and 159.3G159.3G159.3\>\textrm{G}159.3 G. (Note the nonuniform step size of B𝐵Bitalic_B.) ν𝜈\nuitalic_ν represents the REMPI laser frequency and ν0subscript𝜈0\nu_{0}italic_ν start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is a frequency reference, see Methods. Each dip in a trace for a chosen B𝐵Bitalic_B-field corresponds to a state-specific product molecule signal. Colored diamonds mark the known resonance positions predicted from coupled-channel calculations.

Refer to caption
Figure 3: Opening up a product spin channel. (a) Molecule detection rates for three product states for which the quantum numbers |/(v,LR)|\uparrow/\downarrow\rangle(\textrm{v},L_{R})| ↑ / ↓ ⟩ ( v , italic_L start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ) are given in the legend. The gray dashed line marks the experimental detection limit and the light-green and purple shaded areas indicate the position of the Efimov and Feshbach resonance, respectively. The error bars in the plot indicate one standard deviation (1σ𝜎\sigmaitalic_σ). (b) Calculated three-body recombination rate coefficients. The black dashed line is the total rate coefficient L3,totsubscript𝐿3totL_{3,\mathrm{tot}}italic_L start_POSTSUBSCRIPT 3 , roman_tot end_POSTSUBSCRIPT. Solid colored lines correspond to partial rates for the states under discussion. Gray lines correspond to other molecular states. No calculations are shown for 152GB157G152G𝐵157G152\>\textrm{G}\leq B\leq 157\>\textrm{G}152 G ≤ italic_B ≤ 157 G, see text. We expect theoretical errors up to a few tens of percent for the partial rate for vibrational levels down to v =4absent4=-4= - 4, judging from when more vibrational states are included in our effective potentials [46]. (c) The normalized reaction rate coefficients L3/L3,totsubscript𝐿3subscript𝐿3totL_{3}/L_{3,\mathrm{tot}}italic_L start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT / italic_L start_POSTSUBSCRIPT 3 , roman_tot end_POSTSUBSCRIPT do not exhibit a maximum at the Efimov resonance.

The molecular levels are labeled by their spin states (|ket|\uparrow\rangle| ↑ ⟩ or |ket|\downarrow\rangle| ↓ ⟩), and by their vibrational (v) and rotational (LRsubscript𝐿𝑅L_{R}italic_L start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT) quantum numbers [48]. In Fig. 2 we observe one molecular state with spin |ket|\uparrow\rangle| ↑ ⟩ and two molecular states with spin |ket|\downarrow\rangle| ↓ ⟩. The resonance positions change in a characteristic way with the B𝐵Bitalic_B-field due to the Zeeman effect. We use this as fingerprint information for identifying the molecular levels. The strength of each signal roughly reflects the recombination rate towards each respective state. In the magnetic field range up to about 120G120G120\>\textrm{G}120 G each REMPI spectrum exhibits only a single resonance dip which can be unambiguously assigned to the state |ket{|\uparrow\rangle}| ↑ ⟩. At about 120G120G120\>\textrm{G}120 G, two additional molecular signals start to appear stemming from molecular states in the spin state |ket|\downarrow\rangle| ↓ ⟩. The strengths of the signals of these states become similar to the |ket|\uparrow\rangle| ↑ ⟩ signals when approaching the Feshbach resonance at 155G155G155\>\textrm{G}155 G. For magnetic fields above the Feshbach resonance, all signals decrease very quickly within a few Gauss.

From our REMPI spectra, molecule detection rates for each observed molecular state are extracted. These rates are roughly proportional to the partial three-body recombination rates for the flux into individual product channels (see [46] and Methods). The obtained rates for the states in Fig. 2 are shown in Fig. 3(a) for the magnetic field region in the vicinity of the Efimov and Feshbach resonances, located at 140G140G140\>\textrm{G}140 G and 155G155G155\>\textrm{G}155 G, respectively.

The data show that the rates for the |ket|\downarrow\rangle| ↓ ⟩ states indeed strongly increase from below the detection limit (gray dashed line) to about a factor of 50 above the detection limit as the magnetic field is increased from B<115G𝐵115GB<115\>\textrm{G}italic_B < 115 G towards the Efimov resonance. The detection limit is mainly determined by the background noise of our REMPI scheme. By contrast, the rate for the |ket|\uparrow\rangle| ↑ ⟩ state is rather constant for all B𝐵Bitalic_B-fields below the Efimov resonance. At the position of the Efimov resonance at 140 G we observe a clear enhancement of the rates for all molecular states. In fact, the signals for all three states attain similar strength, which demonstrates the large relative tuning range of our scheme. At this point the spin product channel |ket|\downarrow\rangle| ↓ ⟩ has been fully opened up for the reaction flux.

The relatively constant production rate for the |ket|\uparrow\rangle| ↑ ⟩ state below the Efimov resonance might be unexpected at first in view of the known a4superscript𝑎4a^{4}italic_a start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT scaling of the recombination rate in the limit of zero temperature [52, 6], where a𝑎aitalic_a is the scattering length. It can, however, be explained to a large extent as an effect of our finite temperature of 860 nK [53, 54], as further discussed below.

We carried out numerical model calculations for the partial rate coefficients L3subscript𝐿3L_{3}italic_L start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT for each molecular quantum state, using the adiabatic hyperspherical representation [55, 56, 39] (see Methods). This determines the partial recombination rate, L3(f)n3d3r/3subscript𝐿3𝑓superscript𝑛3superscript𝑑3𝑟3L_{3}(f)\cdot\int n^{3}d^{3}r\,/3italic_L start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ( italic_f ) ⋅ ∫ italic_n start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_d start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_r / 3, into a molecular state f𝑓fitalic_f, where n𝑛nitalic_n is the atomic density distribution. The calculations take into account thermal averaging of the partial rate constants. The results are shown in Fig. 3(b). The region from 152152152152 to 157G157G157\>\textrm{G}157 G, i.e., the direct vicinity of the Feshbach resonance is excluded since the numerical calculations quickly become computationally highly demanding in this resonant regime [57].

Among all the possible molecular states produced by recombination (gray dotted lines), we highlight in color the molecular states under discussion. In addition, we present the total three-body recombination rate coefficient (dashed black lines).

Our calculations show that due to thermal averaging, the calculated recombination rate coefficient for the probed |ket|\uparrow\rangle| ↑ ⟩ state increases only moderately towards the Efimov resonance. For more details on how finite temperature affects the recombination rate coefficient, see Supplemental Materials. The increase of the theoretical curves is still faster than for the experimental data. This may be mainly explained by imperfections of the experiment. During the B𝐵Bitalic_B-field ramp atoms are already lost due to three-body recombination and the sample slightly heats up. As a consequence, the density of the atom cloud sinks. In addition, the B-field ramp is not perfect, but tends to lag behind and to overshoot, which can lead to averaging out of signals. Furthermore, there could be a small variation in the REMPI efficiency as a function of magnetic field. These variations hamper a direct comparison between ion rates and L3subscript𝐿3L_{3}italic_L start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT coefficients.

Nevertheless, the main characteristics of the experimental data are qualitatively well described. For example, the observed sharp drop of the recombination rate above the Feshbach resonance is also clearly reproduced by the theory. The reason for this drop is the rapid decrease of the scattering length towards its zero crossing near B=166G𝐵166GB=166\>\textrm{G}italic_B = 166 G and the close-by minimum in L3subscript𝐿3L_{3}italic_L start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT due to Efimov physics [39, 58]. We note that above the Feshbach resonance the Feshbach molecular state appears (see dark-green solid line in Fig. 3(b)), and takes the main fraction of the total reaction flux.

Our calculations show that the effect of the Efimov resonance is to increase the partial three-body recombination rate coefficients with the same overall factor, not favoring particular product channels. This is evident from Fig. 3(b) where all partial rate coefficients exhibit a similar maximum at the location of the Efimov resonance. It also becomes manifest when normalizing the partial rate coefficients to the total rate coefficient (see Fig. 3(c)), as each maximum at the Efimov resonance disappears. The global enhancement is due to the fact that the Efimov resonance is a shape resonance which occurs in a single three-body adiabatic channel. As such, approaching the resonance increases the overall amplitude of the three-body scattering wavefunction at short distances where the reaction takes place, therefore, enhancing all the partial rates by the same factor [43].

Refer to caption
Figure 4: Spin families of molecular products. (a) and (b) Detection signals of product molecules of various spin families by REMPI spectroscopy at a magnetic field of B=4.6G𝐵4.6GB=4.6\>\textrm{G}italic_B = 4.6 G and B=155G𝐵155GB=155\>\textrm{G}italic_B = 155 G, respectively. Vertical lines correspond to calculated resonance positions for molecular states assigned to spin families (F,fa,fb𝐹subscript𝑓𝑎subscript𝑓𝑏F,f_{a},f_{b}italic_F , italic_f start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT , italic_f start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT) according to the legend on top of the figure. The marked individual states have vibrational quantum numbers in the range from v=1v1\textrm{v}=-1v = - 1 to 77-7- 7 and rotational quantum numbers LR=0,2,4subscript𝐿𝑅024L_{R}=0,2,4italic_L start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT = 0 , 2 , 4 or 6666. The REMPI path is via a Πg3superscriptsubscriptΠ𝑔3{}^{3}\Pi_{g}start_FLOATSUPERSCRIPT 3 end_FLOATSUPERSCRIPT roman_Π start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT intermediate state, see Methods. We have νB=ν0=497831.928subscript𝜈𝐵subscript𝜈0497831.928\nu_{B}=\nu_{0}=497831.928italic_ν start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT = italic_ν start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 497831.928 GHz for (a) and νB=ν0228subscript𝜈𝐵subscript𝜈0228\nu_{B}=\nu_{0}-228italic_ν start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT = italic_ν start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - 228 MHz for (b). The 228 MHz shift compensates for the Zeeman shift, allowing for a better comparison of the two spectra. (c) Calculated, summed-up molecular product fraction for each spin family (F,fa,fb)𝐹subscript𝑓𝑎subscript𝑓𝑏(F,f_{a},f_{b})( italic_F , italic_f start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT , italic_f start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ) as indicated next to the curves.

Remarkably, near the Feshbach resonance we also find, both experimentally and theoretically, molecular products in spin states other than |ket|\uparrow\rangle| ↑ ⟩ and |ket|\downarrow\rangle| ↓ ⟩, as shown in Fig. 4. This points towards physics beyond the |ket|\uparrow\rangle| ↑ ⟩-|ket|\downarrow\rangle| ↓ ⟩ Feshbach mixing. In the experiments, these are molecular products with spins (F,fa,fb)=(4,3,2)𝐹subscript𝑓𝑎subscript𝑓𝑏432(F,f_{a},f_{b})=(4,3,2)( italic_F , italic_f start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT , italic_f start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ) = ( 4 , 3 , 2 ) and (5,3,2)532(5,3,2)( 5 , 3 , 2 ). In this notation, we omit mFsubscript𝑚𝐹m_{F}italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT, as it is always mF=4subscript𝑚𝐹4m_{F}=-4italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = - 4. Thus, we observe product states where only one of the fisubscript𝑓𝑖f_{i}italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT has flipped, and where even the total angular momentum F𝐹Fitalic_F can change. Following a similar analysis as used in Refs. [29, 46], the observation of such states can be understood as follows in terms of two-body physics. The spin state (4,3,2)432(4,3,2)( 4 , 3 , 2 ) can be produced via two-body spin-exchange interaction at short distances, starting either from state |(4,2,2)ket422|\uparrow\rangle\equiv(4,2,2)| ↑ ⟩ ≡ ( 4 , 2 , 2 ) or from |(4,3,3)ket433|\downarrow\rangle\equiv(4,3,3)| ↓ ⟩ ≡ ( 4 , 3 , 3 ). Close to the Feshbach resonance the scattering wavefunction amplitude is strongly enhanced at short range and with it also the rate for spin-exchange. Producing the spin state (5,3,2)532(5,3,2)( 5 , 3 , 2 ) is possible due to the presence of a finite B𝐵Bitalic_B-field which breaks global rotational symmetry and couples different F𝐹Fitalic_F quantum numbers. A (5,3,2)532(5,3,2)( 5 , 3 , 2 ) state is typically energetically close to a corresponding (4,3,2)432(4,3,2)( 4 , 3 , 2 ) state, such that coupling between them is resonantly enhanced.

In Fig.  4(a) and (b) REMPI spectra at magnetic fields of B=4.6G𝐵4.6GB=4.6\>\textrm{G}italic_B = 4.6 G and B=155G𝐵155GB=155\>\textrm{G}italic_B = 155 G, respectively, are compared. The spectrum at high magnetic field exhibits many more resonance lines than the spectrum at low field. In a thorough analysis of the spectra, similarly as in Ref. [30], we identified a total of four spin families for B=155G𝐵155GB=155\>\textrm{G}italic_B = 155 G and only a single one for B=4.6G𝐵4.6GB=4.6\>\textrm{G}italic_B = 4.6 G. The individual spin states are marked with colored bars in Fig. 4(a) and (b).

Figure 4(c) shows our numerical calculations for the product fractions of the molecules in the different spin families (F,fa,fb)𝐹subscript𝑓𝑎subscript𝑓𝑏(F,f_{a},f_{b})( italic_F , italic_f start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT , italic_f start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ) as a function of the external magnetic field B𝐵Bitalic_B. Here, for each individual family we sum up the populations for all corresponding molecular states having the same spin characteristics. In agreement with our previous discussion, spin-exchange is strongly enhanced when approaching the Feshbach resonance. Furthermore, Fig. 4(c) reveals a hierarchy in the propensity for the production of spin states. Changing the total angular momentum F𝐹Fitalic_F is more strongly suppressed than changing an atomic f𝑓fitalic_f quantum number, see also [59].

In summary, we have demonstrated a powerful scheme to control the reaction pathway in a three-body recombination process of ultracold atoms. Using a magnetically tunable Feshbach resonance we admixed a well-defined spin state to the reaction complex of three atoms and by this steered the reaction flux between the corresponding spin channels. We find that a large fraction of the total reaction flux can be redirected in this way. Furthermore, we show that in contrast to the Feshbach resonance an Efimov resonance only enhances globally the reaction rate, while maintaining the relative flux between reaction channels. We investigated our control scheme both experimentally and theoretically, using high-resolution state-to-state measurements and state-of-the-art numerical three-body scattering calculations, respectively.

The demonstrated reaction control holds large promise for general few-body reactions, as it is simple and can easily be extended. Feshbach resonances are ubiquitous in cold atomic and molecular gases. The scheme is fully coherent and can thus be used as a central building block in interferometric control, where the Feshbach resonance functions as a beam splitter for the incoming wave function. The split up parts can then potentially follow different pathways towards the same final product state where they interfere. For example, the final product state could have tunable spin-mixed character which can be set by further control methods such as state dressing with optical or microwave fields. In this way, additional tuning of the interference can be achieved.

I acknowledgements

This work was financed by the Baden-Württemberg Stiftung through the Internationale Spitzenforschung program (contract BWST ISF2017-061) and by the German Research Foundation (DFG, Deutsche Forschungsgemeinschaft) within contract 399903135. We acknowledge support from bwFor-Cluster JUSTUS 2 for high performance computing. J. P. D. also acknowledges partial support from the U.S. National Science Foundation, Grant No. PHY-2012125 and PHY-2308791, and NASA/JPL 1502690. S.H. also acknowledges support from Japan Science and Technology Agency Moonshot R&\&&D Grant No. JPMJMS2063 and ASPIRE Grant No. JPMJAP2319.

II Author contributions

S.H. and D.D. have carried out the experiments. J.L. and J.P.D. calculated the three-body recombination rate coefficients. J.H.D. supervised the project. All authors have contributed to the analysis of the experiment and to the writing of the manuscript.

III Competing interests

The authors declare no competing interests.

References

  • Jones et al. [2006] K. M. Jones, E. Tiesinga, P. D. Lett, and P. S. Julienne, Ultracold photoassociation spectroscopy: Long-range molecules and atomic scattering, Rev. Mod. Phys. 78, 483 (2006).
  • Lett et al. [1993] P. D. Lett, K. Helmerson, W. D. Phillips, L. P. Ratliff, S. L. Rolston, and M. E. Wagshul, Spectroscopy of Na2 by photoassociation of laser-cooled Na, Phys. Rev. Lett. 71, 2200 (1993).
  • Miller et al. [1993] J. D. Miller, R. A. Cline, and D. J. Heinzen, Photoassociation spectrum of ultracold Rb atoms, Phys. Rev. Lett. 71, 2204 (1993).
  • Donley et al. [2002] E. A. Donley, N. R. Claussen, S. T. Thompson, and C. E. Wieman, Atom-molecule coherence in a Bose-Einstein condensate, Nature 417, 529 (2002).
  • Herbig et al. [2003] J. Herbig, T. Kraemer, M. Mark, T. Weber, C. Chin, H.-C. Nägerl, and R. Grimm, Preparation of a pure molecular quantum gas, Science 301, 1510 (2003).
  • Weber et al. [2003a] T. Weber, J. Herbig, M. Mark, H.-C. Nägerl, and R. Grimm, Three-body recombination at large scattering lengths in an ultracold atomic gas, Phys. Rev. Lett. 91, 123201 (2003a).
  • Yang et al. [2022] H. Yang, J. Cao, Z. Su, J. Rui, B. Zhao, and J.-W. Pan, Creation of an ultracold gas of triatomic molecules from an atom-diatomic molecule mixture, Science 378, 1009 (2022).
  • Chin et al. [2010] C. Chin, R. Grimm, P. Julienne, and E. Tiesinga, Feshbach resonances in ultracold gases, Rev. Mod. Phys. 82, 1225 (2010).
  • Park et al. [2023] J. J. Park, Y.-K. Lu, O. A. Jamison, T. V. Tscherbul, and W. Ketterle, A Feshbach resonance in collisions between triplet ground-state molecules, Nature 614, 54 (2023).
  • Anderegg et al. [2021] L. Anderegg, S. Burchesky, Y. Bao, S. S. Yu, T. Karman, E. Chae, K.-K. Ni, W. Ketterle, and J. M. Doyle, Observation of microwave shielding of ultracold molecules, Science 373, 779 (2021).
  • Lin et al. [2023] J. Lin, G. Chen, M. Jin, Z. Shi, F. Deng, W. Zhang, G. Quéméner, T. Shi, S. Yi, and D. Wang, Microwave shielding of bosonic NaRb molecules, Phys. Rev. X 13, 031032 (2023).
  • Bigagli et al. [2023] N. Bigagli, C. Warner, W. Yuan, S. Zhang, I. Stevenson, T. Karman, and S. Will, Collisionally stable gas of bosonic dipolar ground-state molecules, Nat. Phys. 19, 1579 (2023).
  • Chen et al. [2024] X.-Y. Chen, S. Biswas, S. Eppelt, A. Schindewolf, F. Deng, T. Shi, S. Yi, T. Hilker, I. Bloch, and X.-Y. Luo, Ultracold field-linked tetratomic molecules, Nature 626, 283 (2024).
  • Yan et al. [2020] Z. Z. Yan, J. W. Park, Y. Ni, H. Loh, S. Will, T. Karman, and M. Zwierlein, Resonant dipolar collisions of ultracold molecules induced by microwave dressing, Phys. Rev. Lett. 125, 063401 (2020).
  • Matsuda et al. [2020] K. Matsuda, L. De Marco, J.-R. Li, W. G. Tobias, G. Valtolina, G. Quéméner, and J. Ye, Resonant collisional shielding of reactive molecules using electric fields, Science 370, 1324 (2020).
  • Ruttley et al. [2023] D. K. Ruttley, A. Guttridge, S. Spence, R. C. Bird, C. R. Le Sueur, J. M. Hutson, and S. L. Cornish, Formation of ultracold molecules by merging optical tweezers, Phys. Rev. Lett. 130, 223401 (2023).
  • Yu et al. [2021] Y. Yu, K. Wang, J. D. Hood, L. R. B. Picard, J. T. Zhang, W. B. Cairncross, J. M. Hutson, R. Gonzalez-Ferez, T. Rosenband, and K.-K. Ni, Coherent optical creation of a single molecule, Phys. Rev. X 11, 031061 (2021).
  • Reynolds et al. [2020] L. A. Reynolds, E. Schwartz, U. Ebling, M. Weyland, J. Brand, and M. F. Andersen, Direct measurements of collisional dynamics in cold atom triads, Phys. Rev. Lett. 124, 073401 (2020).
  • Cheuk et al. [2020] L. W. Cheuk, L. Anderegg, Y. Bao, S. Burchesky, S. S. Yu, W. Ketterle, K.-K. Ni, and J. M. Doyle, Observation of collisions between two ultracold ground-state CaF molecules, Phys. Rev. Lett. 125, 043401 (2020).
  • Tobias et al. [2022] W. G. Tobias, K. Matsuda, J.-R. Li, C. Miller, A. N. Carroll, T. Bilitewski, A. M. Rey, and J. Ye, Reactions between layer-resolved molecules mediated by dipolar spin exchange, Science 375, 1299 (2022).
  • de Miranda et al. [2011] M. H. G. de Miranda, A. Chotia, B. Neyenhuis, D. Wang, G. Quéméner, S. Ospelkaus, J. L. Bohn, J. Ye, and D. S. Jin, Controlling the quantum stereodynamics of ultracold bimolecular reactions, Nat. Phys. 7, 502 (2011).
  • Drews et al. [2017] B. Drews, M. Deiß, K. Jachymski, Z. Idziaszek, and J. Hecker Denschlag, Inelastic collisions of ultracold triplet Rb2 molecules in the rovibrational ground state, Nat. Commun. 8, 14854 (2017).
  • Goban et al. [2018] A. Goban, R. B. Hutson, G. E. Marti, S. L. Campbell, M. A. Perlin, P. S. Julienne, J. P. D’Incao, A. M. Rey, and J. Ye, Emergence of multi-body interactions in a fermionic lattice clock, Nature 563, 369 (2018).
  • Sala et al. [2013] S. Sala, G. Zürn, T. Lompe, A. N. Wenz, S. Murmann, F. Serwane, S. Jochim, and A. Saenz, Coherent molecule formation in anharmonic potentials near confinement-induced resonances, Phys. Rev. Lett. 110, 203202 (2013).
  • Lee et al. [2023] Y. K. Lee, H. Lin, and W. Ketterle, Spin dynamics dominated by resonant tunneling into molecular states, Phys. Rev. Lett. 131, 213001 (2023).
  • Son et al. [2022] H. Son, J. J. Park, Y.-K. Lu, A. O. Jamison, T. Karman, and W. Ketterle, Control of reactive collisions by quantum interference, Science 375, 1006 (2022).
  • [27] Y.-X. Liu, L. Zhu, J. Luke, J. J. A. Houwman, M. C. Babin, M.-G. Hu, and K.-K. Ni, Quantum interference and entanglement in ultracold atom-exchange reactions, arXiv:2310.07620 (2023) .
  • Hu et al. [2021] M.-G. Hu, Y. Liu, M. A. Nichols, L. Zhu, G. Quéméner, O. Dulieu, and K.-K. Ni, Nuclear spin conservation enables state-to-state control of ultracold molecular reactions, Nat. Chem. 13, 435 (2021).
  • Haze et al. [2022] S. Haze, J. P. D’Incao, D. Dorer, M. Deiß, E. Tiemann, P. S. Julienne, and J. Hecker Denschlag, Spin-conservation propensity rule for three-body recombination of ultracold Rb atoms, Phys. Rev. Lett. 128, 133401 (2022).
  • Wolf et al. [2017] J. Wolf, M. Deiß, A. Krükow, E. Tiemann, B. P. Ruzic, Y. Wang, J. P. D’Incao, P. S. Julienne, and J. Hecker Denschlag, State-to-state chemistry for three-body recombination in an ultracold rubidium gas, Science 358, 921 (2017).
  • Wolf et al. [2019] J. Wolf, M. Deiß, and J. Hecker Denschlag, Hyperfine magnetic substate resolved state-to-state chemistry, Phys. Rev. Lett. 123, 253401 (2019).
  • Liu et al. [2021] Y. Liu, M.-G. Hu, M. A. Nichols, D. Yang, D. Xie, H. Guo, and K.-K. Ni, Precision test of statistical dynamics with state-to-state ultracold chemistry, Nature 593, 379 (2021).
  • Rui et al. [2017] J. Rui, H. Yang, L. Liu, D.-C. Zhang, Y.-X. Liu, J. Nan, Y.-A. Chen, B. Zhao, and J.-W. Pan, Controlled state-to-state atom-exchange reaction in an ultracold atom–dimer mixture, Nat. Phys. 13, 699 (2017).
  • Hoffmann et al. [2018] D. K. Hoffmann, T. Paintner, W. Limmer, D. S. Petrov, and J. Hecker Denschlag, Reaction kinetics of ultracold molecule-molecule collisions, Nat. Commun. 9, 5244 (2018).
  • Dürr et al. [2004] S. Dürr, T. Volz, A. Marte, and G. Rempe, Observation of molecules produced from a Bose-Einstein condensate, Phys. Rev. Lett. 92, 020406 (2004).
  • Xu et al. [2003] K. Xu, T. Mukaiyama, J. R. Abo-Shaeer, J. K. Chin, D. E. Miller, and W. Ketterle, Formation of quantum-degenerate sodium molecules, Phys. Rev. Lett. 91, 210402 (2003).
  • Thalhammer et al. [2006] G. Thalhammer, K. Winkler, F. Lang, S. Schmid, R. Grimm, and J. Hecker Denschlag, Long-lived Feshbach molecules in a three-dimensional optical lattice, Phys. Rev. Lett. 96, 050402 (2006).
  • Jochim et al. [2003] S. Jochim, M. Bartenstein, A. Altmeyer, G. Hendl, C. Chin, J. Hecker Denschlag, and R. Grimm, Pure gas of optically trapped molecules created from fermionic atoms, Phys. Rev. Lett. 91, 240402 (2003).
  • Xie et al. [2020] X. Xie, M. J. Van de Graaff, R. Chapurin, M. D. Frye, J. M. Hutson, J. P. D’Incao, P. S. Julienne, J. Ye, and E. A. Cornell, Observation of Efimov universality across a nonuniversal Feshbach resonance in K39superscriptK39{}^{39}\mathrm{K}start_FLOATSUPERSCRIPT 39 end_FLOATSUPERSCRIPT roman_KPhys. Rev. Lett. 125, 243401 (2020).
  • Braaten and Hammer [2006] E. Braaten and H.-W. Hammer, Universality in few-body systems with large scattering length, Physics Reports 428, 259 (2006).
  • Weber et al. [2003b] T. Weber, J. Herbig, M. Mark, H.-C. Nägerl, and R. Grimm, Bose-Einstein condensation of cesium, Science 299, 232 (2003b).
  • Ferlaino et al. [2011] F. Ferlaino, A. Zenesini, M. Berninger, B. Huang, H.-C. Nägerl, and R. Grimm, Efimov resonances in ultracold quantum gases, Few-Body Syst. 51, 113 (2011).
  • D’Incao [2018] J. P. D’Incao, Few-body physics in resonantly interacting ultracold quantum gases, J. Phys. B: At. Mol. Opt. Phys. 51, 043001 (2018).
  • Hermsmeier et al. [2021] R. Hermsmeier, J. Kłos, S. Kotochigova, and T. V. Tscherbul, Quantum spin state selectivity and magnetic tuning of ultracold chemical reactions of triplet alkali-metal dimers with alkali-metal atoms, Phys. Rev. Lett. 127, 103402 (2021).
  • Tscherbul and Krems [2015] T. V. Tscherbul and R. V. Krems, Tuning bimolecular chemical reactions by electric fields, Phys. Rev. Lett. 115, 023201 (2015).
  • Haze et al. [2023] S. Haze, J. P. D’Incao, D. Dorer, J. Li, M. Deiß, E. Tiemann, P. S. Julienne, and J. Hecker Denschlag, Energy scaling of the product state distribution for three-body recombination of ultracold atoms, Phys. Rev. Res. 5, 013161 (2023).
  • Blackley et al. [2013] C. L. Blackley, C. R. Le Sueur, J. M. Hutson, D. J. McCarron, M. P. Köppinger, H.-W. Cho, D. L. Jenkin, and S. L. Cornish, Feshbach resonances in ultracold 85Rb, Phys. Rev. A 87, 033611 (2013).
  • Köhler et al. [2006] T. Köhler, K. Góral, and P. S. Julienne, Production of cold molecules via magnetically tunable Feshbach resonances, Rev. Mod. Phys. 78, 1311 (2006).
  • [49] In order to meet this regime, certain conditions need to be met. First, the considered molecular state must be bound weakly enough, so that the molecular size is not much smaller than the van der Waals length. Second, spin admixtures due to spin-exchange interaction in the two-atom collision wavefunction need to be negligible for atomic distances larger than about one half of the van der waals length. These conditions are well fulfilled for weakly bound states for Rb, but not for Li [Li2022].
  • [50] Furthermore, the level has rotational angular momentum LR=0subscript𝐿𝑅0{L_{R}}=0italic_L start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT = 0 and its vibrational quantum number is v=3v3\text{v}=-3v = - 3, counting down from the fa=fb=3subscript𝑓𝑎subscript𝑓𝑏3f_{a}=f_{b}=3italic_f start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT = italic_f start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT = 3 atomic threshold, starting with v=1v1\text{v}=-1v = - 1 for the most weakly-bound state. The vibrational levels for |ket|\uparrow\rangle| ↑ ⟩ states are counted analogously, however, starting from the fa=fb=2subscript𝑓𝑎subscript𝑓𝑏2f_{a}=f_{b}=2italic_f start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT = italic_f start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT = 2 atomic threshold.
  • [51] This argument is closely linked to a spin conservation propensity rule that was observed in recent work of ours [29], stating that the three-body recombination process for Rb conserves the hyperfine spin state of the atom pair forming the molecule.
  • Fedichev et al. [1996] P. O. Fedichev, M. W. Reynolds, and G. V. Shlyapnikov, Three-body recombination of ultracold atoms to a weakly bound s𝑠\mathit{s}italic_s level, Phys. Rev. Lett. 77, 2921 (1996).
  • D’Incao et al. [2004] J. P. D’Incao, H. Suno, and B. D. Esry, Limits on universality in ultracold three-boson recombination, Phys. Rev. Lett. 93, 123201 (2004).
  • Rem et al. [2013] B. S. Rem, A. T. Grier, I. Ferrier-Barbut, U. Eismann, T. Langen, N. Navon, L. Khaykovich, F. Werner, D. S. Petrov, F. Chevy, and C. Salomon, Lifetime of the Bose gas with resonant interactions, Phys. Rev. Lett. 110, 163202 (2013).
  • Wang et al. [2011] J. Wang, J. P. D’Incao, and C. H. Greene, Numerical study of three-body recombination for systems with many bound states, Phys. Rev. A 84, 052721 (2011).
  • Chapurin et al. [2019] R. Chapurin, X. Xie, M. J. Van de Graaff, J. S. Popowski, J. P. D’Incao, P. S. Julienne, J. Ye, and E. A. Cornell, Precision test of the limits to universality in few-body physics, Phys. Rev. Lett. 123, 233402 (2019).
  • [57] For the scattering calculations we need to calculate the three-body potentials and diabatic couplings up to a hyperradius at least an order of magnitude larger than the scattering length, all at a high resolution in length. Therefore, performing scattering calculations become increasingly difficult as we approach the Feshbach resonance, where the scattering length diverges.
  • Zaccanti et al. [2009] M. Zaccanti, B. Deissler, C. D’Errico, M. Fattori, M. Jona-Lasinio, S. Müller, G. Roati, M. Inguscio, and G. Modugno, Observation of an Efimov spectrum in an atomic system, Nat. Phys. 5, 586 (2009).
  • Li et al. [2024] J. Li, P. S. Julienne, J. Hecker Denschlag, and J. P. D’Incao, Spin hierarchy in van der Waals molecule formation via ultracold three-body recombination (2024), arXiv:2407.18567 [physics.atom-ph] .
  • Suno et al. [2002] H. Suno, B. D. Esry, C. H. Greene, and J. P. Burke, Three-body recombination of cold helium atoms, Phys. Rev. A 65, 042725 (2002).
  • Strauss et al. [2010] C. Strauss, T. Takekoshi, F. Lang, K. Winkler, R. Grimm, J. Hecker Denschlag, and E. Tiemann, Hyperfine, rotational, and vibrational structure of the a3Σu+superscript𝑎3superscriptsubscriptΣ𝑢{a}^{3}{\Sigma}_{u}^{+}italic_a start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT roman_Σ start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT state of Rb87superscriptRb87{}^{87}\mathrm{Rb}start_FLOATSUPERSCRIPT 87 end_FLOATSUPERSCRIPT roman_Rb2Phys. Rev. A 82, 052514 (2010).
  • [62] J.-L. Li, P. S. Julienne, J. Hecker Denschlag, and J. P. D’Incao, Spin structure of diatomic van der Waals molecules of alkali atoms, in preparation.

IV Methods

IV.1 Preparation of ultracold atomic sample

The experimental sequence starts with capturing 85Rb atoms in a magneto-optical trap. After a magnetic transport over 40cm the atoms are subsequently loaded into an optical dipole trap where evaporative cooling is performed. They are then transported to the center of the Paul trap via a moving 1D-optical lattice. At the final stage of the sample preparation, the atoms are confined in a far-detuned crossed-dipole trap formed by 1064  nm lasers. The trapping frequency is ωx,y,z=2π×(156,148,18)Hzsubscript𝜔𝑥𝑦𝑧2𝜋15614818Hz\omega_{x,y,z}=2\pi\times(156,148,18)\>\textrm{Hz}italic_ω start_POSTSUBSCRIPT italic_x , italic_y , italic_z end_POSTSUBSCRIPT = 2 italic_π × ( 156 , 148 , 18 ) Hz. The resulting atom cloud consists of a pure sample in (fi,mfi)=(2,2)subscript𝑓𝑖subscript𝑚subscript𝑓𝑖22(f_{i},m_{f_{i}})=(2,-2)( italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT ) = ( 2 , - 2 ) hyperfine spin state with the typical particle number of 2.5 ×\times× 105. The temperature of atoms is 860  nK. This temperature was chosen as it provided the strongest recombination signals at a reasonably cold temperature.

IV.2 REMPI detection

In order to state-selectively detect the product molecules, we apply two-step resonance-enhanced multiphoton ionization (REMPI) with a cw-laser which has a linewidth of 1MHzabsent1MHz\approx 1\>\textrm{MHz}≈ 1 MHz. The laser beam is roughly an equal mixture of σ𝜎\sigmaitalic_σ- and π𝜋\piitalic_π-polarized light. It has a power of 100 mW and a beam waist (1/e21superscript𝑒21/e^{2}1 / italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT radius) of 0.1mm at the location of the atom cloud. We use identical photons for the two REMPI steps at a wavelength around 602 nm. For Figs. 2 and 3 the intermediate REMPI states are levels of (2)1Σu+superscript21subscriptsuperscriptΣ𝑢(2)^{1}\Sigma^{+}_{u}( 2 ) start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT roman_Σ start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT with J=3superscript𝐽3J^{\prime}=3italic_J start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = 3 for |ket|\uparrow\rangle| ↑ ⟩ states and J=1superscript𝐽1J^{\prime}=1italic_J start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = 1 for |ket|\downarrow\rangle| ↓ ⟩ states [29], where Jsuperscript𝐽J^{\prime}italic_J start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT represents the total angular momentum excluding nuclear spin. The J=1superscript𝐽1J^{\prime}=1italic_J start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = 1 and J=3superscript𝐽3J^{\prime}=3italic_J start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = 3 levels are split by 2.9 GHz. The photoassociation laser frequency towards the intermediate level J=1superscript𝐽1J^{\prime}=1italic_J start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = 1 is ν0=497603.591subscript𝜈0497603.591\nu_{0}=497603.591italic_ν start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 497603.591 GHz at B=4.6G𝐵4.6GB=4.6\>\textrm{G}italic_B = 4.6 G. The binding energies of the experimentally observed molecular states in Figs. 2 and 3 are 4.7GHz×h4.7GHz4.7\>\textrm{GHz}\times h4.7 GHz × italic_h for |ket|\uparrow\rangle| ↑ ⟩, and span a range between 6.4 to 7.3GHz×hGHz\>\textrm{GHz}\times hGHz × italic_h for |ket|\downarrow\rangle| ↓ ⟩. Here, the binding energy is determined relative to the B𝐵Bitalic_B-field dependent (4,2,2) threshold. For Fig. 4, the intermediate REMPI states are deeply-bound levels of (2)3Π  0g+superscript23Πsuperscriptsubscript  0𝑔(2)^{3}\Pi\,\,0_{g}^{+}( 2 ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT roman_Π 0 start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT with J=1,3,5superscript𝐽135J^{\prime}=1,3,5italic_J start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = 1 , 3 , 5 [29]. Here, ν0=497831.928subscript𝜈0497831.928\nu_{0}=497831.928italic_ν start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 497831.928 GHz is the photoassociation frequency towards J=1superscript𝐽1J^{\prime}=1italic_J start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = 1 at B=4.6G𝐵4.6GB=4.6\>\textrm{G}italic_B = 4.6 G. The binding energies of the molecular states observed in Figs. 4 (a) and (b) span a range between 0.60.60.60.6 and 12.6GHz×h12.6GHz12.6\>\textrm{GHz}\times h12.6 GHz × italic_h. Again, the binding energy is determined relative to the (4,2,2) threshold. In general, the Zeeman effects of our intermediate states are negligible compared to the ones of the ground state. We make an effort to ensure that the REMPI efficiencies are similar for the states that we probe, also at various magnetic fields, but a precise calibration of the REMPI efficiency has not been done yet. We note that the first REMPI step is generally not saturated.

When ions are produced via REMPI, they are directly trapped and detected in an eV-deep Paul trap which is centered on the atom cloud. Elastic atom-ion collisions inflict tell-tale atom loss while the ions remain trapped. From the atom loss which is measured via absorption imaging of the atom cloud the ion number can be inferred, for details see [29]. From the ion numbers and the interaction time we obtain an ion production rate (i.e. the molecular detection rate) which is generally proportional to the state-selective molecular production rate and the three-body recombination loss rate constant.

IV.3 Model calculations

Our numerical simulations use the adiabatic hyperspherical representation approach where the coordinates of three particles are given in terms of the hyperradius R𝑅Ritalic_R for the overall size of the system and a set of hyperangles ΩΩ\Omegaroman_Ω for the internal motion [60, 55, 56, 39]. The three-body Schrödinger equation is solved by adiabatically separating the hyperradial motion

[22μd2dR2+Uν(R)]Fν(R)delimited-[]superscriptPlanck-constant-over-2-pi22𝜇superscript𝑑2𝑑superscript𝑅2subscript𝑈𝜈𝑅subscript𝐹𝜈𝑅\displaystyle\left[-\frac{\hbar^{2}}{2\mu}\frac{d^{2}}{dR^{2}}+U_{\nu}(R)% \right]F_{\nu}(R)[ - divide start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_μ end_ARG divide start_ARG italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_d italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + italic_U start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( italic_R ) ] italic_F start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( italic_R )
+νWνν(R)Fν(R)=EFν(R),subscriptsuperscript𝜈subscript𝑊𝜈superscript𝜈𝑅subscript𝐹superscript𝜈𝑅𝐸subscript𝐹𝜈𝑅\displaystyle~{}~{}~{}~{}~{}~{}+\sum_{\nu^{\prime}}W_{\nu\nu^{\prime}}(R)F_{% \nu^{\prime}}(R)=EF_{\nu}(R),+ ∑ start_POSTSUBSCRIPT italic_ν start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT italic_W start_POSTSUBSCRIPT italic_ν italic_ν start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( italic_R ) italic_F start_POSTSUBSCRIPT italic_ν start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( italic_R ) = italic_E italic_F start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( italic_R ) , (1)

from the internal motion

H^adΦν(R;Ω)=Uν(R)Φν(R;Ω),subscript^𝐻adsubscriptΦ𝜈𝑅Ωsubscript𝑈𝜈𝑅subscriptΦ𝜈𝑅Ω\hat{H}_{\rm ad}\Phi_{\nu}(R;\Omega)=U_{\nu}(R)\Phi_{\nu}(R;\Omega),over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT roman_ad end_POSTSUBSCRIPT roman_Φ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( italic_R ; roman_Ω ) = italic_U start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( italic_R ) roman_Φ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( italic_R ; roman_Ω ) , (2)

where the hyperradius R𝑅Ritalic_R appears only as a parameter. The diagonalization of the hyperangular adiabatic Hamiltonian H^adsubscript^𝐻ad\hat{H}_{\rm ad}over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT roman_ad end_POSTSUBSCRIPT gives the three-body potentials Uνsubscript𝑈𝜈U_{\nu}italic_U start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT and the channel functions ΦνsubscriptΦ𝜈\Phi_{\nu}roman_Φ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT, which are also used for computing the nonadiabatic couplings Wννsubscript𝑊𝜈superscript𝜈W_{\nu\nu^{\prime}}italic_W start_POSTSUBSCRIPT italic_ν italic_ν start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT, for the hyperradial equation.

In our model, the hyperangular adiabatic Hamiltonian reads

H^ad=Λ^2(Ω)+15/42μR22+i,j=a,b,cijV^ab(R,Ω)+i=a,b,cH^isp(B),subscript^𝐻adsuperscript^Λ2Ω1542𝜇superscript𝑅2superscriptPlanck-constant-over-2-pi2subscriptformulae-sequence𝑖𝑗𝑎𝑏𝑐𝑖𝑗subscript^𝑉𝑎𝑏𝑅Ωsubscript𝑖𝑎𝑏𝑐subscriptsuperscript^𝐻sp𝑖𝐵\displaystyle\hat{H}_{\rm ad}=\frac{\hat{\Lambda}^{2}(\Omega)+15/4}{2\mu R^{2}% }\hbar^{2}+\sum_{\begin{subarray}{c}i,j=a,b,c\\ i\neq j\end{subarray}}\hat{V}_{ab}(R,\Omega)+\sum_{i=a,b,c}\hat{H}^{\rm sp}_{i% }(B),over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT roman_ad end_POSTSUBSCRIPT = divide start_ARG over^ start_ARG roman_Λ end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( roman_Ω ) + 15 / 4 end_ARG start_ARG 2 italic_μ italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + ∑ start_POSTSUBSCRIPT start_ARG start_ROW start_CELL italic_i , italic_j = italic_a , italic_b , italic_c end_CELL end_ROW start_ROW start_CELL italic_i ≠ italic_j end_CELL end_ROW end_ARG end_POSTSUBSCRIPT over^ start_ARG italic_V end_ARG start_POSTSUBSCRIPT italic_a italic_b end_POSTSUBSCRIPT ( italic_R , roman_Ω ) + ∑ start_POSTSUBSCRIPT italic_i = italic_a , italic_b , italic_c end_POSTSUBSCRIPT over^ start_ARG italic_H end_ARG start_POSTSUPERSCRIPT roman_sp end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_B ) , (3)

where Λ^^Λ\hat{\Lambda}over^ start_ARG roman_Λ end_ARG denotes the hyperangular momentum operator [60, 55] and μ=m/3𝜇𝑚3\mu=m/\sqrt{3}italic_μ = italic_m / square-root start_ARG 3 end_ARG is the reduced mass of three identical atoms of mass m𝑚mitalic_m. The atomic spin Hamiltonian H^ispsubscriptsuperscript^𝐻sp𝑖\hat{H}^{\rm sp}_{i}over^ start_ARG italic_H end_ARG start_POSTSUPERSCRIPT roman_sp end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT for atom i𝑖iitalic_i contains the hyperfine and Zeeman interaction, and to a very good approximation within the present work its eigenstates are |fi,mfiketsubscript𝑓𝑖subscript𝑚subscript𝑓𝑖|f_{i},m_{f_{i}}\rangle| italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT ⟩. For two Rb atoms (e.g., i𝑖iitalic_i and j𝑗jitalic_j) of the 5S1/2+5S1/25subscript𝑆125subscript𝑆125S_{1/2}+5S_{1/2}5 italic_S start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT + 5 italic_S start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT asymptote, the pairwise interaction V^ijsubscript^𝑉𝑖𝑗\hat{V}_{ij}over^ start_ARG italic_V end_ARG start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT can be expressed in terms of the electronic singlet and triplet Born-Oppenheimer potentials. We use the potentials from Ref. [61] with an additional repulsive term C/rij12𝐶subscriptsuperscript𝑟12𝑖𝑗C/r^{12}_{ij}italic_C / italic_r start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT to reduce the number of bound states in our simulation. Here, rijsubscript𝑟𝑖𝑗r_{ij}italic_r start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT is the interatomic distance. Removing deeply bound states mitigates the computational hardship without affecting too much the results, as generally more deeply bound states play a less important role in the three-body recombination process [46]. The truncation of the potentials shall be explained in more detail in a separate publication [62]. In brief, two parameters C𝐶Citalic_C (Cssubscript𝐶sC_{\mathrm{s}}italic_C start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT and Ctsubscript𝐶tC_{\mathrm{t}}italic_C start_POSTSUBSCRIPT roman_t end_POSTSUBSCRIPT) are adjusted individually for the truncated singlet and triplet potentials so that they contain 6 and 5 s𝑠sitalic_s-wave bound states, respectively, and so that the known singlet and triplet scattering lengths are reproduced. Additional fine-tuning of the two C𝐶Citalic_C parameters together with the atomic hyperfine splitting aims at reproducing the Feshbach resonance at about 155 G. As a result, the atomic hyperfine splitting is reduced by about 5%percent55\%5 % compared to the literature value. We use Cs=(0.3242030rvdw)6C6subscript𝐶ssuperscript0.3242030subscript𝑟vdw6subscript𝐶6C_{\mathrm{s}}=(0.3242030\>r_{\mathrm{vdw}})^{6}\cdot C_{6}italic_C start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT = ( 0.3242030 italic_r start_POSTSUBSCRIPT roman_vdw end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT ⋅ italic_C start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT and Ct=(0.3258900rvdw)6C6subscript𝐶tsuperscript0.3258900subscript𝑟vdw6subscript𝐶6C_{\mathrm{t}}=(0.3258900\>r_{\mathrm{vdw}})^{6}\cdot C_{6}italic_C start_POSTSUBSCRIPT roman_t end_POSTSUBSCRIPT = ( 0.3258900 italic_r start_POSTSUBSCRIPT roman_vdw end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT ⋅ italic_C start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT, where rvdw=12(mC62)1/4subscript𝑟vdw12superscript𝑚subscript𝐶6superscriptPlanck-constant-over-2-pi214r_{\mathrm{vdw}}=\frac{1}{2}(\frac{mC_{6}}{\hbar^{2}})^{1/4}italic_r start_POSTSUBSCRIPT roman_vdw end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( divide start_ARG italic_m italic_C start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT end_ARG start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) start_POSTSUPERSCRIPT 1 / 4 end_POSTSUPERSCRIPT is the van der Waals length and C6subscript𝐶6C_{6}italic_C start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT is the van der Waals coefficient.

Interactions between the particles and with the external magnetic field B𝐵Bitalic_B couple various angular momenta. Therefore, the incoming spin channel |2,2|2,2|2,2ket22ket22ket22|2,-2\rangle|2,-2\rangle|2,-2\rangle| 2 , - 2 ⟩ | 2 , - 2 ⟩ | 2 , - 2 ⟩ can in principle be coupled to a range of spin channels |fa,mfa|fb,mfb|fc,mfcketsubscript𝑓𝑎subscript𝑚subscript𝑓𝑎ketsubscript𝑓𝑏subscript𝑚subscript𝑓𝑏ketsubscript𝑓𝑐subscript𝑚subscript𝑓𝑐|f_{a},m_{f_{a}}\rangle|f_{b},m_{f_{b}}\rangle|f_{c},m_{f_{c}}\rangle| italic_f start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT end_POSTSUBSCRIPT ⟩ | italic_f start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT end_POSTSUBSCRIPT ⟩ | italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_POSTSUBSCRIPT ⟩, where fisubscript𝑓𝑖f_{i}italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT can be 2 or 3. We essentially only have the restriction that Mtot=mfa+mfb+mfcsubscript𝑀totsubscript𝑚subscript𝑓𝑎subscript𝑚subscript𝑓𝑏subscript𝑚subscript𝑓𝑐M_{\rm tot}=m_{f_{a}}+m_{f_{b}}+m_{f_{c}}italic_M start_POSTSUBSCRIPT roman_tot end_POSTSUBSCRIPT = italic_m start_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT end_POSTSUBSCRIPT + italic_m start_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT end_POSTSUBSCRIPT + italic_m start_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_POSTSUBSCRIPT is conserved, as long as spin-spin interaction can be neglected. However, motivated by previous work [29], where we found a spin conservation propensity rule in three-body recombination of Rb atoms we restrict the spin of the third atom to be |fc,mfc=|2,2ketsubscript𝑓𝑐subscript𝑚subscript𝑓𝑐ket22|f_{c},m_{f_{c}}\rangle=|2,-2\rangle| italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_POSTSUBSCRIPT ⟩ = | 2 , - 2 ⟩ in our calculations. One reason for this restriction could be that the third atom (c𝑐citalic_c) interacts mainly mechanically with the other two, (a,b𝑎𝑏a,bitalic_a , italic_b), while they are forming a molecule. This approximation leads to a model of five coupled three-body channels with the quantum numbers (F,fa,fb𝐹subscript𝑓𝑎subscript𝑓𝑏F,f_{a},f_{b}italic_F , italic_f start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT , italic_f start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT) = (4, 2, 2), (4, 3, 2), (4, 3, 3), (5, 3, 2), and (6, 3, 3).

V Supplemental Materials

V.1 Feshbach resonance at 155 Gauss

The s𝑠sitalic_s-wave Feshbach resonance used in this work is located at 155.3  G. It couples the incoming (F,fa,fb,mF)=(4,2,2,4)𝐹subscript𝑓𝑎subscript𝑓𝑏subscript𝑚𝐹4224(F,f_{a},f_{b},m_{F})=(4,2,2,-4)( italic_F , italic_f start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT , italic_f start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) = ( 4 , 2 , 2 , - 4 ) state and the closed-channel bound state (F,fa,fb,mF)=(4,3,3,4)𝐹subscript𝑓𝑎subscript𝑓𝑏subscript𝑚𝐹4334(F,f_{a},f_{b},m_{F})=(4,3,3,-4)( italic_F , italic_f start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT , italic_f start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) = ( 4 , 3 , 3 , - 4 ). The scattering length across the Feshbach resonance is well characterized by the relation a(B)=abg(1ΔBBB0)𝑎𝐵subscript𝑎bg1Δ𝐵𝐵subscript𝐵0a(B)=a_{\textrm{bg}}(1-\frac{\Delta B}{B-B_{0}})italic_a ( italic_B ) = italic_a start_POSTSUBSCRIPT bg end_POSTSUBSCRIPT ( 1 - divide start_ARG roman_Δ italic_B end_ARG start_ARG italic_B - italic_B start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ). Here, abg=443a0subscript𝑎bg443subscript𝑎0a_{\textrm{bg}}=-443a_{0}italic_a start_POSTSUBSCRIPT bg end_POSTSUBSCRIPT = - 443 italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, ΔB=10.9GΔ𝐵10.9G\Delta B=10.9\>\textrm{G}roman_Δ italic_B = 10.9 G, B0=155.3Gsubscript𝐵0155.3GB_{0}=155.3\>\textrm{G}italic_B start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 155.3 G is the background scattering length, the width and the position of the resonance [47]. The scattering length is shown in Fig. S1.

Refer to caption
Fig. S 1: Scattering length in the vicinity of the s𝑠sitalic_s-wave Feshbach resonance. The scattering length in units of Bohr radius is plotted as a function of magnetic field.

V.2 Temperature dependence of L3subscript𝐿3L_{3}italic_L start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT

Refer to caption
Fig. S 2: The three-body recombination rate constant L3subscript𝐿3L_{3}italic_L start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT for the |(3,4)ket34|\uparrow\rangle(-3,4)| ↑ ⟩ ( - 3 , 4 ) state at 80808080 nK is compared to that at 860860860860 nK. The dashed line indicates the L3a4proportional-tosubscript𝐿3superscript𝑎4L_{3}\propto a^{4}italic_L start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ∝ italic_a start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT scaling.

The scaling of L3a4proportional-tosubscript𝐿3superscript𝑎4L_{3}\propto a^{4}italic_L start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ∝ italic_a start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT is better perceivable at lower temperatures. In Fig. S2, we compare the L3subscript𝐿3L_{3}italic_L start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT for the |(3,4)ket34|\uparrow\rangle(-3,4)| ↑ ⟩ ( - 3 , 4 ) state at 80 nK with the 860 nK result presented in the main manuscript. The Efimov resonance at about 140 G perturbs the overall scaling. At 150 G the L3subscript𝐿3L_{3}italic_L start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT value for 80 nK is again close to the a4superscript𝑎4a^{4}italic_a start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT prediction.