control of ultracold photodissociation with magnetic fields · quantum chemistry model that...

8
Control of Ultracold Photodissociation with Magnetic Fields M. McDonald, 1, * I. Majewska, 2 C.-H. Lee, 1 S. S. Kondov, 1 B. H. McGuyer, 1, R. Moszynski, 2 and T. Zelevinsky 1, 1 Department of Physics, Columbia University, 538 West 120th Street, New York, NY 10027-5255, USA 2 Quantum Chemistry Laboratory, Department of Chemistry, University of Warsaw, Pasteura 1, 02-093 Warsaw, Poland (Dated: September 15, 2017) Photodissociation of a molecule produces a spatial distribution of photofragments determined by the molecular structure and the characteristics of the dissociating light. Performing this basic chem- ical reaction at ultracold temperatures allows its quantum mechanical features to dominate. In this regime, weak applied fields can be used to control the reaction. Here, we photodissociate ultracold diatomic strontium in magnetic fields below 10 G and observe striking changes in photofragment angular distributions. The observations are in excellent qualitative agreement with a multichannel quantum chemistry model that includes nonadiabatic effects and predicts strong mixing of partial waves in the photofragment energy continuum. The experiment is enabled by precise quantum-state control of the molecules. Chemical reactions at cold and ultracold temperatures exhibit quantum mechanical behavior, since at these low kinetic energies reactions possess a strong sensitiv- ity to the details of intermolecular interactions. More- over, when the reactants are prepared at ultracold tem- peratures, their internal quantum states can be well con- trolled, leading to a much greater understanding of the reaction and potentially enabling a complete theoretical description. When a reaction proceeds at such low tem- peratures, it becomes possible to control its outcome by applying modest electric or magnetic fields. This occurs because the size of Stark or Zeeman shifts can be much greater than the kinetic energy [1], and the density of molecular states is high near the threshold, facilitating mixing by external fields [2]. Field control of diatomic- molecule collisions and reactions has been investigated recently for polar molecules, largely focusing on rate con- stants [3–5]. Photodissociation of a diatomic molecule is a basic chemical reaction where a bond breaks under the influ- ence of light. It is related to photoassociation of an atom pair [6] by time reversal, but has advantages for stud- ies of ultracold chemistry. In photodissociation, thermal averaging of the atomic collision energies is avoided and the internal and motional states of the initial molecules can be precisely engineered, leading to fully quantum- state-controlled reactions and strictly nonclassical phe- nomena such as matter-wave interference of the reaction products [7]. Here, we photodissociate ultracold diatomic strontium molecules, 88 Sr 2 , and induce dramatic changes in reaction outcomes by applying magnetic fields. The study of photodissociation in the ultracold regime and in the presence of external fields requires us to explicitly include field-induced angular-momentum mixing into the theoretical treatment of this process. While the theory of photodissociation has been extensively developed [8–11] including the effects of magnetic fields [12], previously the total angular momentum was considered a conserved quantum number. In the regime explored here, this is no longer the case. Combined with a multichannel quantum- chemistry molecular model [13, 14], the theoretical treat- ment we have developed here faithfully reproduces all our experimental observations. In the experiment we directly observe and record the photofragment angular distributions (PADs) in the mil- likelvin energy regime. The molecules are prepared at microkelvin temperatures in an optical lattice, and are subsequently fragmented with laser light [7]. The one- dimensional lattice is a standing wave of far-off-resonant light at 910 nm and is approximately 1 MHz (or 50 μK) deep. The geometry of the setup is defined in Fig. 1(a). Photodissociation results in two counter- propagating photofragments, an atom in the ground state 1 S 0 and an atom in the electronically excited state 3 P 1 which decays to 1 S 0 with a 10 μs lifetime. These atoms are absorption imaged using a charge-coupled device camera on the strong Sr transition at 461 nm. The imag- ing light is turned on for a short duration of 10 μs, at a time τ (between 250 and 600 μs) after the 20-50 μs photodissociation pulse at 689 nm. During this time, the photofragments freely expand and effectively form spherical shells with radii determined by the frequency of the photodissociation light and the Zeeman shifts of the atomic continua. The camera is nearly on-axis with the lattice, thus capturing a two-dimensional projection of the spherical shells since the atoms effectively origi- nate from a point source. The laboratory quantum axis points along the applied magnetic field ~ B, which has a vertical orientation that defines the polar angle θ and az- imuthal angle φ. The dependence of the photofragment density on these angles is our key observable and encodes the quantum mechanics of the reaction. The photodis- sociation light polarization is set to be either vertical or horizontal. Figure 1(b) illustrates the Sr 2 molecular structure rel- evant to this work. The molecules are created from ultra- cold atoms via photoassociation [15] in the least-bound vibrational level, denoted by v = -1, of the electronic arXiv:1709.04527v1 [physics.atom-ph] 13 Sep 2017

Upload: others

Post on 06-Jul-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Control of Ultracold Photodissociation with Magnetic Fields · quantum chemistry model that includes nonadiabatic e ects and predicts strong mixing of partial waves in the photofragment

Control of Ultracold Photodissociation with Magnetic Fields

M. McDonald,1, ∗ I. Majewska,2 C.-H. Lee,1 S. S. Kondov,1 B. H. McGuyer,1, † R. Moszynski,2 and T. Zelevinsky1, ‡

1Department of Physics, Columbia University, 538 West 120th Street, New York, NY 10027-5255, USA2Quantum Chemistry Laboratory, Department of Chemistry,University of Warsaw, Pasteura 1, 02-093 Warsaw, Poland

(Dated: September 15, 2017)

Photodissociation of a molecule produces a spatial distribution of photofragments determined bythe molecular structure and the characteristics of the dissociating light. Performing this basic chem-ical reaction at ultracold temperatures allows its quantum mechanical features to dominate. In thisregime, weak applied fields can be used to control the reaction. Here, we photodissociate ultracolddiatomic strontium in magnetic fields below 10 G and observe striking changes in photofragmentangular distributions. The observations are in excellent qualitative agreement with a multichannelquantum chemistry model that includes nonadiabatic effects and predicts strong mixing of partialwaves in the photofragment energy continuum. The experiment is enabled by precise quantum-statecontrol of the molecules.

Chemical reactions at cold and ultracold temperaturesexhibit quantum mechanical behavior, since at theselow kinetic energies reactions possess a strong sensitiv-ity to the details of intermolecular interactions. More-over, when the reactants are prepared at ultracold tem-peratures, their internal quantum states can be well con-trolled, leading to a much greater understanding of thereaction and potentially enabling a complete theoreticaldescription. When a reaction proceeds at such low tem-peratures, it becomes possible to control its outcome byapplying modest electric or magnetic fields. This occursbecause the size of Stark or Zeeman shifts can be muchgreater than the kinetic energy [1], and the density ofmolecular states is high near the threshold, facilitatingmixing by external fields [2]. Field control of diatomic-molecule collisions and reactions has been investigatedrecently for polar molecules, largely focusing on rate con-stants [3–5].

Photodissociation of a diatomic molecule is a basicchemical reaction where a bond breaks under the influ-ence of light. It is related to photoassociation of an atompair [6] by time reversal, but has advantages for stud-ies of ultracold chemistry. In photodissociation, thermalaveraging of the atomic collision energies is avoided andthe internal and motional states of the initial moleculescan be precisely engineered, leading to fully quantum-state-controlled reactions and strictly nonclassical phe-nomena such as matter-wave interference of the reactionproducts [7]. Here, we photodissociate ultracold diatomicstrontium molecules, 88Sr2, and induce dramatic changesin reaction outcomes by applying magnetic fields. Thestudy of photodissociation in the ultracold regime andin the presence of external fields requires us to explicitlyinclude field-induced angular-momentum mixing into thetheoretical treatment of this process. While the theory ofphotodissociation has been extensively developed [8–11]including the effects of magnetic fields [12], previouslythe total angular momentum was considered a conservedquantum number. In the regime explored here, this is no

longer the case. Combined with a multichannel quantum-chemistry molecular model [13, 14], the theoretical treat-ment we have developed here faithfully reproduces all ourexperimental observations.

In the experiment we directly observe and record thephotofragment angular distributions (PADs) in the mil-likelvin energy regime. The molecules are prepared atmicrokelvin temperatures in an optical lattice, and aresubsequently fragmented with laser light [7]. The one-dimensional lattice is a standing wave of far-off-resonantlight at 910 nm and is approximately 1 MHz (or 50µK) deep. The geometry of the setup is defined inFig. 1(a). Photodissociation results in two counter-propagating photofragments, an atom in the ground state1S0 and an atom in the electronically excited state 3P1

which decays to 1S0 with a 10 µs lifetime. These atomsare absorption imaged using a charge-coupled devicecamera on the strong Sr transition at 461 nm. The imag-ing light is turned on for a short duration of ∼ 10 µs, ata time τ (between 250 and 600 µs) after the 20-50 µsphotodissociation pulse at 689 nm. During this time,the photofragments freely expand and effectively formspherical shells with radii determined by the frequencyof the photodissociation light and the Zeeman shifts ofthe atomic continua. The camera is nearly on-axis withthe lattice, thus capturing a two-dimensional projectionof the spherical shells since the atoms effectively origi-nate from a point source. The laboratory quantum axispoints along the applied magnetic field ~B, which has avertical orientation that defines the polar angle θ and az-imuthal angle φ. The dependence of the photofragmentdensity on these angles is our key observable and encodesthe quantum mechanics of the reaction. The photodis-sociation light polarization is set to be either vertical orhorizontal.

Figure 1(b) illustrates the Sr2 molecular structure rel-evant to this work. The molecules are created from ultra-cold atoms via photoassociation [15] in the least-boundvibrational level, denoted by v = −1, of the electronic

arX

iv:1

709.

0452

7v1

[ph

ysic

s.at

om-p

h] 1

3 Se

p 20

17

Page 2: Control of Ultracold Photodissociation with Magnetic Fields · quantum chemistry model that includes nonadiabatic e ects and predicts strong mixing of partial waves in the photofragment

2

PDz

y

x

(a)

Pot

enti

al E

nerg

y

Internuclear Separation

(b)

10.15 G0.5 G

(c)

Exp

.T

h.

1S0+3P1

1S0+1S0

FIG. 1. (a) Geometry of the photodissociation process. Themolecules are trapped in an optical lattice at the origin, whilethe photodissociation (PD) laser propagates along the x axis.The polar angle θ and azimuthal angle φ are defined as shownto describe the photofragment angular distributions (PADs).The radii of the spherical shells containing the fragments af-ter a fixed expansion time are defined by the frequency of thePD laser and the Zeeman shifts of different atomic continua.The largest shell corresponds to a negative shift (yellow), themedium shell to an absence of shift (red), and the smallestshell to a positive shift (blue). A camera points in the −x di-rection and images a two-dimensional projection of the nestedshells. (b) Molecular potentials and quantum states relevantto the experiment. The photodissociation process is desig-nated by the double arrow. The detuning of the PD lightfrom the m = 0 Zeeman component of the continuum is ∆,and the symmetric Zeeman splitting has a magnitude ∆Z .The barrier of the 1u potential has a height of ∼ 30 MHz.The numbers in parentheses are v and Ji. (c) An exampleof calculated and measured PAD images for a process wherea small applied magnetic field drastically alters the outcomeof the reaction. The two pairs of images differ only by themagnitude of the applied field: B = 0.5 G on the left and10.15 G on the right.

ground state X1Σ+g (correlating to the 1S0+1S0 atomic

threshold). Initially, the molecules occupy two rotationalstates with the total angular-momentum quantum num-bers Ji = 0 and 2, but the Ji = 2 population is mostlyremoved prior to fragmentation by a laser pulse resonantwith an excited molecular state. The Ji = 0 molecules(with a projection quantum number Mi = 0) are cou-pled by the photodissociation laser to the singly-excitedcontinuum above the 0u and 1u ungerade potentials (cor-relating to the 1S0+3P1 atomic threshold), where thenumbers refer to the total atomic angular momentumprojections onto the molecular axis. Under an appliedfield B > 0, the atomic energy levels split by the Zeemaninteraction into the m = −1, 0, and 1 sublevels, where

the energy separation between the neighboring sublevelsis h∆Z = 1.5µBB and µB is the Bohr magneton. Theradius of each photofragment shell is vτ where the veloc-ity v =

√h(∆−m∆Z)/mSr, h is the Planck constant, ∆

is the frequency detuning of the photodissociation lightfrom the m = 0 component of the continuum, and mSr

is the atomic mass of Sr.

If the photodissociation laser detuning is large andnegative, ∆ < −∆Z , no photofragments should be de-tectable because the target energy is below the low-est threshold. If the detuning is small and negative,−∆Z < ∆ < 0, then only one fragment shell shouldbe visible, corresponding to m = −1. If the detuningis small and positive, 0 < ∆ < ∆Z , we expect to ob-serve two fragment shells, with m = −1 and 0. Finally,if the detuning is large and positive, ∆Z < ∆, we ex-pect three fragment shells with all possible values of m.This is the case in the example of Fig. 1(c) that showsa strong alteration of the PAD for B = 10.15 G com-pared to 0.5 G. Here we make the distinction between theangular-momentum projection quantum numbers m andM , the latter denoting the projection of the total angu-lar momentum J in the continuum. Electric-dipole (E1)selection rules require M = Mi = 0 if the photodissocia-tion laser polarization is parallel to the quantum axis andM = ±1 if the polarization is perpendicular. In contrast,there are no such selection rules for the atomic magneticsublevels m which can be superpositions of several M .

When a Ji = 0 diatomic molecule is photodissociatedvia a one-photon E1 process without an applied field, weexpect and observe a dipolar-shaped PAD with an axisset by the laser polarization [7], as in the nearly field-freecase of Fig. 1(c). This can be understood either by visu-alizing a spherically symmetric molecule absorbing lightwith a dipolar probability distribution, or by applyingangular-momentum selection rules that require J = 1 forthe outgoing channel, which has a dipolar angular distri-bution with a single spatial node. We find that with anonzero B this is no longer the case, and instead observecomplicated structures with multiple nodes.

The main results of the experiment and theory aresummarized in Fig. 2. The two-dimensional projectionsof the PADs onto the imaging plane, with the detuning∆ = 29.2 MHz, are shown in Fig. 2(a) for a progressionof magnetic fields B from 0.5 to 10.15 G. The removalof the Ji = 2 molecules is imperfect which results in thefaint outermost shell that can be ignored. The top pairof rows corresponds to parallel light polarization and thebottom pair to perpendicular polarization. We observea transformation from simple dipolar patterns at B = 0to more complex patterns that exhibit a multiple-nodestructure at 10.15 G. Figure 2(b) shows PADs that areobserved when B is kept fixed at 10.15 G while ∆ is var-ied from -13.8 to 50.0 MHz, again for both cases of linearlight polarization. For the entire range of continuum en-ergies, we observe PADs that exhibit a multinode struc-

Page 3: Control of Ultracold Photodissociation with Magnetic Fields · quantum chemistry model that includes nonadiabatic e ects and predicts strong mixing of partial waves in the photofragment

3

(a)

(b)

0.5 1.55 2.78 5.24 7.69 10.15E

xp.

Th.

Exp

.T

h.

B EPD

-13.8 -2.8 +5.0 +17.2 +29.2 +50.0

Th.

Th.

Exp

.E

xp.

Magnetic Field (G)

Detuning above threshold (MHz)m = -1m = 0m = +1

B

EPD

B EPD

B

EPD

FIG. 2. Tuning of the photodissociation reaction with smallmagnetic fields, across a range of energies. The color codingfor the continuum Zeeman components is consistent betweenFigs. 1 and 2. (a) Theoretical and experimental images ofPADs as the magnetic field B is increased from 0.5 to 10.15G, for the detuning ∆ = 29.2 MHz. (b) Theoretical and ex-perimental PAD images at B = 10.15 G, covering a range of∆ from -13.8 to 50.0 MHz. As indicated in Fig. 1, additionalchannels (m = −1, 0, and 1) become available in the contin-uum as ∆ increases, leading to extra photofragment shells. Inall experimental images, the faint outermost shell is the resultof incidental photodissociation of residual J = 2 moleculesand can be ignored. The top and bottom pairs of rows inboth (a) and (b) correspond to the light polarization parallel

and perpendicular to ~B, respectively. Typically, 300 exper-iments with atoms and 300 without atoms (for backgroundsubtraction) are averaged to obtain each experimental PADimage. The experimental images use an arbitrary brightnessscale, and the relative transition strengths for different imagescan be inferred from this data only qualitatively. Within eachPAD, however, relative transition strengths to different m’sare more accurately reflected in the relative brightness of therings.

ture. As ∆ and B are varied, the angular dependence,or anisotropy, of the outgoing PAD is strongly affected.The zero-field evolution of the PADs with energy for thiscontinuum is discussed in detail in [7]. All additionalfeatures observed here are due to the continuum partialwaves J being strongly mixed by the applied field.

As Fig. 2 demonstrates, our theoretical results arein excellent qualitative agreement with the experimentaldata. The theory involves extending the standard treat-ment of diatomic photodissociation to the case of mixedangular momenta in the presence of a magnetic field, andapplying it to the quantum-chemistry model of the 88Sr2

molecule [13, 14]. As detailed in [16], the PADs can bedescribed by the expansion

I(θ, φ) ∝ β0

(1 +

∞∑µ=1

µ∑ν=0

βµνPνµ (cos θ) cos(νφ)

)(1)

where P νµ (cos θ) are the associated Legendre polynomialsand the βµν coefficients are called anisotropy parameters.In the case of parallel light polarization, the PADs arecylindrically symmetric (no φ dependence) [7], and weset βµ ≡ βµ0 while all other βµν vanish. The µ are evenfor homonuclear dimers. The anisotropy parameters inexpression (1) can be evaluated from Fermi’s golden ruleafter properly representing the initial (bound-state) andfinal (continuum) wave functions, including mixing of theangular momenta Ji and J by the magnetic field.

The most salient feature of ultracold photodissociationin nonzero magnetic fields is the dramatic change of thePADs which tend to become significantly more complexas the field is increased. Figure 3(a,b) compares the pho-todissociation outcome at B = 10.15 G (also in the topright of Fig. 2(a)) to that at B = 0. Besides the appear-ance of an outer shell caused by Zeeman splitting in thecontinuum, the central m = 0 shell gains additional lobesas compared to the purely dipolar (J = 1) pattern forB = 0. This effect arises from the magnetic field admix-ing higher partial waves in the continuum, as the densityof states is particularly high near the dissociation thresh-old [2]. We show this directly by simulating the imageof the PAD on the right panel of Fig. 3(a) while usinga series of cutoff partial waves Jmax that are includedin the continuum wave function. The result is in Fig.3(c). The PAD evolves with increasing Jmax, only repro-ducing the data at Jmax = 5. We have confirmed thatincreasing Jmax further does not alter the PAD apprecia-bly. (max = 1 if B = 0.) The evolution of the PADs withincreasing magnetic field can be alternatively describedby plotting the anisotropy parameters βµ as functions ofB. Figure 3(d-h) shows this for the PAD in Fig. 3(a), foranisotropies of order µ = 0 through 8 that we can resolvein the experiment. The curves correspond to contribu-tions from pure and mixed exit-channel partial waves ofEq. (10) in [16], with Jk, J

′k varying from 1 to 5. These

plots directly show that higher-order anisotropy (µ > 4)

Page 4: Control of Ultracold Photodissociation with Magnetic Fields · quantum chemistry model that includes nonadiabatic e ects and predicts strong mixing of partial waves in the photofragment

4

0.05

0.04

0.03

0.02

0.01

0.000.00

0.04

0.08

0.12

0.16

0.20

-0.4

-0.2

0.0

0.2

0.4

0.6

1.0

0.8

0.6

0.4

0.2

0.00 2 4 6 8 10 0 2 4 6 8 10

0 2 4 6 8 10 0 2 4 6 8 10 0 2 4 6 8 10

0.0

-0.2

-0.4

-0.6

-0.8

-1.0

-1.2

1 2 3 4 5

(d)(d)

(c)

(e)(e)

(f)(f ) (g) (h)

(a) (b)B = 10.15 G B = 0 G

FIG. 3. Mixing of the continuum partial waves with a smallmagnetic field. (a) Experimental and calculated PAD imagesfor photodissociation at B = 10.15 G. The laser polarizationis parallel to ~B and its detuning is ∆ = 29.2 MHz. Thebrightest fragment shell here corresponds to m = 0. (b) Forcomparison, a calculated PAD for B = 0 consists of a sim-ple dipolar pattern corresponding to J = 1. The quantumnumbers m = −1, 0, 1 are unresolved. (c) Calculated PADsof the process in (a), each image including contributions onlyup to the maximum partial wave Jmax = 1 through 5 thatis included in the multichannel continuum wave function, aslabeled. Agreement with experiment is reached for Jmax = 5.(d-h) Plots of the anisotropy parameters βµ versus the fieldstrength B for the dominant m = 0 component of the PADin (a). The contributions of individual continuum angular-momentum pairs (J, J ′) are shown (with Jmax = 5), wherethe indexed notation corresponds to that of Eq. (10) in [16].

arises already at ∼ 1 G and is dominated by the admix-ing of increasingly higher angular momenta in the con-tinuum. Note in the plots of Fig. 3(d-h) that if B = 0,the maximum anisotropy order is µ = 4 for our quantumnumbers.

We have shown that the chemical reaction of photodis-sociation can be strongly altered in the ultracold regimeby applied magnetic fields. In this work, the fragmen-

tation of 88Sr2 molecules was explored for a range offields from 0 to 10 G, and for a variety of energies abovethreshold in the 0–2 mK range. The near-threshold con-tinuum has a high density of partial waves that are read-ily mixed by the field, resulting in pronounced changesof the photofragment angular distributions. The the-ory of photodissociation, after explicit accounting forfield-induced mixing of angular momenta in the boundand continuum states, and combined with an accuratequantum-chemistry molecular model, has yielded excel-lent agreement with experimental data. The experimentand its clear interpretation was made possible by prepar-ing the molecules in well-defined quantum states. Wehave shown that ultracold molecule techniques allow ahigh level of control over basic chemical reactions withweak applied fields. Moreover, this work serves as a testof ab initio molecular theory in the continuum.

We acknowledge the ONR Grants No. N00014-16-1-2224 and N00014-17-1-2246, the NSF Grant No. PHY-1349725, and the NIST Grant No. 60NANB13D163 forpartial support of this work. R. M. and I. M. alsoacknowledge the Polish National Science Center GrantNo. 2016/20/W/ST4/00314 and M. M. the NSF IGERTGrant No. DGE-1069240.

SUPPLEMENTAL MATERIAL

This supplement summarizes our extension of thequantum mechanical theory of photodissociation to thesituation where the total angular momentum is not a con-served quantum number, as is the case in our ultracold-molecule experiments with applied magnetic fields.

Notation

In our theoretical description of quantum mechani-cal photodissociation, the notation follows [11, 17] withslight changes. The main symbols are as follows:

• R = (R,Θ,Φ): vector that connects the pair ofatomic fragments. The angles Θ,Φ are defined rel-ative to the molecular axis.• r: set of electronic coordinates of the atoms.• k = (k, θ, φ): scattering wave vector of the

photofragments. The angles θ, φ are defined rel-ative to the z axis, the quantum axis in the labframe.• j = j1 + j2: combined angular momentum of the

atomic fragments.• mj ≡ m: projection of j onto the lab z axis.• l: orbital angular momentum of the atomic frag-

ments about their center of mass.• ml: projection of l onto the lab z axis.• J = j+l: total angular momentum of the photodis-

sociated system.

Page 5: Control of Ultracold Photodissociation with Magnetic Fields · quantum chemistry model that includes nonadiabatic e ects and predicts strong mixing of partial waves in the photofragment

5

• M = mj +ml: projection of J onto the lab z axis.

• Ω: projection of J onto the molecular axis.

A novel aspect of this work is that the total angularmomentum J is not conserved, while M is the rigorouslyconserved quantum number. To account for this, we in-troduce the indexed angular momenta JR and Jk, wherethe subscripts R and k denote the entrance and exit chan-nels for the continuum wave function of the photofrag-ments.

Parametrization of the photofragment angulardistribution

For photodissociation of a diatomic molecule, thephotofragment angular distribution (PAD) is given bythe intensity function of the polar angle θ and azimuthalangle φ as

I(θ, φ) ∝ β0

(1 +

∞∑µ=1

µ∑ν=0

βµνPνµ (cos θ) cos(νφ)

),

(1 revisited)

where P νµ (cos θ) are the associated Legendre polynomialsand βµν are anisotropy parameters. For the parallel po-larization of light, the PADs are cylindrically symmetric(no φ dependence) [7], and we set βµ ≡ βµ0 while allother βµν vanish.

Theory of photodissociation in a magnetic field

The photodissociation process is characterized by a dif-ferential cross section σ(k) = |f |2, defined by Fermi’sgolden rule with the electric-dipole (E1) transition oper-

ator. The corresponding scattering amplitude is

f ∝ 〈Ψpf jmj

k (r,R)|T 1E1|Ψ

piJiMi

(r,R)〉, (2)

where ΨpiJiMi

(r,R) and Ψpf jmj

k (r,R) are the ini-tial (bound-staet) and final (continuum) wave func-tions. This description was first applied in the Born-Oppenheimer approximation in [8]. Furthermore, thetreatment of triatomic photodissociation [9, 18] is use-ful for our diatomic case with additional internal atomicstructure. Detailed derivation of photodissociation the-ory for individual magnetic sublevels is available in lit-erature [11, 17, 19]. However, to the best of our knowl-edge, the wave functions in the presence of a magneticfield (eigenfunctions of the Zeeman Hamiltonian) havenot been previously incorporated into the theory. Pho-todissociation in a magnetic field was discussed in [12],but J was assumed to be a good quantum number, whichis not the case in our experimental regime even for weakfields.

In this work we consider E1 photodissociation ofweakly bound ground-state 88Sr2 molecules into the0+u /1u ungerade continuum correlating to the 1S0+3P1

atomic threshold. The transition operator connecting theinitial and final wave functions is assumed to be constantand proportional to the atomic value. This approxima-tion is valid for weakly bound molecules. It is assumedthat the field affects only the excited states, since theground state (correlating to 1S0+1S0) is nearly nonmag-netic.

Bound state wave function

Since the initial (bound-state) wave function is not af-fected by the magnetic field B, it is given by the standardform using the electronic coordinates r and the inter-nuclear vector R = (R, θ, φ),

ΨpiJiMi

(r,R) =1√2

√2Ji + 1

8π2

+Ji∑Ωi=−Ji

1√1 + δΩi0

(3)

×(D

(Ji)?

MiΩi(φ, θ, 0)ψpiJiΩi

(r, R) + σiD(Ji)

?

Mi−Ωi(φ, θ, 0)ψpiJi−Ωi

(r, R)),

where the subscript i denotes the initial molecular state,σ is the spectroscopic parity defined as p(−1)J , p is theparity with respect to the space-fixed inversion, andDJ

are the Wigner rotation matrices. In Hund’s case (c) theinternal wave function ψpiJiΩi

(r, R) can be represented

by the Born-Huang expansion [20, 21],

ψpiJiΩi(r, R) =

∑ni

φniΩi(r;R)χpiniJiΩi(R), (4)

where φniΩi(r;R) are the solutions of the elec-tronic Schrodinger equation including spin-orbit cou-pling, χpiniJiΩi

(R) are the rovibrational wave functions,

Page 6: Control of Ultracold Photodissociation with Magnetic Fields · quantum chemistry model that includes nonadiabatic e ects and predicts strong mixing of partial waves in the photofragment

6

and the index n labels all relativistic electronic channelsthat are included in the model. The rovibrational wavefunctions are solutions of a system of coupled differentialequations as detailed in [13].

Continuum state wave function

The correct description of the final (continuum) wavefunction is crucial to explaining and predicting the out-come of photodissociation in a magnetic field. Zeemanmixing of rovibrational levels was responsible for observa-tions of forbidden molecular (bound-to-bound) E1 tran-sitions that violate the ∆J = 0,±1 selection rule [2].Similar effects are expected for bound-to-continuum pho-todissociation transitions.

In a magnetic field, the only conserved quantities arethe projection of the total angular momentum M and

the total parity. For dissociation to the ungerade contin-uum (1S0+3P1), the atomic angular-momentum quan-tum number is j = 1. Its magnetic sublevels mj =1, 0,−1 are split by the field, and therefore the photodis-sociation cross-section calculations have to be performedfor each sublevel individually. The 0+

u and 1u states,corresponding to Ω = 0 and |Ω| = 1, are coupled bythe nonadiabatic Coriolis interaction, while the Zeemaninteraction couples the ∆J = 0,±1 states. For thesereasons, two sets of additional numbers are introduced:ΩR, JR, nR correspond to the entrance channels of themultichannel continuum wave function and Ωk, Jk, nkcorrespond to the exit channels. In this work, selectionrules fix ΩR = 1.

As a result, the continuum wave function correspond-ing to the wave vector k is

Ψpjmj

k (r,R) =∑

JkMJR

∑ΩkΩR

∑lml

(−1)Jk+ΩkY ?lml(k)

√2l + 1

8π2

(Jk j l−M mj ml

)(Jk j lΩk −Ωk 0

)(5)

×√

2Jk + 1√1 + δΩk0

1√1 + δΩR0

(DJR

ΩRM(R)ψjJkΩkp

JRΩR(r, R) + pDJR

−ΩRM(R)ψjJkΩkp

JR−ΩR(r, R)

),

where Ylmlare the spherical harmonics. The detailed

derivation of this wave function is found in [9, 17], butfor the simpler case when J is a good quantum numberand the Jk, JR channel numbers are not needed.

The function ψjJkΩkpJRΩR

(r, R) can be expressed by theBorn-Huang expansion as

ψjJkΩkpJRΩR

(r, R) =∑nknR

φnRΩR(r, R)χjJkΩknkp

JRΩRnR(R),

(6)

where the wave functions φnRΩR(r, R) are the solutions

of the electronic Schrodinger equation. The rovibrationalwave functions χjJkΩknkp

JRΩRnR(R) are obtained by solving the

nuclear Schrodinger equation with the Hamiltonian

H = − ~2

2µR2

∂RR2 ∂

∂R+

~2l2

2µR2+ V (R), (7)

where V (R) is the potential matrix including the Coriolisand Zeeman couplings, ~ is the reduced Planck constant,and µ is the reduced atomic mass.

At large interatomic distances in the presence of ex-ternal fields, the asymptotic value of V (R), or Vas,is not diagonal in the basis of the wave functionsφnRΩR

(r, R). It is then necessary to introduce a trans-formation C that diagonalises Vas. The rovibrationalfunctions χjJkΩknkp

JRΩRnR(R) form a matrix X that is prop-

agated to large distances and transformed to the basis

that diagonalizes the asymptotic potential, X = CT XC.Then the boundary conditions are imposed [22, 23] as

X(R)→ J(R) +N(R) ·K, (8)

where K is the reaction matrix, J(R) and N(R) are thediagonal matrices containing the spherical Bessel func-tions for the open channels,

[J(R)]ij = δij1√kjjl(kjR), [N(R)]ij = δij

1√kjnl(kjR),

(9)

kj is the wave number of the jth channel, and l is theorbital angular momentum of the jth channel. A moredetailed description of the close-coupled equations in amagnetic field can be found in [24].

Anisotropy parameters

After inserting the wave functions (3) and (5) intoFermi’s golden rule (2) and transforming the cross sec-tion for the photodissociation process using the Clebsch-Gordan series and properties of the Wigner 3j symbols,

Page 7: Control of Ultracold Photodissociation with Magnetic Fields · quantum chemistry model that includes nonadiabatic e ects and predicts strong mixing of partial waves in the photofragment

7

we get the following expansion for the PAD:

I(θ, φ) ∝ β0

(1 +

∞∑µ=1

µ∑ν=0

βµνPνµ (cos θ) cos(νφ)

),

(1 revisited)

where the anisotropy parameters are given by

βµν =1

β0

∑JkJRJ′

kJ′R

∑ll′MM ′

tJkJRtJ′k?

J′RUJklJRM

UJ′kl

J′RM

′ [µ]

√(µ− ν)!

(µ+ ν)!(2− δMM ′)

(l l′ µ

M −mj mj −M ′ ν

)(l l′ µ0 0 0

)(10)

and [A] ≡ 2A + 1. The symbols UJklJRMin Eq. (10) are defined as

UJklJRM=∑

PΩkml

(−1)Jk+Ωk−mj [l]

√[Jk]√

1 + δΩk0

(Jk j l−M mj ml

)(Jk j lΩk −Ωk 0

)(JR 1 Ji−M P Mi

), (11)

and the symbols tJkJR are the scaled matrix elements of the asymptotic body-fixed E1 transition operator dBF withthe initial and final rovibrational wave functions,

tJkJR =1

2√

2

Ji∑Ωi=−Ji

J∑ΩR=−J

1∑q=−1

∑nknR

(−1)M−ΩR

√[Ji]√

1 + δΩi0

√[JR]√

1 + δΩR0

(12)

×(JR 1 Ji−ΩR q Ωi

)〈χjJkΩknkpJRΩRnR

(R)|dBF|χpiniJiΩi(R)〉.

The normalization factor β0 is given by

β0 =∑lM

∣∣∣∣∣∑JkJR

tJkJRUJklJRM

√2l + 1(−1)Jk

∣∣∣∣∣2

. (13)

In Eq. (11), the polarization index P = 0 if the pho-todissociation light is polarized along the z axis, whileP = −1, 1 if the light is polarized perpendicularly to thez axis.

The properties of the 3j symbols force the followingrules for the µ, ν indices:

• µ is even for homonuclear dimers.• µmax = 2Jk,max + 2j = 2Jk,max + 2 for resolvedm sublevels. Thus the number of terms in the ex-pansion (1) is limited by the number of channelsused to construct the continuum wave function.

(When the m sublevels are degenerate and are ob-served simultaneously, additional symmetry leadsto µmax = 2Jk,max.) If B = 0, then Jk,max = 1and µmax = 4, as can be seen in Fig. 3(d-h) of themanuscript.

• ν = M ′−M . Since M = M ′ = Mi for parallel lightpolarization, ν = 0 and thus the photodissociationcross section is cylindrically symmetric.

The anisotropy parameters presented in Fig. 3(d-h) ofthe manuscript are calculated using Eq. (10), but insteadof summing over Jk, J

′k, the contributions of each combi-

nation Jk, J′k are individually plotted, and subsequently

divided by β0 from Eq. (13).

Page 8: Control of Ultracold Photodissociation with Magnetic Fields · quantum chemistry model that includes nonadiabatic e ects and predicts strong mixing of partial waves in the photofragment

8

∗ Present address: Department of Physics, University ofChicago, 929 East 57th Street GCIS ESB11, Chicago, IL60637, USA

† Present address: Facebook, Inc., 1 Hacker Way, MenloPark, CA 94025, USA

[email protected][1] N. Balakrishnan, “Perspective: Ultracold molecules and

the dawn of cold controlled chemistry,” J. Chem. Phys.145, 150901 (2016).

[2] B. H. McGuyer, M. McDonald, G. Z. Iwata, W. Sko-morowski, R. Moszynski, and T. Zelevinsky, “Control ofoptical transitions with magnetic fields in weakly boundmolecules,” Phys. Rev. Lett. 115, 053001 (2015).

[3] K.-K. Ni, S. Ospelkaus, D. Wang, G. Quemener,B. Neyenhuis, M. H. G. de Miranda, J. L. Bohn, J. Ye,and D. S. Jin, “Dipolar collisions of polar molecules inthe quantum regime,” Nature 464, 1324–1328 (2010).

[4] M. H. G. de Miranda, A. Chotia, B. Neyenhuis, D. Wang,G. Quemener, S. Ospelkaus, J. L. Bohn, J. Ye, andD. S. Jin, “Controlling the quantum stereodynamics ofultracold bimolecular reactions,” Nature Phys. 7, 502–507 (2011).

[5] G. Quemener and J. L. Bohn, “Ultracold molecular col-lisions in combined electric and magnetic fields,” Phys.Rev. A 88, 012706 (2013).

[6] K. M. Jones, E. Tiesinga, P. D. Lett, and P. S. Juli-enne, “Ultracold photoassociation spectroscopy: Long-range molecules and atomic scattering,” Rev. Mod. Phys.78, 483–535 (2006).

[7] M. McDonald, B. H. McGuyer, F. Apfelbeck, C.-H.Lee, I. Majewska, R. Moszynski, and T. Zelevin-sky, “Photodissociation of ultracold diatomic strontiummolecules with quantum state control,” Nature 534, 122–126 (2016).

[8] R. N. Zare, “Photoejection dynamics,” Mol. Photochem.4, 1–37 (1972).

[9] G. G. Balint-Kurti and M. Shapiro, “Photofragmentationof triatomic molecules. Theory of angular and state dis-tribution of product fragments,” Chem. Phys. 61, 137–155 (1981).

[10] S. E. Choi and R. B. Bernstein, “Theory of orientedsymmetric-top molecule beams: Precession, degree of ori-entation, and photofragmentation of rotationally state-selected molecules,” J. Chem. Phys. 85, 150–161 (1986).

[11] V. V. Kuznetsov and O. S. Vasyutinskii, “Photofragmentangular momentum distribution beyond the axial recoilapproximation: The role of molecular axis rotation,” J.Chem. Phys. 123, 034307 (2005).

[12] J. A. Beswick, “Angular distribution of photo-predissociation fragments in the presence of a mangeticfield,” Chem. Phys. 42, 191–199 (1979).

[13] W. Skomorowski, F. Paw lowski, C. P. Koch, andR. Moszynski, “Rovibrational dynamics of the strontiummolecule in the A1Σ+

u , c3Πu, and a3Σ+u manifold from

state-of-the-art ab initio calculations,” J. Chem. Phys.136, 194306 (2012).

[14] M. Borkowski, P. Morzynski, R. Ciury lo, P. S. Julienne,M. Yan, B. J. DeSalvo, and T. C. Killian, “Mass scal-ing and nonadiabatic effects in photoassociation spec-troscopy of ultracold strontium atoms,” Phys. Rev. A90, 032713 (2014).

[15] G. Reinaudi, C. B. Osborn, M. McDonald, S. Ko-tochigova, and T. Zelevinsky, “Optical production ofstable ultracold 88Sr2 molecules,” Phys. Rev. Lett. 109,115303 (2012).

[16] See Supplemental Material for the quantum mechani-cal description of photodissociation in applied magneticfields.

[17] P. S. Shternin and O. S. Vasyutinskii, “The parity-adapted basis set in the formulation of the photofrag-ment angular momentum polarization problem: The roleof the Coriolis interaction,” J. Chem. Phys. 128, 194314(2008).

[18] G. G. Balint-Kurti and M. Shapiro, Quantum Theoryof Molecular Photodissociation in Advances in ChemicalPhysics: Photodissociation and Photoionization, vol. 60,ed. K. P. Lawley (John Wiley & Sons, Inc., Hoboken, NJ,1985).

[19] J. G. Underwood and I. Powis, “Photodissociation of po-larized diatomic molecules in the axial recoil limit: Con-trol of atomic polarization,” J. Chem. Phys. 113, 7119–7130 (2000).

[20] M. Born and K. Huang, Dynamical Theory of CrystalLattices (Oxford University Press, Oxford, 1956).

[21] B. Bussery-Honvault, J.-M. Launay, T. Korona, andR. Moszynski, “Theoretical spectroscopy of the calciumdimer in the A1Σ+

u , c3Πu, and a3Σ+u manifolds: An

ab initio nonadiabatic treatment,” J. Chem. Phys. 125,114315 (2006).

[22] R. D. Levine, Quantum Mechanics of Molecular RateProcesses (Oxford University Press, Oxford, 1969).

[23] B. R. Johnson, “The multichannel log-derivative methodfor scattering calculations,” J. Comput. Phys. 13, 445–449 (1973).

[24] R. V. Krems and A. Dalgarno, “Quantum-mechanicaltheory of atom-molecule and molecular collisions in amagnetic field: Spin depolarization,” J. Chem. Phys.120, 2296–2307 (2004).