Quantum dynamical calculations of ultracold collisions induced by nonlinearly chirped light

9
PHYSICAL REVIEW A 85, 013424 (2012) Quantum dynamical calculations of ultracold collisions induced by nonlinearly chirped light J. L. Carini, 1 J. A. Pechkis, 1,2 C. E. Rogers III, 1 P. L. Gould, 1 S. Kallush, 3 and R. Kosloff 4 1 Department of Physics, University of Connecticut, Storrs, Connecticut 06269, USA 2 Currently with the Naval Research Laboratory, 4555 Overlook Avenue S.W., Washington, DC 20375, USA 3 Department of Physics and Optical Engineering, ORT Braude, P.O. Box 78, Karmiel, Israel 4 Department of Physical Chemistry and the Fritz Haber Research Center for Molecular Dynamics, The Hebrew University, 91094 Jerusalem, Israel (Received 27 May 2011; published 31 January 2012) We describe quantum dynamical calculations of ultracold 85 Rb trap-loss collisions induced by pulses of light whose frequency is chirped on the nanosecond time scale. The chirped light excites the ground-state collisional wave function to the long-range attractive potential and escape from the trap is modeled by an absorbing boundary at short range. Both positive and negative chirps are considered and various chirp shapes and detunings are examined. For positive chirps, the loss rates are rather independent of the chirp shape. Negative chirps, on the other hand, show a dependence on chirp shape for detunings where collisional flux can be coherently returned to the ground state. These trends are consistent with the results of a recent experiment. DOI: 10.1103/PhysRevA.85.013424 PACS number(s): 32.80.Qk, 34.50.Rk I. INTRODUCTION Efforts to bring the techniques of coherent control to bear on ultracold molecule formation [1] have attracted significant attention in recent years. This has been motivated in large part by the interest in improving the production efficiency and manipulation of ultracold molecules [25] for the myriad of applications in ultracold chemistry, many-body dipolar systems, quantum information, and precision spectroscopy. Direct laser cooling of atoms is now a standard technique [6]. Although progress is being made [7], extending this concept to molecules is, in general, not straightforward. The impediment is the multitude of internal levels [8]. Since coherent control techniques [9,10] typically deal with internal degrees of freedom, while atom cooling schemes [6] manipulate the translational degrees of freedom, combining the two offers the prospect of complete control over molecular systems. The process of photoassociation [1114], in which laser light converts colliding atoms into bound molecules, is one specific possibility. A number of such schemes [1529], using coherent control with ultrafast pulses to form molecules, have been considered. Experimental progress, however, has been limited. Early experiments demonstrated that the photodestruction of ultracold molecules could be coherently controlled using shaped [30] and chirped [31] ultrafast pulses. More recently, coherent transients were observed in excited-state molecules photoassociated with femtosecond pulses [3236]. In addition, some evidence for the production of molecules in the electronic ground state was reported [33]. Photoassociative ionization studies with short pulses have also been carried out [37,38]. We have recently performed a series of experiments with frequency-chirped light [3942] that are related to coher- ently controlled molecule formation, but that differ in two important aspects. Instead of detecting molecules formed by photoassociation, we measure the rate of laser-induced trap-loss collisions between ultracold Rb atoms. Also, in contrast to the short-pulse work, our experiments take place on a nanosecond time scale, which is better matched to the motion of atoms undergoing long-range collisions. We have demonstrated coherent control of these collisions by varying the direction of the frequency chirp [41]. Because the atom pair always accelerates inward on the attractive potential, and a negative chirp provides an excitation (or Condon) radius which can follow this motion, there can be multiple interactions between the atom pair and the light, as shown in Fig. 1(b). This can lead to coherent deexcitation and thus the suppression of the short-range inelastic collision rate, a process we denote as “coherent collision blocking.” This cannot happen for a positive chirp where the excitation radius moves outward with time [Fig. 1(a)]. In our most recent work [42], we demonstrate further coherent control by varying not only the direction, but also the shape of the frequency chirp. We find that under certain conditions, the details of the chirp nonlinearity significantly affect the rate of trap-loss collisions. This type of control of the collisional wave function as it evolves from long range to short range may benefit the process of photoassociative molecule formation. Dynamical studies of trap-loss collisions may also help to improve our understanding of the mechanisms of these important loss processes [43]. In the present work, we discuss the results of quantum dynamical calculations of these most recent experiments. The dynamics of the collisional wave functions, on both ground-state and excited-state potentials, are followed in the presence of various frequency chirps. Collisional loss, as measured in the experiment, is assumed to occur when the atom pair reaches short range in the excited state. In general, the simulations reproduce the trends observed in the experimental data. A dependence of the collisional loss rate on the shape of the negative chirp is demonstrated for certain parameters, while the loss rate for positive chirps is found to be relatively insensitive to the details of the chirp shape. These variations with negative chirp shape occur in the regime of coherent collision blocking, indicating that details of the chirp are most important when the internuclear separation and the Condon radius vary on similar time scales. This concurrence allows multiple interactions between the collisional wave function and the chirped light, resulting in an increased sensitivity of the excited-state wave function evolution to the chirp shape. We also find that the dependence on negative chirp shape 013424-1 1050-2947/2012/85(1)/013424(9) ©2012 American Physical Society

Transcript of Quantum dynamical calculations of ultracold collisions induced by nonlinearly chirped light

PHYSICAL REVIEW A 85, 013424 (2012)

Quantum dynamical calculations of ultracold collisions induced by nonlinearly chirped light

J. L. Carini,1 J. A. Pechkis,1,2 C. E. Rogers III,1 P. L. Gould,1 S. Kallush,3 and R. Kosloff4

1Department of Physics, University of Connecticut, Storrs, Connecticut 06269, USA2Currently with the Naval Research Laboratory, 4555 Overlook Avenue S.W., Washington, DC 20375, USA

3Department of Physics and Optical Engineering, ORT Braude, P.O. Box 78, Karmiel, Israel4Department of Physical Chemistry and the Fritz Haber Research Center for Molecular Dynamics,

The Hebrew University, 91094 Jerusalem, Israel(Received 27 May 2011; published 31 January 2012)

We describe quantum dynamical calculations of ultracold 85Rb trap-loss collisions induced by pulses of lightwhose frequency is chirped on the nanosecond time scale. The chirped light excites the ground-state collisionalwave function to the long-range attractive potential and escape from the trap is modeled by an absorbingboundary at short range. Both positive and negative chirps are considered and various chirp shapes and detuningsare examined. For positive chirps, the loss rates are rather independent of the chirp shape. Negative chirps, on theother hand, show a dependence on chirp shape for detunings where collisional flux can be coherently returned tothe ground state. These trends are consistent with the results of a recent experiment.

DOI: 10.1103/PhysRevA.85.013424 PACS number(s): 32.80.Qk, 34.50.Rk

I. INTRODUCTION

Efforts to bring the techniques of coherent control to bearon ultracold molecule formation [1] have attracted significantattention in recent years. This has been motivated in largepart by the interest in improving the production efficiencyand manipulation of ultracold molecules [2–5] for the myriadof applications in ultracold chemistry, many-body dipolarsystems, quantum information, and precision spectroscopy.Direct laser cooling of atoms is now a standard technique [6].Although progress is being made [7], extending this concept tomolecules is, in general, not straightforward. The impedimentis the multitude of internal levels [8]. Since coherent controltechniques [9,10] typically deal with internal degrees offreedom, while atom cooling schemes [6] manipulate thetranslational degrees of freedom, combining the two offersthe prospect of complete control over molecular systems.The process of photoassociation [11–14], in which laser lightconverts colliding atoms into bound molecules, is one specificpossibility. A number of such schemes [15–29], using coherentcontrol with ultrafast pulses to form molecules, have beenconsidered. Experimental progress, however, has been limited.Early experiments demonstrated that the photodestructionof ultracold molecules could be coherently controlled usingshaped [30] and chirped [31] ultrafast pulses. More recently,coherent transients were observed in excited-state moleculesphotoassociated with femtosecond pulses [32–36]. In addition,some evidence for the production of molecules in the electronicground state was reported [33]. Photoassociative ionizationstudies with short pulses have also been carried out [37,38].

We have recently performed a series of experiments withfrequency-chirped light [39–42] that are related to coher-ently controlled molecule formation, but that differ in twoimportant aspects. Instead of detecting molecules formedby photoassociation, we measure the rate of laser-inducedtrap-loss collisions between ultracold Rb atoms. Also, incontrast to the short-pulse work, our experiments take placeon a nanosecond time scale, which is better matched to themotion of atoms undergoing long-range collisions. We havedemonstrated coherent control of these collisions by varying

the direction of the frequency chirp [41]. Because the atompair always accelerates inward on the attractive potential, and anegative chirp provides an excitation (or Condon) radius whichcan follow this motion, there can be multiple interactionsbetween the atom pair and the light, as shown in Fig. 1(b). Thiscan lead to coherent deexcitation and thus the suppression ofthe short-range inelastic collision rate, a process we denoteas “coherent collision blocking.” This cannot happen for apositive chirp where the excitation radius moves outward withtime [Fig. 1(a)]. In our most recent work [42], we demonstratefurther coherent control by varying not only the direction, butalso the shape of the frequency chirp. We find that under certainconditions, the details of the chirp nonlinearity significantlyaffect the rate of trap-loss collisions. This type of control ofthe collisional wave function as it evolves from long rangeto short range may benefit the process of photoassociativemolecule formation. Dynamical studies of trap-loss collisionsmay also help to improve our understanding of the mechanismsof these important loss processes [43].

In the present work, we discuss the results of quantumdynamical calculations of these most recent experiments.The dynamics of the collisional wave functions, on bothground-state and excited-state potentials, are followed in thepresence of various frequency chirps. Collisional loss, asmeasured in the experiment, is assumed to occur when the atompair reaches short range in the excited state. In general, thesimulations reproduce the trends observed in the experimentaldata. A dependence of the collisional loss rate on the shapeof the negative chirp is demonstrated for certain parameters,while the loss rate for positive chirps is found to be relativelyinsensitive to the details of the chirp shape. These variationswith negative chirp shape occur in the regime of coherentcollision blocking, indicating that details of the chirp are mostimportant when the internuclear separation and the Condonradius vary on similar time scales. This concurrence allowsmultiple interactions between the collisional wave functionand the chirped light, resulting in an increased sensitivity ofthe excited-state wave function evolution to the chirp shape.We also find that the dependence on negative chirp shape

013424-11050-2947/2012/85(1)/013424(9) ©2012 American Physical Society

J. L. CARINI et al. PHYSICAL REVIEW A 85, 013424 (2012)

1 10 100 1000 10000 100000-5000

0

5000

10000

15000

20000

25000

5S1/2

+5P3/2

E(c

m-1)

R (units of a0)

5S1/2

+5S1/2

1 10 100 1000 10000 100000-5000

0

5000

10000

15000

20000

25000

5S1/2

+5P3/2

E(c

m-1)

R (units of a0)

5S1/2

+5S1/2

Positive Chirp Negative Chirp(a) (b)

FIG. 1. Ultracold collisions in Rb induced by frequency-chirped light. The ground- and excited-state potentials are shown as well as theexcited-state wave packet, the region of R swept over by the chirp, and the absorbing boundary (dashed line). Note that the horizontal axis islogarithmic. For the positive chirp (a), the excitation radius (upward arrow) moves outward in time while the excited-state wave packet movesinward. However, for the negative chirp (b), the excitation radius moves inward in time, following the excited wave packet trajectory. Thisleads to multiple interactions, which can return a portion of the wave packet to the ground state (downward arrow).

involves the efficiency of collisional loss for a given amountof excitation, and not simply the amount of excitation.

The paper is organized as follows. In Sec. II, we presentthe theory used in the quantum dynamical calculations of theultracold collisions. The specific nonlinear frequency chirpsused in the simulations are also described. In Sec. III, wepresent results of the simulations, including not only the overallcollisional loss rate constants, but also examples of populationand wave function dynamics. The results are also comparedto the experimental measurements. Section IV comprisesconcluding remarks.

II. THEORY

The time-dependent Schrodinger equation for two inter-acting 85Rb atoms is solved with the Chebychev polynomialexpansion method [44] for the temporal operators, and with amapped Fourier grid to address the wide range of length scalesin the system [45]. In the dressed-state picture, the two-channelHamiltonian has the form

H =(

T + Vg(R) h�(t)h�∗(t) T + Ve(R) + h�p

). (1)

In Eq. (1), T is the kinetic energy operator, �p is thecenter detuning of the linearly chirped pulse with respect tothe atomic asymptote [defined as 5S1/2 (F = 3) → 5P3/2

(F′ = 4)], and Vg/e are the ground/excited-state potentials.

We restrict the treatment to a single excited-state potential,0+

u [18,46], and ignore hyperfine structure. To account forthe two possible initial scattering states the computation isperformed separately for the singlet X 1�+

g and the tripleta 3�+

u states [18,46], each adjusted for its proper scatteringlength [47], and then a weighted average is taken. At longrange, hyperfine mixing relaxes the u-g selection rule, allowingboth of these ground states to be excited. Within the rotating-wave approximation, the time-dependent coupling between theground and excited states due to the chirped pulse is given by

h� = μE0 exp

[−

(t − tcenter

)2

+ iνchirp(t − tcenter)2

], (2)

where μ is the transition dipole moment, E0 is the peakelectric field, σ is the pulse width, νchirp is the chirp rate,and tcenter defines the center of the intensity pulse. Dueto the nature of the nonlinear chirps, instead of using achirp rate, we input instantaneous frequency offsets, ω(t),from the center frequency of the linearly chirped pulse, �p.Therefore,

h� = μE0 exp

[−

(t − tcenter

)2

+ iω(t)(t − tcenter)

]. (3)

The chirps (concave-down, concave-up, and linear for bothpositive and negative directions, chirping ∼1 GHz in 100 ns)are characterized by a heterodyne signal [42]. A referencebeam is combined with the modulated beam on a photodiode.The resulting beat frequency yields the instantaneous offsetfrequency as a function of time, ω(t), through analysisof the maxima and minima. These data points are theninterpolated using a piecewise cubic Hermite interpolatingpolynomial (PCHIP) to yield a frequency value every 0.1 ns.The resulting values are smoothed using a 5 ns full widthat half maximum (FWHM) Gaussian function, ensuring thatthe slope is always <0 for negative chirps or >0 for positivechirps. The smoothed frequencies are then interpolated againusing a PCHIP time step of ∼0.45 ps, corresponding to thetemporal propagation step of the simulations. The resultinginstantaneous frequencies, in 2 ns steps, are shown in Fig. 2.The corresponding 40 ns FWHM intensity profiles of thepulses with peak intensity I = 67 W/cm2 are also displayed.We run the simulations for 200 ns to ensure full convergence,keeping the frequencies constant at the value they reachedat 100 ns for the duration of the run. In Sec. III we presentresults for the first 100 ns since this more closely matchesthe time scale of the experiment [42] and convergence isalready met to within a few percent at 100 ns. Based on therange of the frequency chirps, any positive (negative) chirpwith �p closer to the atomic resonance than ∼−635 (−500)MHz would go through resonance during the 100 ns chirp.The atomic resonance is problematic because it correspondsto an infinite excitation radius. Also, we have not includeda repulsive excited-state potential that would be relevant for

013424-2

QUANTUM DYNAMICAL CALCULATIONS OF ULTRACOLD . . . PHYSICAL REVIEW A 85, 013424 (2012)

FIG. 2. (Color online) Shaped positive (a) and negative (b)frequency chirps. The curves shown are smoothed interpolations ofthe actual heterodyne signals. Indicated frequencies ω(t) are relativeto the centers of the linear chirps. Also shown are the corresponding40 ns FWHM Gaussian intensity pulses I (t).

positive detunings. For our set of �p values, passage throughthe atomic resonance only happens for the positive chirp with�p/2π = −550 MHz. Even here, this occurs near the endof the 100 ns period, where the pulse intensity is greatlyreduced.

The initial s-wave scattering state is computed by diago-nalization of the time-independent Hamiltonian of the ground-state potential in the mapped Fourier grid basis. Observablesare computed by taking a single initial state that correspondsto the 50 μK temperature of the sample and factorizingaccording to [20]. This thermal averaging is discussed inSec. III. Contributions to the loss from higher partial wavesare found to be negligible.

We model the trap loss as follows. If the atom pair reachesshort range in the excited state (5S1/2 + 5P3/2), the atomscan gain sufficient kinetic energy to be ejected from thetrap. This can occur either by radiative escape, where theexcited atom pair spontaneously decays at short range, orby fine-structure predissociation, where a short-range curvecrossing results in the atom pair emerging on the lowerfine-structure asymptote (5S1/2 + 5P1/2) [48]. Decay into abound molecular state, which would occur primarily inside theabsorbing boundary, also results in loss of the colliding atomsbecause molecules are not confined in the atom trap. We modelthese loss mechanisms by placing an absorbing boundary onthe excited-state potential at R = 100 a0 to absorb incomingcollisional flux. Any collisional flux reaching this internuclearseparation is considered to result in the loss of the atoms fromthe trap. We have checked that the calculated losses are ratherinsensitive to the exact location of this boundary, as discussedin Sec. III. We note that this short-range absorbing boundaryis distinct from an absorbing wall of a calculational box usedto prevent population from leaving the box.

In the collisions we are simulating, the atoms initially movevery slowly and interact at long range. The relevant time scalecan therefore be quite long and spontaneous emission cannotbe ignored. We model this decay process by coupling the

excited state to a sink channel with a damping rate of =(22 ns)−1 [48]. This sink channel represents loss of excited-state population occurring outside the absorbing boundary andthus not resulting in loss from the trap. We note that thistreatment takes into account only a unidirectional decay fromthe excited state, and does not account for any incoherentinteractions that result from decay of this population back tothe ground state. Therefore, some incoherent effects that arerelevant to our discussion, such as flux enhancement (see nextsection), are not included.

III. RESULTS AND ANALYSIS

To understand the results of the quantum dynamical calcu-lations, we first examine the population transfers and lossesshown in Fig. 3. As discussed in Sec. II, the trap loss is definedas the cumulative fraction of the initial ground-state populationwhich encounters the excited-state absorbing boundary. Ascan be seen from these plots, most of the population remainsin the ground state throughout the 100 ns chirp. At longtimes, the total population (ground state plus excited state)does not sum to unity because of the irreversible transfer tothe trap-loss channel and to the spontaneous emission sink.The excited-state maxima for the positive chirps [Figs. 3(a)and 3(b)] come slightly after the peak of the intensity pulse at60 ns whereas the excited-state maxima for the negative chirps[Figs. 3(c) and 3(d)] come slightly before the peak of theintensity pulse at 40 ns. This time ordering was also observedin simulations of chirped photoassociation on much fastertime scales [16]. The population plots from the positive chirps[Figs. 3(a) and 3(b)] are smooth with some slight variationsfor the various shapes whereas the negative chirps [Figs. 3(c)and 3(d)] show rapid small-amplitude oscillations. These areout of phase for the ground and excited states and are likelyresidual Rabi oscillations resulting from extended interactionsas the wave packet and excitation radius both move inward.Although we have only shown population plots for chirpswith �p/2π = −750 MHz, detunings of −550 MHz and−950 MHz yield similar results. The losses for the singletpotential [Figs. 3(a) and 3(c)] considerably exceed those forthe triplet potential [Figs. 3(b) and 3(d)] because the largersinglet scattering length yields more wave function amplitudeat the relevant values of R. The positive chirps all yield similarlosses for each potential at a given detuning [Figs. 3(a) and3(d)], whereas the losses show more variation with shapefor the negative chirps, especially at �p/2π = −750 MHz[Figs. 3(c) and 3(d)].

To gain further insight, we examine the loss per excitedatom pair for each case. This is defined as the trap lossdivided by the average excited-state fraction during the100 ns observation window and is a measure of the likelihoodfor an excited atom pair to undergo trap loss as opposed tospontaneous emission. The weighted average of the singlet andtriplet loss per excited atom pair for the positive and negativeconcave-down, concave-up, and linear chirps is shown inFig. 4. The amount of loss per excited atom pair decreases forboth the positive and negative chirps as the pulse detuning,�p, gets closer to resonance. This is due to the fact thatexcitation at larger R results in less acceleration and therefore agreater chance for spontaneous emission before encountering

013424-3

J. L. CARINI et al. PHYSICAL REVIEW A 85, 013424 (2012)

FIG. 3. (Color online) Ground- and excited-state population, spontaneous emission sink population, and trap loss for positive chirps [singlet(a) and triplet (b) potentials] and negative chirps [singlet (c) and triplet (d) potentials] with �p/2π = −750 MHz.

the absorbing boundary. The largest separation occurs forthe negative chirps with �p/2π = −750 MHz. Here theconcave-down chirp has the most loss per excited atom pair,followed by the linear chirp and then the concave-up chirp.This trend is consistent with that of the trap loss [Figs. 3(c)and 3(d)], indicating that the efficiency of loss for a givenamount of excitation, and not simply the amount of excitation,is important. We find that for the positive chirps there is nosignificant difference among the shapes for each detuning.The variation that occurs at �p/2π = −550 MHz for thepositive chirps is due to these shapes going through the atomicresonance (R = ∞) at slightly different times and with slightly

different slopes, as expected for the frequency chirps shown inFig. 2(a).

Examples of the excited-state wave packet evolution canbe seen in Fig. 5. We show snapshots of | e |2R2 for R <

1000 a0 for positive chirps [Figs. 5(a)–5(c)] and negative chirps[Figs. 5(e)–5(g)], starting on the singlet potential. The resultsfor the triplet potential are similar and therefore not shown.We focus first on the overall difference between the positiveand negative chirps. For the positive chirp with �p/2π =−750 MHz [Figs. 5(a)–5(c)], the excitation of the wave packetmoves outward as time progresses. For the corresponding neg-ative chirp, however, the excitation of the wave packet moves

013424-4

QUANTUM DYNAMICAL CALCULATIONS OF ULTRACOLD . . . PHYSICAL REVIEW A 85, 013424 (2012)

FIG. 4. (Color online) Weighted average of singlet and tripletloss per excited atom pair, defined in the text. This quantity is shownfor the positive and negative concave-down, concave-up, and linearchirps. The lines connecting the points are to guide the eye.

inward as the chirp proceeds, which is the same direction asthe motion caused by the attractive nature of the excited-statepotential. This difference between the two chirp directions isclearly seen by comparing Figs. 5(d) and 5(h), which show〈Re〉, the excited-state expectation value of R, as a function oftime for the positive and negative chirps, respectively. Thesetrends are similar for linear chirps and for the other detunings,neither of which, for reasons of clarity, are shown in Fig. 5.Note that toward the end of the positive chirp [Fig. 5(d)], theintensity is low and excitation is no longer occurring, so thealready existing excited-state wave function begins to moveinward on the attractive potential. Note also the significantdifference between Figs. 5(c) and 5(g). Both are at timesapproximately 20 ns after the peak of the pulse, but Fig. 5(g)has a much smaller amplitude. This is due to coherent collisionblocking whereby the negative chirp drives population backdown to the ground state as the excitation radius follows thewave packet motion inward, and is consistent with the gener-ally smaller trap loss seen for negative versus positive chirps.

We now examine the dependence on chirp shape in Fig. 5.For the positive chirp, there is not much difference in theevolution of 〈Re〉 between concave-down and concave-upshapes [Fig. 5(d)]. The results are also similar for the linearchirp (not shown). However, for the negative chirp [Fig. 5(h)],the differences in 〈Re〉 for the two chirp shapes are morepronounced. Comparing the wave packets at the beginning ofthe shaping, we see that for the negative chirp with �p/2π =−750 MHz [Fig. 5(e)] the concave-up chirp provides moreoverall excitation. This is likely due to the wave packetmotion being less synchronized with the higher instantaneouschirp rate, resulting in less deexcitation [see also Fig. 3(c)].Despite this, a larger portion of the wave packet created by theconcave-down chirp is able to more quickly accelerate inwardon the excited-state potential than for the other shapes. At thesame time, for �p/2π = −950 MHz [shown in the insert ofFig. 5(e)], there is less of a difference between concave-downand concave-up wave packets.

In order to estimate the trap-loss rate constant β from asingle energy eigenstate calculation, we start by finding thenumber of molecules per chirped pulse leading to loss Nmol.Since the gas is dilute and the Hamiltonian is represented ona finite grid in coordinate space, we divide the trap volume Vinto many small boxes of volume νbox where the number ofsmall boxes is larger than the total number of atoms in the trap,N . Then the number of excited-state molecules leading to lossat a given temperature T (taken to be 50 μK) is given by

Nmol = 1

2N2 νbox

V〈Pe〉T ,box (4)

[20]. Following the procedure in Sec. 5 of Ref. [20], thethermally averaged probability to make such a molecule,〈Pe〉T ,box, is given by

〈Pe〉T ,box = 2π2h3

μ3/2νbox

PE0

�E0

√kT

, (5)

where E0 is the single energy eigenstate, �E0 is the energywidth of this level (the inverse of the density of states in thebox), μ is the reduced mass, and PE0 is the probability toinduce an excited-atom loss event, examples of which areshown as losses in Fig. 3. We note that Eq. (5) applies whenthe Wigner-Kramers-Brillouin (WKB) approximation is valid,which is the case in our energy range [20]. This corresponds tothe trap loss coming primarily from atom pairs whose energyE0 is close to kT . The number of atoms lost per pulse, Natoms,is twice the number of corresponding molecules Nmol:

Natoms = 2π2h3N2

μ3/2V

PE0

�E0

√kT

. (6)

The collisonal trap-loss rate constant β, where β × N/V isthe time-averaged loss rate per atom, can then be defined as

β = NatomsPcycle

t

V

N2, (7)

where Pcycle is the number of pulses per cycle of theexperiment (typically 60) and t is the repetition time, 722 μs.Plugging in the relevant numbers for this experiment [42]leads to

β = PE0 × 1.04 × 10−8 cm3

s. (8)

Using this conversion [Eq. (8)], and performing the weightedaverage of the singlet and triplet ground states, we find theabsolute values for β as shown in Figs. 6(a) and 6(b). We notethat the efficiency of the absorbing boundary, which we assumeto be 100%, will affect the absolute values of β. However,the results are rather insensitive to the boundary’s location.For example, shifting the absorbing boundary from 100 a0 to150 a0 increased the singlet-state trap loss for the �p/2π =−750 MHz positive linear chirp [Fig. 3(a)] by only 0.66%.This is because the attractive potential is sufficiently steep thatany excited-state collisional flux reaching these small valuesof R is accelerated very quickly to even smaller values of R.

Focusing first on the results for the nonlinear positive chirps[Fig. 6(a)], we see that the values of β for the various chirpshapes are similar at a given �p, showing that there is littledependence of β on the details of the nonlinearity of thepositive chirp. These results are consistent with the excitation

013424-5

J. L. CARINI et al. PHYSICAL REVIEW A 85, 013424 (2012)

FIG. 5. (Color online) Evolution of | e |2R2 for concave-down and concave-up positive [(a)–(c)] and negative [(e)–(g)] chirps. For allframes, the singlet potential is assumed and the chirp has �p/2π = −750 MHz. Successive frames show various times during the chirp, asindicated. The insert in (e) shows �p/2π = −950 MHz at the same time as the main frame (19.05 ns). The center of the 40 ns FWHM intensitypulse occurs at 60 ns (b) for the positive chirps and at 40 ns (f) for the negative chirps (see Fig. 2). Also shown, in (d) and (h), are the temporalevolutions of the excited-state expectation value of R.

radius of the positive chirp sweeping outward, away from theexcited-state wave packet trajectory, as seen in Figs. 5(a)–5(c).These results are also consistent with the smooth nature ofthe population plots in Figs. 3(a) and 3(b). There is also ageneral trend of β decreasing as the detuning gets farther from

resonance. This is due to fewer atom pairs being available forexcitation at the corresponding smaller values of R.

Now focusing on the results for the nonlinear negativechirps [Fig. 6(b)], we see that the values for the negativechirps are smaller than those of the positive chirps. As

013424-6

QUANTUM DYNAMICAL CALCULATIONS OF ULTRACOLD . . . PHYSICAL REVIEW A 85, 013424 (2012)

-950 -750 -5500

2

4

6

8

10

12

-950 -750 -550

Concave-downConcave-upLinear

(10-1

2 cm3 /s

)

p/(2 ) (MHz)

Positive(d)(c)

Concave-downConcave-upLinear

p/(2 ) (MHz)

-950 -750 -5500

2

4

6

8

10

12

14

16

18

-950 -750 -550

(a)

Concave-downConcave-upLinear

(10-1

2 cm3 /s

)

p/(2 ) (MHz)

Positive(b)

Concave-downConcave-upLinear

p/(2 ) (MHz)

Negative

Negative

FIG. 6. (Color online) Quantum dynamical calculations of β(�p)for concave-down, concave-up, and linear positive (a) and negative(b) chirps. Experimental β(�p) for concave-down, concave-up, andlinear positive (c) and negative (d) chirps. The lines connecting pointsare to guide the eye.

mentioned earlier, this is attributed to the positive chirps havingefficient excitations with no further interactions, whereas thenegative chirp may have further interactions which drive thepopulation back down to the ground state. The general trendin this detuning range is a reduction in β as we approachresonance, which is opposite to the behavior for the positivechirps. Although there are more atom pairs available at largeR (i.e., closer to resonance), this enhancement in initialatom-pair excitation is more than compensated for by theincreased likelihood of coherent collision blocking for theseslow-moving pairs, resulting in less overall loss. We note againthat flux enhancement, due to long-range excitation followedby spontaneous emission and then reexcitation at shorter range[41], is not included in the quantum dynamical calculations.As mentioned in Sec. II, the simulations do take into accountspontaneous emission, but do not allow for the decayed wavefunction to be reexcited. The evolution of the population in the

spontaneous emission sink is shown in Figs. 3(c) and 3(d) forthe negative chirp with �p/2π = −750 MHz. At the variousnegative chirp detunings examined, the sink populations att = 100 ns, averaged over chirp shape and singlet and tripletground states, are ∼0.008 for −950 MHz, ∼0.014 for −750MHz, and ∼0.031 for −550 MHz. This trend indicates thatflux enhancement effects will be most important at the smallestnegative detuning (−550 MHz). In our previous measurementswith linear chirps, and in the corresponding classical MonteCarlo simulations [41], we found that this flux enhancementmechanism dominates the loss for negative chirps centeredcloser to resonance than �p/2π = −600 MHz, consistentwith the sink population trend discussed above. Excludingthe contribution of flux enhancement will thus have the largesteffect at −550 MHz.

We now examine the effect of chirp shape for the negativechirps. At �p/2π = −550 MHz, the various shapes yieldsimilar values of β. However, �p/2π = −750 MHz and−950 MHz are in the collision blocking regime [41]. Specifi-cally, at �p/2π = −750 MHz, the concave-down chirp yieldsa higher β than that of the concave-up or linear chirps. Thereis a similar but smaller separation at �p/2π = −950 MHz,indicating that the values of β are starting to converge. Thisis consistent with measurements at larger detunings [42]where β becomes independent of chirp shape. Since theexcitation occurs at short range where the attractive potentialis steep, the changing excitation radius cannot keep up withthe wave packet motion and multiple interactions are lesslikely to occur. From the wave packet dynamics for �p/2π =−750 MHz [Figs. 5(e) and 5(g)], we see that a larger portion ofthe wave packet created by the concave-down chirp is able tomore quickly accelerate inward on the excited-state potential,giving rise to the larger β. These dynamics also lead to alarger loss per excited atom pair as seen in Fig. 4. In fact, bycomparing Fig. 6(b) and Fig. 4, we conclude that the chirpshape dependence for �p/2π = −750 MHz is due mainly tothe efficiency of loss for a given excitation, as opposed to theamount of initial excitation. The further interactions possiblein this regime are further exemplified through the oscillationsseen in Figs. 3(c) and 3(d).

The measured values of β induced by these nonlinearchirps [42] are shown in Figs. 6(c) and 6(d). As can beseen from the data for the positive chirps in Fig. 6(c), thereis no significant difference in β due to the shaping of thepositive chirps, except possibly at �p/2π = −950 MHz. Thisis expected since the positive chirp sweeps outward, awayfrom the evolving excited-state wave packet. These results aresimilar to those of the simulations shown in Fig. 6(a).

The data shown in Fig. 6(d) for the negative chirps showa dependence on the nonlinearity of the chirp. Specifically,the concave-down chirp is found to have a value of β thatis 50(20)% larger than those of the concave-up and linearchirps at �p/2π = −750 MHz [42]. There is a similar trendat �p/2π = −950 MHz. These variations are also seen inthe simulations [Fig. 6(b)], although to a lesser extent. At�p/2π = −550 MHz, comparing the simulations with themeasurements is problematic because of the important roleof flux enhancement. Nevertheless, it is interesting that bothshow little dependence on the chirp shape. However, thedetuning dependence of the simulations and the measurements

013424-7

J. L. CARINI et al. PHYSICAL REVIEW A 85, 013424 (2012)

is quite different because the former does not include fluxenhancement.

IV. CONCLUSION

In summary, we have presented quantum dynamicalcalculations of excited-state trap-loss collisions caused byfrequency-chirped light. The dependencies on chirp direction,chirp shape, and center detuning have been examined andcompared to the results of a recent experiment. The basic trendsare in agreement. In particular, the loss rate for positive chirpsshows little dependence on chirp shape, while for negativechirps, a dependence is seen for certain center detunings. Thisis attributed to the coherent return of collisional flux to theground state as the resonant radius for excitation follows the ac-celerating excited-state flux inward. Although the simulations

do account for spontaneous emission, only a single attractiveexcited-state potential is considered and the hyperfine structureis ignored. Despite these simplifications, qualitative agreementwith the experiment is found and evidence for coherentcontrol is seen in the dependence on the shape for thenegative chirp. An interesting future direction is to extend thesimulations from trap-loss collisions to photoassociation andto utilize coherent control to optimize the formation of boundmolecules.

ACKNOWLEDGMENTS

The work at the University of Connecticut is supportedby the Chemical Sciences, Geosciences, and BiosciencesDivision, Office of Basic Energy Sciences, US Departmentof Energy.

[1] E. Luc-Koenig and F. Masnou-Seeuws, in Cold Molecules:Theory, Experiment, Applications, edited by R. V. Krems,W. C. Stwalley, and B. Friedrich (CRC Press, Boca Raton, FL,2009).

[2] O. Dulieu, M. Raoult, and E. Tiemann, Special Issue of J. Phys.B 39, No. 19 (2006).

[3] L. D. Carr and J. Ye, Special Issue of New J. Phys. 11, 055009(2009).

[4] Cold and Ultracold Molecules, Faraday Discuss. 142 (2009).[5] R. V. Krems, W. C. Stwalley, and B. Friedrich, eds. Cold

Molecules: Theory, Experiment, Applications (CRC Press, BocaRaton, FL, 2009).

[6] H. J. Metcalf and P. van der Straten, Laser Cooling and Trapping(Springer, New York, 1999).

[7] E. S. Shuman, J. F. Barry, and D. DeMille, Nature (London)467, 820 (2010).

[8] J. T. Bahns, W. C. Stwalley, and P. L. Gould, J. Chem. Phys.104, 9689 (1996).

[9] S. A. Rice and M. Zhao, Optimal Control of Molecular Dynamics(Wiley, New York, 2000).

[10] M. Shapiro and P. Brumer, Principles of Quantum Control ofMolecular Processes (Wiley, New York, 2003).

[11] H. R. Thorsheim, J. Weiner, and P. S. Julienne, Phys. Rev. Lett.58, 2420 (1987).

[12] W. C. Stwalley and H. Wang, J. Mol. Spectrosc. 195, 194 (1999).[13] K. M. Jones, E. Tiesinga, P. D. Lett, and P. S. Julienne, Rev.

Mod. Phys. 78, 483 (2006).[14] W. C. Stwalley, P. L. Gould, and E. E. Eyler, in Cold Molecules:

Theory, Experiment, Applications, edited by R. V. Krems, W. C.Stwalley, and B. Friedrich (CRC Press, Boca Raton, FL, 2009).

[15] J. Vala, O. Dulieu, F. Masnou-Seeuws, P. Pillet, and R. Kosloff,Phys. Rev. A 63, 013412 (2000).

[16] E. Luc-Koenig, R. Kosloff, F. Masnou-Seeuws, andM. Vatasescu, Phys. Rev. A 70, 033414 (2004).

[17] E. Luc-Koenig, M. Vatasescu, and F. Masnou-Seeuws, Eur. Phys.J. D 31, 239 (2004).

[18] C. P. Koch, R. Kosloff, and F. Masnou-Seeuws, Phys. Rev. A 73,043409 (2006).

[19] U. Poschinger, W. Salzmann, R. Wester, M. Weidemuller, C. P.Koch, and R. Kosloff, J. Phys. B 39, S1001 (2006).

[20] C. P. Koch, R. Kosloff, E. Luc-Koenig, F. Masnou-Seeuws, andA. Crubellier, J. Phys. B 39, S1017 (2006).

[21] B. L. Brown and I. A. Walmsley, J. Phys. B 39, S1055(2006).

[22] C. P. Koch, E. Luc-Koenig, and F. Masnou-Seeuws, Phys. Rev.A 73, 033408 (2006).

[23] J. Mur-Petit, E. Luc-Koenig, and F. Masnou-Seeuws, Phys. Rev.A 75, 061404(R) (2007).

[24] S. Kallush, R. Kosloff, and F. Masnou-Seeuws, Phys. Rev. A 75,043404 (2007).

[25] S. Kallush and R. Kosloff, Phys. Rev. A 76, 053408 (2007).[26] S. Kallush and R. Kosloff, Phys. Rev. A 77, 023421 (2008).[27] C. P. Koch, M. Ndong, and R. Kosloff, Faraday Discuss. 142,

389 (2009).[28] S. Ghosal, R. J. Doyle, C. P. Koch, and J. M. Hutson, New J.

Phys. 11, 055011 (2009).[29] E. Luc-Koenig, F. Masnou-Seeuws, and R. Kosloff, Phys. Rev.

A 76, 053415 (2007).[30] W. Salzmann, U. Poschinger, R. Wester, M. Weidemuller,

A. Merli, S. M. Weber, F. Sauer, M. Plewicki, F. Weise, A. M.Esparza, L. Woste, and A. Lindinger, Phys. Rev. A 73, 023414(2006).

[31] B. L. Brown, A. J. Dicks, and I. A. Walmsley, Phys. Rev. Lett.96, 173002 (2006).

[32] W. Salzmann, T. Mullins, J. Eng, M. Albert, R. Wester,M. Weidemuller, A. Merli, S. M. Weber, F. Sauer, M. Plewicki,F. Weise, L. Woste, and A. Lindinger, Phys. Rev. Lett. 100,233003 (2008).

[33] T. Mullins, W. Salzmann, S. Gotz, M. Albert, J. Eng, R. Wester,M. Weidemuller, F. Weise, A. Merli, S. M. Weber, F. Sauer,L. Woste, and A. Lindinger, Phys. Rev. A 80, 063416 (2009).

[34] A. Merli, F. Eimer, F. Weise, A. Lindinger, W. Salzmann,T. Mullins, S. Gotz, R. Wester, M. Weidemuller, R. Aganoglu,and C. P. Koch, Phys. Rev. A 80, 063417 (2009).

[35] D. J. McCabe, D. G. England, H. E. L. Martay, M. E. Friedman,J. Petrovic, E. Dimova, B. Chatel, and I. A. Walmsley, Phys.Rev. A 80, 033404 (2009).

[36] H. E. L. Martay, D. J. McCabe, D. G. England, M. E. Friedman,J. Petrovic, and I. A. Walmsley, Phys. Rev. A 80, 033403(2009).

013424-8

QUANTUM DYNAMICAL CALCULATIONS OF ULTRACOLD . . . PHYSICAL REVIEW A 85, 013424 (2012)

[37] F. Fatemi, K. M. Jones, H. Wang, I. Walmsley, and P. D. Lett,Phys. Rev. A 64, 033421 (2001).

[38] G. Veshapidze, M. L. Trachy, H. U. Jang, C. W. Fehrenbach,and B. D. DePaola, Phys. Rev. A 76, 051401(R) (2007).

[39] M. J. Wright, S. D. Gensemer, J. Vala, R. Kosloff, and P. L.Gould, Phys. Rev. Lett. 95, 063001 (2005).

[40] M. J. Wright, J. A. Pechkis, J. L. Carini, and P. L. Gould, Phys.Rev. A 74, 063402 (2006).

[41] M. J. Wright, J. A. Pechkis, J. L. Carini, S. Kallush,R. Kosloff, and P. L. Gould, Phys. Rev. A 75, 051401(R)(2007).

[42] J. A. Pechkis, J. L. Carini, C. E. Rogers, P. L. Gould, S. Kallush,and R. Kosloff, Phys. Rev. A 83, 063403 (2011).

[43] J. Weiner, V. S. Bagnato, Z. Zilio, and P. S. Julienne, Rev. Mod.Phys. 71, 1 (1999).

[44] R. Kosloff, Annu. Rev. Phys. Chem. 45, 145 (1994).[45] S. Kallush and R. Kosloff, Chem. Phys. Lett. 433, 221 (2006).[46] M. Aymar and O. Dulieu, J. Chem. Phys. 122, 204302 (2005).[47] J. L. Roberts, N. R. Claussen, James P. Burke Jr.,

Chris H. Greene, E. A. Cornell, and C. E. Wieman, Phys. Rev.Lett. 81, 5109 (1998).

[48] P. S. Julienne and J. Vigue, Phys. Rev. A 44, 4464 (1991).

013424-9