calculation of transition probabilities and ac stark...

13
PHYSICAL REVIEW A 81, 062508 (2010) Calculation of transition probabilities and ac Stark shifts in two-photon laser transitions of antiprotonic helium Masaki Hori Max-Planck-Institut f¨ ur Quantenoptik, Hans-Kopfermann-Strasse 1, D-85748 Garching, Germany and Department of Physics, University of Tokyo, Hongo, Bunkyo-ku, Tokyo 113-0033, Japan Vladimir I. Korobov Joint Institute for Nuclear Research 141980, Dubna, Russia (Received 9 February 2010; published 24 June 2010) Numerical ab initio variational calculations of the transition probabilities and ac Stark shifts in two-photon transitions of antiprotonic helium atoms driven by two counter-propagating laser beams are presented. We found that sub-Doppler spectroscopy is, in principle, possible by exciting transitions of the type (n,L)(n 2,L 2) between antiprotonic states of principal and angular momentum quantum numbers n L 1 35, first by using highly monochromatic, nanosecond laser beams of intensities 10 4 –10 5 W/cm 2 , and then by tuning the virtual intermediate state close (e.g., within 10–20 GHz) to the real state (n 1,L 1) to enhance the nonlinear transition probability. We expect that ac Stark shifts of a few MHz or more will become an important source of systematic error at fractional precisions of better than a few parts in 10 9 . These shifts can, in principle, be minimized and even canceled by selecting an optimum combination of laser intensities and frequencies. We simulated the resonance profiles of some two-photon transitions in the regions n = 30–40 of the p 4 He + and p 3 He + isotopes to find the best conditions that would allow this. DOI: 10.1103/PhysRevA.81.062508 PACS number(s): 36.10.k, 31.15.A, 32.70.Jz I. INTRODUCTION The transition frequencies ν exp of antiprotonic helium atoms [15]( pHe + p + e + He 2+ ) have recently been measured by single-photon laser spectroscopy to a fractional precision of 1 part in 10 8 [68]. By comparing these results with three-body QED calculations [913], the antiproton-to- electron mass ratio has been determined as 1836.152674(5) [8,14]. To further increase the experimental precision on ν exp , we have proposed future experiments [15] of sub-Doppler two-photon spectroscopy of pHe + by irradiating the atom with two counterpropagating laser beams [16]. Dynamic (ac) Stark effects are expected to become one of the important sources of systematic error in these future experiments, as is the case with other high-precision laser spectroscopy measurements of atomic hydrogen [1722] and antihydrogen [23,24], molecular hydrogen [25,26], helium [2730], and muonium [3133]. In this paper we calculate the transition probability and ac Stark shift involved in these two-photon transitions using precise three-body wave functions of pHe + . The pHe + atoms [2] can be easily synthesized by simply allowing antiprotons to slow down [3439] and come to rest in a helium target. Some of the antiprotons are captured [4049] into Rydberg pHe + states with large principal (n 38) and angular momentum (L n 1) quantum numbers that have microsecond-scale lifetimes against antiproton annihilation in the helium nucleus. The longevity is due to the ground-state electron in pHe + , which protects the antiproton during colli- sions with other helium atoms [5052]. All laser spectroscopy experiments [8] reported so far have used pulsed lasers [53] to induce single-photon transitions of antiprotons occupying these metastable states to short-lived states with nanosecond- scale lifetimes against Auger emission of the electron [5458]. A Rydberg pHe 2+ ion [5962] then remained after Auger decay, which was rapidly destroyed by collisional Stark effects. The resulting resonance profiles of pHe + had Doppler widths ω D /2π 0.3–1.2 GHz corresponding to the thermal motion of pHe + in the experimental target at T 10 K. This broadening limited the experimental precision on ν exp . The first-order Doppler broadening can, in principle, be reduced (Fig. 1)[15] by irradiating pHe + atoms with two counterpropagating laser beams of angular frequencies ω 1 and ω 2 and inducing e.g., the two-photon transition (n,L) (n 2,L 2). This results in a reduction of ω D by a factor of |(ω 1 ω 2 )/(ω 1 + ω 2 )|. Among a number of possible two-photon transitions, a particularly strong signal is expected for (n,L) = (36,34) (34,32) (Fig. 1), as this involves a large antiproton population [40,41] in the resonance parent state (36,34). Whereas the states (36,34) and (35,33) are metastable with 1-µs-scale lifetimes, the resonance daughter state (34,32) is Auger-dominated with the lifetime τ 4 ns [6,10,54,55] that corresponds to the natural width of a spectral line of 30 MHz. The probability of inducing the transition can be enhanced [15,6369] by tuning ω 1 and ω 2 so that the virtual intermediate state of the two-photon transition lies close (e.g., |ω d /2π | < 10–20 GHz) to the real state (35,33) such that, ω 1 (SI) = 4πR c[E (35,33) E (34,32) (a.u.)] + ω d , (1) ω 2 (SI) = 4πR c[E (36,34) E (35,33) (a.u.)] ω d . Here R c = 3.289842 × 10 15 Hz denotes the Rydberg con- stant and E (n,L) the binding energy of the p 4 He + state (n,L) in atomic units. At experimental conditions where this offset is much larger than the Doppler width (|ω d | ω D ), the two-photon transition is expected to directly transfer the antiprotons populating the parent state (36,34) to the daughter state (34,32), whereas the population in the intermediate state (35,33) will be unaffected. 1050-2947/2010/81(6)/062508(13) 062508-1 ©2010 The American Physical Society

Upload: others

Post on 29-Sep-2020

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Calculation of transition probabilities and ac Stark ...theor.jinr.ru/~korobov/papers/Two_photon_Stark_PRA10.pdf · MASAKI HORI AND VLADIMIR I. KOROBOV PHYSICAL REVIEW A 81, 062508

PHYSICAL REVIEW A 81, 062508 (2010)

Calculation of transition probabilities and ac Stark shifts in two-photon laser transitionsof antiprotonic helium

Masaki HoriMax-Planck-Institut fur Quantenoptik, Hans-Kopfermann-Strasse 1, D-85748 Garching, Germany and

Department of Physics, University of Tokyo, Hongo, Bunkyo-ku, Tokyo 113-0033, Japan

Vladimir I. KorobovJoint Institute for Nuclear Research 141980, Dubna, Russia

(Received 9 February 2010; published 24 June 2010)

Numerical ab initio variational calculations of the transition probabilities and ac Stark shifts in two-photontransitions of antiprotonic helium atoms driven by two counter-propagating laser beams are presented. We foundthat sub-Doppler spectroscopy is, in principle, possible by exciting transitions of the type (n,L)→(n − 2,L − 2)between antiprotonic states of principal and angular momentum quantum numbers n ∼ L − 1 ∼ 35, first byusing highly monochromatic, nanosecond laser beams of intensities 104–105 W/cm2, and then by tuning thevirtual intermediate state close (e.g., within 10–20 GHz) to the real state (n − 1,L − 1) to enhance the nonlineartransition probability. We expect that ac Stark shifts of a few MHz or more will become an important sourceof systematic error at fractional precisions of better than a few parts in 109. These shifts can, in principle, beminimized and even canceled by selecting an optimum combination of laser intensities and frequencies. Wesimulated the resonance profiles of some two-photon transitions in the regions n = 30–40 of the p4He+ andp3He+ isotopes to find the best conditions that would allow this.

DOI: 10.1103/PhysRevA.81.062508 PACS number(s): 36.10.−k, 31.15.A−, 32.70.Jz

I. INTRODUCTION

The transition frequencies νexp of antiprotonic heliumatoms [1–5] (pHe+ ≡ p− + e− + He2+) have recently beenmeasured by single-photon laser spectroscopy to a fractionalprecision of ∼1 part in 108 [6–8]. By comparing these resultswith three-body QED calculations [9–13], the antiproton-to-electron mass ratio has been determined as 1836.152674(5)[8,14]. To further increase the experimental precision on νexp,we have proposed future experiments [15] of sub-Dopplertwo-photon spectroscopy of pHe+ by irradiating the atom withtwo counterpropagating laser beams [16]. Dynamic (ac) Starkeffects are expected to become one of the important sourcesof systematic error in these future experiments, as is the casewith other high-precision laser spectroscopy measurements ofatomic hydrogen [17–22] and antihydrogen [23,24], molecularhydrogen [25,26], helium [27–30], and muonium [31–33]. Inthis paper we calculate the transition probability and ac Starkshift involved in these two-photon transitions using precisethree-body wave functions of pHe+.

The pHe+ atoms [2] can be easily synthesized by simplyallowing antiprotons to slow down [34–39] and come to rest ina helium target. Some of the antiprotons are captured [40–49]into Rydberg pHe+ states with large principal (n ∼ 38) andangular momentum (L ∼ n − 1) quantum numbers that havemicrosecond-scale lifetimes against antiproton annihilation inthe helium nucleus. The longevity is due to the ground-stateelectron in pHe+, which protects the antiproton during colli-sions with other helium atoms [50–52]. All laser spectroscopyexperiments [8] reported so far have used pulsed lasers [53]to induce single-photon transitions of antiprotons occupyingthese metastable states to short-lived states with nanosecond-scale lifetimes against Auger emission of the electron [54–58].A Rydberg pHe2+ ion [59–62] then remained after Augerdecay, which was rapidly destroyed by collisional Stark

effects. The resulting resonance profiles of pHe+ had Dopplerwidths �ωD/2π ∼ 0.3–1.2 GHz corresponding to the thermalmotion of pHe+ in the experimental target at T ∼ 10 K. Thisbroadening limited the experimental precision on νexp.

The first-order Doppler broadening can, in principle, bereduced (Fig. 1) [15] by irradiating pHe+ atoms with twocounterpropagating laser beams of angular frequencies ω1

and ω2 and inducing e.g., the two-photon transition (n,L) →(n − 2,L − 2). This results in a reduction of �ωD by afactor of |(ω1 − ω2)/(ω1 + ω2)|. Among a number of possibletwo-photon transitions, a particularly strong signal is expectedfor (n,L) = (36,34) → (34,32) (Fig. 1), as this involves a largeantiproton population [40,41] in the resonance parent state(36,34). Whereas the states (36,34) and (35,33) are metastablewith 1-µs-scale lifetimes, the resonance daughter state (34,32)is Auger-dominated with the lifetime τ ∼ 4 ns [6,10,54,55]that corresponds to the natural width of a spectral line of∼30 MHz.

The probability of inducing the transition can be enhanced[15,63–69] by tuning ω1 and ω2 so that the virtual intermediatestate of the two-photon transition lies close (e.g., |�ωd/2π | <

10–20 GHz) to the real state (35,33) such that,

ω1(SI) = 4πR∞c[E(35,33) − E(34,32)(a.u.)] + �ωd,(1)

ω2(SI) = 4πR∞c[E(36,34) − E(35,33)(a.u.)] − �ωd.

Here R∞c = 3.289842 × 1015 Hz denotes the Rydberg con-stant and E(n,L) the binding energy of the p4He

+state (n,L)

in atomic units. At experimental conditions where this offsetis much larger than the Doppler width (|�ωd | � �ωD),the two-photon transition is expected to directly transfer theantiprotons populating the parent state (36,34) to the daughterstate (34,32), whereas the population in the intermediate state(35,33) will be unaffected.

1050-2947/2010/81(6)/062508(13) 062508-1 ©2010 The American Physical Society

Page 2: Calculation of transition probabilities and ac Stark ...theor.jinr.ru/~korobov/papers/Two_photon_Stark_PRA10.pdf · MASAKI HORI AND VLADIMIR I. KOROBOV PHYSICAL REVIEW A 81, 062508

MASAKI HORI AND VLADIMIR I. KOROBOV PHYSICAL REVIEW A 81, 062508 (2010)

∆ωd

ω2 laser

ω1 laser

(n,L)=(36,34)

(35,33)

Virtual state

(34,32)

FIG. 1. Energy level diagram indicating the two-photon transition(n,L) = (36,34) → (34,32) in p4He+. The relative position of thevirtual intermediate state for a detuning frequency �ωd of the twocounterpropagating laser beams is shown.

This paper is organized in the following way. Some detailsof the numerical methods are described in Sec. II. Thetransition amplitude of the two-photon resonance (36,34) →(34,32) of p4He, and the polarizabilities of the parent anddaughter states (36,34) and (34,32) at various offsets �ωd areestimated in Secs. III A and III B. We next calculate the “back-ground” polarizability due to the contributions of pHe+ statesother than the resonant intermediate state (35,33) (Sec. III C).Based on these results, the ac Stark shift and broadening aresemi-analytically estimated in Sec. III D. We next discuss thehyperfine structure in two-photon transitions of the p4He

+and

p3He+

isotopes (Sec. III E) [70–75]. We numerically simulatethe profile of several two-photon resonances (Sec. III F) beforeconcluding the paper. Atomic units (a.u.) are used to evaluatethe state polarizabilities and transition amplitudes [76–81],whereas the International System of Units (SI) is used for thetransition frequencies, rates, and laser intensities relevant forfuture spectroscopy experiments.

II. DETAILS OF THE CALCULATION

For simplicity we take the linearly polarized laser fieldaligned along the z axis, such that the perturbation Hamiltonianin a laser field of frequency ω and amplitude F has the form

H ′ = −eF(t) d, F(t) = ezF cos(wt). (2)

Here d = ∑3a=1 ZaRa is the electric dipole moment operator.

The second-order correction E(2) to the unperturbed eigenen-ergy E0 of a pHe+ state vector |0〉 may then be expressedas

E(2) = − 12αzz

d (ω,M2)F 2, (3)

where aij

d (ω) is a tensor of the dynamic dipole polarizability

αij

d = −∑

q

[ 〈0|di |q〉〈q|dj |0〉E0 − Eq + ω

+ 〈0|di |q〉〈q|dj |0〉E0 − Eq − ω

]. (4)

The energy of a pHe+ state vector |q〉 is denoted by Eq , andthe summation of q is over all states which are accessible viasingle-photon transition from the resonance parent state |0〉.

The tensor aij

d (ω) may be rewritten in terms of theirreducible scalar and tensor polarizability operators

αij

d = αs + αt

[LiLj + Lj Li − 2

3L2

],

αs = 1

3[a+ + a0 + a−], (5)

αt = − a+L(2L − 1)

3(L + 1)(2L + 3)+ a0(2L − 1)

3(L + 1)− a−

3,

where the angular momentum operator is denoted by L. Thecoefficients a+, a0, and a− are defined as follows

a+ = − 2

2L + 1

∑q

(E0 − Eq)|〈0L‖d‖q(L + 1)〉|2(E0 − Eq)2 − ω2

,

a0 = − 2

2L + 1

∑q

(E0 − Eq)|〈0L‖d‖qL〉|2(E0 − Eq)2 − ω2

, (6)

a− = − 2

2L + 1

∑q

(E0 − Eq)|〈0L‖d‖q(L − 1)〉|2(E0 − Eq)2 − ω2

.

Here a+ and a− represent the contributions from antiprotontransitions to states of normal parity which change the orbitalangular momentum quantum number L of the antiproton by 1or −1. The contribution a0 involves transitions to pHe+ statesof anomalous parity in which the L value is unchanged andthe 1s electron is excited to the 2p state.

For our analysis it is convenient to define “background”polarizabilities using the previous equations, where the dom-inant contribution from the intermediate state of the two-photon transition is subtracted. For example, the correspondingscalar and tensor background polarizabilities of state (n,L) =(36,34) can be calculated as

βs = αs + 2 (E0 − Ei)

3(2L + 1)

|〈0L‖d‖i(L − 1)〉|2(E0 − Ei)2 − ω2

,

(7)

βt = αt − 2(E0 − Ei)

3(2L + 1)

|〈0L‖d‖i(L − 1)〉|2(E0 − Ei)2 − ω2

.

Here E0 and Ei denote the energies of state (36,34) and theintermediate state (35,32).

The transition matrix element κL,L−2,M of the pHe+ two-photon transition (n,L,M) → (n − 2,L − 2,M) induced bytwo linearly polarized laser beams of total frequency w1 +w2 ≈ 4πR∞c(E0 − E1) can be calculated using the Wigner3j symbols as

κL,L−2,M = −(

L − 1 1 L

M 0 −M

)(L − 1 1 L − 2

M 0 −M

)

×∑

q

[〈1(L− 2)‖d‖q(L− 1)〉〈q(L− 1)‖d‖0L〉E0 − Eq − ω2

+ 〈1(L − 2)‖d‖q(L − 1)〉〈q(L − 1)‖d‖0L〉E0 − Eq − ω1

],

(8)

062508-2

Page 3: Calculation of transition probabilities and ac Stark ...theor.jinr.ru/~korobov/papers/Two_photon_Stark_PRA10.pdf · MASAKI HORI AND VLADIMIR I. KOROBOV PHYSICAL REVIEW A 81, 062508

CALCULATION OF TRANSITION PROBABILITIES AND . . . PHYSICAL REVIEW A 81, 062508 (2010)

wherein |1〉 denotes the state vector of the resonance daughterstate of energy E1. This is related to the two-photon Rabioscillation frequency (in atomic units) of this laser transitionvia the equation

2γM (a.u.)

2π= 1

2 |κL,L−2,M |F1F2. (9)

The last term in Eq. (8) can be neglected at small offsetfrequencies �wd .

To calculate these quantities we must evaluate the reducedmatrix elements for the dipole operator d and diagonalizethe Hamiltonian. For this we use the variational exponentialexpansion described in Ref. [9]. The wave function for a statewith a total orbital angular momentum L and of a total spatialparity π = (−1)L is expanded as follows

�πLM (R,r1) =

∑l1+l2=L

Y l1l2LM (R,r1)GLπ

l1l2(R,r1,r2),

GLπl1l2

(R,r1,r2) =N∑

n=1

{CnRe[e−αnR−βnr1−γnr2 ]

+ DnIm[e−αnR−βnr1−γnr2 ]},

(10)

where the complex exponents α, β, and γ are generated ina pseudorandom way, R and r1 are position vectors of anantiproton and an electron with respect to a helium nucleus, andr2 the distance between the antiproton and electron. Furtherdetails may be found in Refs. [9,10].

This method has been previously employed to calculatethe nonrelativistic energies of pHe+ with a relative precisionof around one part in 1012 [9–11]. We here determined thenonrelativistic values of the transition matrix elements usingthe same wave functions to a precision of ∼10−6. This wasmore than adequate for our purpose of roughly estimatingthe two-photon transition probabilities and ac Stark shiftsrelevant to future spectroscopy experiments. The pHe+ statesof unnatural (or anomalous) parity (−1)L+1 involve an electronin an excited state and therefore are not metastable [82]. Theywill not be considered here.

III. RESULTS

A. Transition matrix element

We first calculated the transition amplitude κL,L−2,M for thetwo-photon resonance (n,L) = (36,34) → (34,32) in p4He

+

at various offsets ωd from the intermediate (35,33) state andestimated the laser intensities needed to drive this transition.In Fig. 2, sequences of single-photon transitions connectingthe states (36,34) and (34,32) are indicated by straightarrows, together with the corresponding dipole moment|〈0L‖d‖q(L − 1)〉|. The atomic units shown here can beconverted to SI units, the corresponding spontaneous decayrates in s−1 obtained using the equation

γ (SI) = e2a20

4πε0

4ω3q0

3hc3

|〈0L‖d‖q(L − 1)〉|2(a.u.)

2L + 1. (11)

The SI-unit constants that appear in the above equation are e,the elementary charge; a0, the Bohr radius; ε0, the dielectricconstant of vacuum; h, the reduced Planck constant; and c thespeed of light. The angular transition frequency between states|q〉 and |0〉 are denoted by ωq0.

Two types of transitions (n,L) → (n,L − 1) and (n,L) →(n − 1,L − 1) have the largest amplitudes of ∼1 a.u., but thelatter kind, which conserve the vibrational quantum numberv = n − L − 1 and involve fluorescence photons of frequencyωq0/2π ∼ 1015 Hz, constitute the dominant channels of spon-taneous decay. These transitions are most favorable for laserspectroscopy. The transition frequencies, dipole moments,and decay rates of some single-photon resonances in p4He

+

and p3He+

of the type (n,L) → (n − 1,L − 1) are shown inTable I. The dipole moments for the higher-lying infraredtransitions involving states with n ∼ 40 are relatively large(>2 a.u.), whereas for UV transitions in the n � 33 regions itis reduced to <1 a.u. On the other hand, the radiative decayrates increase for lower-n transitions [e.g., from 4.7 × 105 s−1

for (40,36)→(39,35)] to 6.6 × 105 s−1 for (32,31) → (31,30)due to the ω3

q0 dependence.Using the single-photon dipole moments calculated previ-

ously, we derive the two-photon transition amplitude κL,L−2,M

Leve

l en

ergy

eV

L=29 30 31 32 33 34

-88

-86

-84

-82

n=34

35

36

1.16

4.9

x 10

-3

0.62

1.38

1.13

6.2

x 10

-3

30

n=31

FIG. 2. (Color online) Portion of the energy level diagram of p4He+. The solid lines indicate radiation-dominated metastable states, thewavy lines Auger-dominated short-lived states. The broken lines indicate pHe2+ ionic states formed after Auger emission, and the curvedarrows Auger transitions with minimum |�L|. Calculated values of the dipole moments for some single-photon radiative transitions are shownin atomic units.

062508-3

Page 4: Calculation of transition probabilities and ac Stark ...theor.jinr.ru/~korobov/papers/Two_photon_Stark_PRA10.pdf · MASAKI HORI AND VLADIMIR I. KOROBOV PHYSICAL REVIEW A 81, 062508

MASAKI HORI AND VLADIMIR I. KOROBOV PHYSICAL REVIEW A 81, 062508 (2010)

TABLE I. Some single-photon transitions in p4He+ and p3He+:their vibrational quantum number v, transition frequencies, dipolemoments, and radiative rates.

Trans. freq. Dipole Rad. rate(n,L) → (n′,L′) v (THz) moment (a.u.) (105 s−1)

p4He+ states(40,36) → (39,35) 3 444.8 2.28 4.71(39,35) → (38,34) 3 501.9 2.02 5.44(38,35) → (37,34) 2 566.1 1.82 6.36(37,34) → (36,33) 2 636.9 1.58 7.01(37,35) → (36,34) 1 638.6 1.61 7.15(36,34) → (35,33) 1 717.5 1.38 7.63(35,33) → (34,32) 1 804.6 1.16 7.92(33,32) → (32,31) 0 1012.4 0.79 7.42(32,31) → (31,30) 0 1132.6 0.62 6.62

p3He+ states(39,35) → (38,34) 3 445.8 2.30 4.96(38,34) → (37,33) 3 505.2 2.03 5.77(37,34) → (36,33) 2 572.0 1.83 6.81(36,33) → (35,32) 2 646.2 1.58 7.55(35,33) → (34,32) 1 730.8 1.37 8.27(34,32) → (33,31) 1 822.8 1.15 8.59(33,32) → (32,31) 0 928.8 0.95 8.44(32,31) → (31,30) 0 1043.1 0.77 7.94

of the resonance (36,34) → (34,32) in p4He+

for caseswhere the virtual intermediate state is offset over a largerange between �ωd/2π = −0.6 and 0.6 PHz from the state(35,33). The atom is excited by two linearly polarized,counterpropagating laser beams. Figures 3(a) through 3(c)show the amplitude |κL,L−2| averaged over all ∼70 transi-tions between the magnetic substates which conserve the M

value

|κL,L−2|2 = 1

2L + 1

∑M

|κL,L−2,M |2. (12)

The |κL,L−2| values are usually small [e.g., a few a.u. forlasers of equal frequency (ω1 = ω2)]. This is smaller thanthe amplitude ∼7.8 a.u. [20] for the 1s–2s two-photontransition of atomic hydrogen excited by 243-nm laser light.Gigawatt-scale laser intensities would be needed to induce theantiprotonic transition within the microsecond-scale lifetimeof pHe+. On the other hand, the transition probabilities canbe strongly enhanced to � 103 a.u. by tuning the virtualintermediate state within ∼20 GHz of the real states (n,L) =(34,33), (35,33), or (36,33). Equation (9) indicates that thetransition can then be induced using nanosecond laser pulsesof electric field F ∼ (1 − 2) × 10−6 a.u. According to theequation,

I (SI) = 1

2ε0c

(eF (a.u.)

4πε0a20

)2

, (13)

this corresponds to a peak intensity of I ∼ 104–105 W/cm2,which is achievable using titanium sapphire lasers of narrowlinewidth [16]. The second maximum at ωd/2π ∼ −85 THzin Fig. 3(b) corresponds to the case of the ω1 and ω2 lasers

0

0.1

0.2

0.3

0.4

-0.5 -0.25 0 0.25 0.5

Virtual state offset ∆ωd / 2π (PHz)

Tra

nsi

tio

n a

mp

litu

de

(a

.u.)

(a)

0

10

20

-100 0 100Virtual state offset ∆ωd / 2π (THz)

Tra

nsi

tio

n a

mp

litu

de

(a

.u.)

ω1 = ω2

(b)

0

1

2

-10 0 10

Virtual state offset ∆ωd / 2π (GHz)

Tra

nsi

tio

n a

mp

litu

de

(a

.u.) x104

(c)

FIG. 3. Transition matrix amplitude |κL,L−2| of the two-photonresonance (n,L) = (36,34) → (34,32) as a function of the offset�ωd/2π in the two laser frequencies ω1 and ω2 from the virtual in-termediate state. Transition probability is averaged over the magneticquantum number M , see text. Arrow indicates the position where thetwo laser beams have equal frequencies.

resonating with the respective transitions (36,34) → (35,33)and (35,33) → (34,32).

B. Polarizabilities

We next evaluate the polarizabilities of the parent anddaughter states of the transition. In Figs. 4(a) and 4(c),the scalar and tensor polarizability components αs(ω2)(36,34),αt (ω2)(36,34), of state (n,L) = (36,34) are shown. To simplifythe calculation, we initially assume that the atom is irradiatedby a single laser field of frequency ω2 (see Fig. 1) correspond-ing to offsets between �ωd/2π = −20 and 20 GHz from thestate (35,33). A similar plot for the polarizabilities αs(ω1)(34,32)

and αt (ω1)(34,32) of state (34,32) under irradiation by a laserfield of ω1 are shown in Figs. 4(b) and 4(d).

062508-4

Page 5: Calculation of transition probabilities and ac Stark ...theor.jinr.ru/~korobov/papers/Two_photon_Stark_PRA10.pdf · MASAKI HORI AND VLADIMIR I. KOROBOV PHYSICAL REVIEW A 81, 062508

CALCULATION OF TRANSITION PROBABILITIES AND . . . PHYSICAL REVIEW A 81, 062508 (2010)

-2

-1

0

1

2

-20 0 20

(n,L)=(36,34)

(a)x104

Virtual state offset ∆ωd / 2π (GHz)

Po

lari

zab

ility

αs

(a.

u.)

-2

-1

0

1

2

-20 0 20

(n,L)=(34,32)

(b)x104

Virtual state offset ∆ωd / 2π (GHz)

Po

lari

zab

ility

αs

(a.

u.)

-2

-1

0

1

2

-20 0 20

(c)x104

Virtual state offset ∆ωd / 2π (GHz)

Po

lari

zab

ility

αt

(a.

u.)

-2

-1

0

1

2

-20 0 20

(d)x104

Virtual state offset ∆ωd / 2π (GHz)

Po

lari

zab

ility

αt

(a.

u.)

FIG. 4. (a)–(b) Scalar dipole polarizability of states (n,L) = (36,34) and (34,32) versus frequency offset �ωd/2π of the virtual intermediatestate from the real state (35,33): αs(w2)(36,34) and αs(w1)(34,32) as defined in Eq. (5). (c)–(d) Tensor polarizabilities αt (w2)(36,34) and αt (w1)(34,32).

As the laser frequencies are offset from �ωd/2π = −100to −6 GHz, the scalar polarizabilities decrease from ∼ − 500to ∼ − 1 × 104 a.u., whereas the tensor polarizabilities havethe opposite sign and increase from ∼500 to ∼1 × 104 a.u.(Table II). These polarizabilities of 1000–10,000 a.u. corre-spond to 50–500 Hz/(W/cm2) in SI units according to theequation

α(a.u.) = α(SI)h2

mea40α

. (14)

The graphs follow a reciprocal �ω−1d dependence and are

approximately symmetric with respect to the origin. This

simple behavior suggests that the ac Stark shift is primarilycaused by the contribution from the intermediate state (35,33).

C. Contribution of nonresonant states

We next study the nonresonant or “background” contri-butions [Eq. (7)] to the polarizabilities from all pHe+ statesother than the intermediate state (n,L) = (35,33). This willallow us to estimate how far the measured ac-Stark shiftwill deviate from the predictions of a simple three-levelmodel which include only the resonance parent, daughter, andintermediate states. Figures 5(a) and 5(c) show the backgroundscalar and tensor polarizabilities βs(ω2)(36,34) and βt (ω2)(36,34)

of state (36,34) when irradiated with a single laser field

TABLE II. Scalar and tensor polarizabilities of the p4He+ states (36,34) and (34,32) at offset frequencies of the lasers between �ωd/2π =−100 and 100 GHz, see text.

Polarizabilities (a.u.)

�ωd/2π −100 GHz −12 GHz −6 GHz 6 GHz 12 GHz 100 GHz

(n,L) = (36,34) stateαs(ω2)(36,34) −602 −5.02 × 103 −1.00 × 104 1.00 × 104 5.02 × 103 602αs(ω1)(36,34) −1.61 −1.61 −1.61 −1.61 −1.61 −1.61αt (ω2)(36,34) 601 5.02 × 103 1.00 × 104 −1.00 × 104 −5.02 × 103 −603αt (ω1)(36,34) 0.204 0.203 0.203 0.203 0.203 0.203βs(ω2)(36,34) −0.403 −0.401 −0.401 −0.401 −0.401 −0.400βt (ω2)(36,34) −0.945 −0.946 −0.946 −0.947 −0.947 −0.948

(n,L) = (34,32) stateαs(ω2)(34,32) −2.06 −2.06 −2.06 −2.06 −2.06 −2.06αs(ω1)(34,32) −459 −3.81 × 103 −7.62 × 103 7.62 × 103 3.81 × 103 456αt (ω2)(34,32) 0.265 0.265 0.265 0.265 0.265 0.264αt (ω1)(34,32) 416 3.47 × 103 6.94 × 103 −6.95 × 103 −3.48 × 103 −417βs(ω1)(34,32) −1.36 −1.36 −1.36 −1.36 −1.36 −1.36βt (ω1)(34,32) −0.401 −0.401 −0.401 −0.401 −0.402 −0.402

062508-5

Page 6: Calculation of transition probabilities and ac Stark ...theor.jinr.ru/~korobov/papers/Two_photon_Stark_PRA10.pdf · MASAKI HORI AND VLADIMIR I. KOROBOV PHYSICAL REVIEW A 81, 062508

MASAKI HORI AND VLADIMIR I. KOROBOV PHYSICAL REVIEW A 81, 062508 (2010)

-0.4015

-0.4014

-0.4013

-0.4012

-0.4011

-0.401

-0.4009

-20 0 20

(n,L)=(36,34)

(a)

Virtual state ofset ∆ωd / 2π (GHz)

Po

lari

zab

ility

βs

(a.

u.)

-1.3614

-1.3613

-1.3612

-1.3611

-1.361

-1.3609

-1.3608

-20 0 20

(n,L)=(34,32)

(b)

Virtual state offset ∆ωd / 2π (GHz)

Po

lari

zab

ility

βs

(a.

u.)

-0.9468

-0.9467

-0.9466

-0.9465

-0.9464

-0.9463

-0.9462

-20 0 20

(c)

Virtual state offset ∆ωd / 2π (GHz)

Po

lari

zab

ility

βt

(a.

u.)

-0.4018

-0.4017

-0.4016

-0.4015

-0.4014

-0.4013

-0.4012

-20 0 20

(d)

Virtual state offset ∆ωd / 2π (GHz)P

ola

riza

bili

ty β

t (

a.u

.)

FIG. 5. Residual scalar dipole polarizability of states (n,L) = (36,34) and (34,32) versus frequency offset �ωd/2π of the virtualintermediate state from the real state (35,33): βs(ω2)(36,34) and βs(ω1)(34,32), and βt (ω2)(36,34) and βt (ω1)(34,32). The contributions of all statesexcept the intermediate one (35,33) are included.

of frequency ω2 at offsets between �ωd/2π = −20 and20 GHz. They remained relatively constant at ∼ − 0.40 and0.95 a.u., respectively (Table II). Figures 5(b) and 5(d) are thecorresponding plots of βs(ω1)(34,32) and βt (ω1)(34,32) for state(34,32) irradiated with the ω1 laser. They are similarly constant(−1.36 and −0.40 a.u.). All these background polarizabilitiesare at least three orders of magnitude smaller than the dominantcontributions to αs and αt arising from the intermediate state(35,33) at small offsets |�ωd/2π | < 12 GHz.

The case of two counterpropagating laser fields of angularfrequencies ω1 and ω2, and amplitudes F1 and F2 irradiatingthe atom simultaneously will next be considered. The pertur-bation Hamiltonian for this can be expressed as,

H ′ = −eF(t)d,

F(t) = ez [F1 cos(w1t) + F2 cos(w2t)] .(15)

As shown in Ref. [17], the interference effect between thetwo laser fields can be neglected in the case of ω1 �= ω2.The contribution to the ac Stark shift of state (36,34)from the ω1 laser [which is far off-resonance with respectto the upper single-photon transition (36,34) → (35,33)] isexpressed by the scalar and tensor polarizabilities αs(ω1)(36,34)

and αt (ω1)(36,34). The calculated values at offsets between�ωd/2π = −20 and 20 GHz were, respectively, −1.6 and−0.2 a.u. (Table II). The corresponding values for the daughterstate αs(ω2)(34,32) and αt (ω2)(34,32) were also small (−2.1 and0.3 a.u.).

We conclude that the two-photon spectroscopy experimentdepicted in Fig. 1 can be accurately simulated by a simplemodel involving three states interacting with two laser beams.Any nonresonant contribution from other pHe+ states are atleast three orders of magnitude smaller.

D. ac Stark shifts of transition frequency

The ac Stark shift in the transition frequency of the(36,34) → (34,32) resonance can be analytically estimatedas

�ωac,M

2π= 2R∞c

[E

(2)(36,34,M) − E

(2)(34,32,M)

], (16)

wherein the ac Stark shifts E(2)(36,34,M) and E

(2)(34,32,M) in the

parent and daughter states of magnetic substate M induced bythe two linearly polarized laser fields can be approximated as

E(2)(36,34,M) ∼ −1

2

[αs(ω2)(36,34) + 3M2 − 1190

2278αt (ω2)(36,34)

]F 2

2 ,

E(2)(34,32,M) ∼ −1

2

[αs(ω1)(34,32) + 3M2 − 1056

2016αt (ω1)(34,32)

]F 2

1 .

(17)

In Fig. 6(a), the ac Stark shift �ωac,M/2π for M values0, ±16, and ±32, and two laser offsets �ωd/2π = −12and 12 GHz obtained from the above equations are plottedas a function of the intensity ratio I2/I1 = F 2

2 /F 21 between

the two laser beams. Here I1 is fixed at ∼5 × 104 W/cm2

while I2 is scanned between (3 − 5) × 104 W/cm2. We find apositive shift (�ωac > 0) at two combinations of laser offsetsand intensities (�ωd > 0, I1 � I2) and (�ωd < 0, I1 � I2).Conversely the shift is negative �ωac < 0 at (�ωd < 0,I1 � I2) and (�ωd > 0, I1 � I2). In addition to the ac Starkshift, the tensor polarizabilities cause the resonance line tosplit depending on the M value of the involved states. Atconditions of I2/I1 < 0.6 or > 1 the ac Stark shift and splittingcan reach values of more than 5–10 MHz. At smaller offsets|�ωd/2π | = 6 GHz, the ac Stark shift and splitting becometwice as large [Fig. 6(b)].

062508-6

Page 7: Calculation of transition probabilities and ac Stark ...theor.jinr.ru/~korobov/papers/Two_photon_Stark_PRA10.pdf · MASAKI HORI AND VLADIMIR I. KOROBOV PHYSICAL REVIEW A 81, 062508

CALCULATION OF TRANSITION PROBABILITIES AND . . . PHYSICAL REVIEW A 81, 062508 (2010)

-15

-10

-5

0

5

10

15

0.6 0.8 1

(a)

12 GHz, M=0

12 GHz, M=±16

12 GHz, M=±32

-12 GHz, M=0

-12 GHz, M=±16

-12 GHz, M=±32

Laser intensity ratio I2 / I1

ac S

tark

sh

ift

∆ω

ac /

(M

Hz)

-15

-10

-5

0

5

10

15

0.6 0.8 1

(b)

6 GHz, M=0

6 GHz, M=±16

6 GHz, M=±32

-6 G

Hz, M=0

-6 GHz,

M=±16

-6 GHz, M=±32

Laser intensity ratio I2 / I1

ac S

tark

sh

ift

∆ω

ac /

(M

Hz)

FIG. 6. (Color online) The ac Stark shifts in the p4He+ resonance (36,34) → (34,32) estimated from the state polarizabilities αs and αt , asa function of the intensity ratio I2/I1 between the two laser beams with I1 fixed at 5 × 104 W/cm2. Shifts for transitions involving the magneticsubstates M = 0, ±16, and ±32 are indicated. Virtual intermediate state is tuned (a) �ωd/2π = ±12 GHz and (b) ±6 GHz from the state(35,33).

The ac Stark shift arising from αs and the splitting due toαt can be minimized by adjusting the laser intensities to thevalues indicated by the filled circles in Figs. 6(a) and 6(b),

I2

I1∼ αs(ω1)(34,32)

αs(ω2)(36,34)∼ 0.76. (18)

An important point is that when the sign of the laser detuning isreversed (e.g., from �ωd/2π = −12 to 12 GHz) the resultingac Stark shift also reverses sign, but its magnitude is thesame

�ωac(ω1,ω2) = −�ωac(ω1 − 2�ωd,ω2 + 2�ωd ). (19)

This means that the residual ac Stark shift can be canceledby comparing the two-photon transition frequencies measuredat laser offsets of opposite sign but the same absolute value(i.e., �ωd and −�ωd ).

Figure 7 shows the ac Stark shifts and transition amplitudes|κL,L−2,M | of all magnetic sublines between M = −32 and32 of this two-photon resonance at laser offsets �ωd/2π =12 GHz (a)–(c) and −12 GHz (d)–(f). The shifts werecalculated at three ratios of the laser intensities I2/I1 = 0.6,0.78, and 1.0 with I1 fixed at 5 × 104 W/cm2. The ac Starkeffect causes a triangular profile with the M = 0 transitionsbeing strongest and shifting the most. In actual experimentsinvolving pulsed laser beams, these shifts and splittings wouldsmear out due to nanosecond-scale changes in the light fieldintensities I1(t) and I2(t); the observed intensities of themagnetic sublines would have a nonlinear dependence on|κL,L−2,M |2, I1(t), and I2(t) as simulated in Sec. III F.

E. Hyperfine structure

We next study the hyperfine lines that appear in thetwo-photon resonance profile. The hyperfine substates of theparent, intermediate, and daughter states (n,L), (n − 1,L − 1),

and (n − 2,L − 2) in p4He+

are shown schematically inFig. 8(a). Due to the dominant interaction between the electronspin Se and the antiproton orbital angular momentum L,a pair of fine-structure sublevels of intermediate angularmomentum quantum number F = L ± 1/2 and a splitting10–15 GHz arise. The interactions involving the antipro-ton spin Sp cause each fine-structure sublevel to split bya few hundred MHz into pairs of hyperfine sublevels oftotal angular momentum quantum number J = F ± 1/2. InFig. 8(a), the spin orientations of the electron and antiproton(Se,Sp) are indicated for the four hyperfine sublevels. Forexample, the energetically highest-lying component consistsof a spin-down electron and spin-up antiproton [i.e., (Se,Sp) =(↓↑)].

In the p3He+

case [Fig. 8(b)], the electronic fine structuresublevels of F = L ± 1/2 are similarly split by ∼10 GHz.Each fine-structure sublevel is then split into pairs of 3Hehyperfine sublevels of intermediate angular momentum K =F ± 1

2 arising from the interactions involving the nuclear spinSh. The antiproton spin gives rise to eight hyperfine sublevelsof total angular momentum J = K ± 1

2 . The spin orientationsof the three constituent particles (Se,Sh,Sp) are indicated foreach substate in Fig. 8(b).

In Figs. 8(a) and 8(b), the four and eight strongest two-photon transitions between the hyperfine sublevels of p4He

+

and p3He+

are indicated by arrows. These transitions passthrough the virtual intermediate state without flipping thespin of any constituent particle. Many other transitions arepossible, but they all involve spin-flip and so their amplitudesare suppressed by three orders of magnitude or more.

F. Optical rate equations

To simulate pHe+ two-photon resonance profiles, weused the following nonlinear rate equations which describe a

062508-7

Page 8: Calculation of transition probabilities and ac Stark ...theor.jinr.ru/~korobov/papers/Two_photon_Stark_PRA10.pdf · MASAKI HORI AND VLADIMIR I. KOROBOV PHYSICAL REVIEW A 81, 062508

MASAKI HORI AND VLADIMIR I. KOROBOV PHYSICAL REVIEW A 81, 062508 (2010)

0

5M=0

M=±32

(a): I2 / I1 = 0.6

0

5M=0

M=±32

(d): I2 / I1 = 0.6

0

5(b): I2 / I1 = 0.78

0

5(e): I2 / I1 = 0.78

0

5

-10 0 10

M=0

M=±32

ac Stark shift ∆ωac,M / 2π (MHz)

Laser detuning ∆ωd / 2π = 12 GHz

(c): I2 / I1 = 1.0

x103

Tw

o-p

ho

ton

tra

nsi

tio

n a

mp

litu

de

(

a.u

.)

0

5

-10 0 10

M=0

M=±32

ac Stark shift ∆ωac,M / 2π (MHz)

Laser detuning ∆ωd / 2π = -12 GHz

(f): I2 / I1 = 1.0

x103

Tw

o-p

ho

ton

tra

nsi

tio

n a

mp

litu

de

(

a.u

.)

FIG. 7. Expected positions and transition amplitudes |κL,L−2,M | of the M sublines of the two-photon transition (36,34) → (34,32) inp4He+, wherein the virtual intermediate state is offset (a)–(c) �ωd/2π ∼ 12 GHz and (d)–(f) −12 GHz from the state (35,33). The intensityratios I2/I1 between the two laser beams were varied between 0.6, 0.78, and 1.0 in each figure, whereas I1 was kept constant at ∼5 ×104 W/cm2.

three-level model,

∂ρaa

∂t= −Im(2Mρab),

∂ρbb

∂t= Im(2Mρab) − Im(1Mρbc),

∂ρcc

∂t= −γcρcc + Im(1Mρbc),

∂ρab

∂t= −idabρab + i

2M

2(ρaa − ρbb) + i

1M

2ρac,

∂ρbc

∂t= −idbcρbc + i

1M

2(ρbb − ρcc) − i

2M

2ρac,

∂ρac

∂t= −idacρac + i

1M

2ρab − i

2M

2ρbc.

(20)

Here the density matrix ρaa , ρbb, and ρcc represent the an-tiproton populations in the parent, intermediate, and daughterstates. The mixing between pairs of states induced by the lasersare denoted by ρab, ρbc, and ρac, the Auger decay rate of thedaughter state by γc. The three detunings that appear in Eq. (20)can be calculated using the equations

dab = Eb − Ea −(

1 + vz

c

)ω2 − i

γa + γb

2,

dbc = Ec − Eb −(

1 − vz

c

)ω1 − i

γb + γc

2,

dac = Ec − Ea −(

1 + vz

c

)ω2 −

(1 − vz

c

)ω1 − i

γa + γc

2,

(21)

where vz denotes the velocity component of the atom in thedirection of the ω2 laser beam, Ea , Eb, and Ec the bindingenergy of the hyperfine states involved in the transition andγa and γb the radiative rates of the parent and intermedi-ate states. The angular Rabi frequencies of single-photontransitions induced between the parent and intermediatestates of magnetic quantum number M , and the intermediateand daughter states are, respectively, denoted by 2M and1M .

Values of 2M/2π (in s−1) for the transition(0,L,F,K,J,M) → (q,L − 1,F ′,K ′,J ′,M) in p3He

+by lin-

early polarized laser light of intensity I1 (in W/cm2) can becalculated as

2M

2π(S.I.) =

√2I1

ε0c

ea0

h|〈0LFKJM|

× d|q(L − 1)F ′K ′J ′M〉| (a.u.), (22)

where the transition matrix element (in atomic units) can bederived using the Wigner 3j and 6j symbols

|〈0LFKJM|d|q(L − 1)F ′K ′J ′M〉

=∣∣∣∣〈0L‖d‖q(L − 1)〉

(J 1 J ′

M 0 −M

)√(2J + 1)(2J ′ + 1)

×{K ′ J ′ 1

2

J K 1

}√(2K + 1)(2K ′ + 1)

{F ′ K ′ 1

2

K F 1

}

×√

(2F + 1)(2F ′ + 1)

{L′ F ′ 1

2

F L 1

}∣∣∣∣. (23)

062508-8

Page 9: Calculation of transition probabilities and ac Stark ...theor.jinr.ru/~korobov/papers/Two_photon_Stark_PRA10.pdf · MASAKI HORI AND VLADIMIR I. KOROBOV PHYSICAL REVIEW A 81, 062508

CALCULATION OF TRANSITION PROBABILITIES AND . . . PHYSICAL REVIEW A 81, 062508 (2010)

He3He3e− e−

e−

p

e− p e−

(n',L−1)

(n'',L−2)

F=L−1/2

(n',L−1)

(n'',L−2)

(b)

(a)

(n,L)

K=L

K=L−1

K=L

K=L+1

K=L−2

K=L−3

K=L−2

K=L−1

K=L

K=L−1

K=L−2

K=L−1

(n,L)

F=L−1/2

F=L+1/2

J=L

J=L−1

J=L+1

J=L

F=L−3/2

F=L−1/2

J=L−1

J=L−2

J=L

J=L−1

J=L−2

J=L−3

J=L−1

J=L−2

F=L−5/2

F=L−3/2

J=L+1/2

J=L−1/2

J=L−1/2

J=L−3/2

J=L+1/2

J=L−1/2

J=L+3/2

J=L+1/2

F=L+1/2

F=L−1/2

J=L−3/2

J=L−5/2

J=L−5/2

J=L−7/2

J=L−3/2

J=L−3/2

J=L+1/2

J=L−1/2

J=L−3/2

J=L−5/2

J=L−5/2

J=L−7/2

J=L−3/2

J=L−5/2

J=L−1/2

J=L−3/2

F=L−3/2

F=L−5/2

F=L−3/2

FIG. 8. (Color online) (a) Energy level diagram showing the hyperfine sublevels of the resonance parent, intermediate, and daughter statesinvolved in the two-photon transitions of p4He+. The spin orientation (Se,Sp) of each hyperfine sublevel is indicated by arrows. The fourstrongest two-photon transitions are indicated with solid arrows. The same figure in the case of p3He+ showing the eight strong lines betweenhyperfine levels (Se,Sh,Sp), see text.

Equation (23) only provides approximate values for the tran-sition matrix elements since F and K are not exact quantumnumbers of the three-body Hamiltonian. The results, however,agree with exact transition amplitudes within 1%, for all thepHe+ transitions of the type (n,L,M) → (n − 1,L − 1,M)studied here.

We numerically integrated Eq. (20) to simulate the antipro-tons depopulated by the two laser beams from the resonanceparent state to the daughter state via a two-photon transition.Prior to laser irradiation, the antiprotons are assumed touniformly populate the ∼70 M sublevels of the parentstate. The atoms follow a Maxwellian thermal distribution

062508-9

Page 10: Calculation of transition probabilities and ac Stark ...theor.jinr.ru/~korobov/papers/Two_photon_Stark_PRA10.pdf · MASAKI HORI AND VLADIMIR I. KOROBOV PHYSICAL REVIEW A 81, 062508

MASAKI HORI AND VLADIMIR I. KOROBOV PHYSICAL REVIEW A 81, 062508 (2010)

(a)

Laser energy ( mJ / cm2 )

An

tip

roto

n d

epo

pu

lati

on

eff

icie

ncy

0

0.1

0 0.1 0.2 0.3 0.4 0.5

(b)

Laser energy ( mJ / cm2 )

An

tip

roto

n d

epo

pu

lati

on

eff

icie

ncy

0

0.5

0 2 4 6 8 10

FIG. 9. Depopulation efficiency in the two-photon transition(n,L) = (36,34) → (34,32) of p4He+ at various intensities of thetwo laser beams. The virtual intermediate state is offset ωd/2π =−12 GHz away from state (35,33), so that the two laser frequenciescoincide with the hyperfine component (Se,Sp) = (↑↑) → (↑↑) ofthe resonance line.

of temperature T ∼ 10 K. The temporal profiles of the laserpulses are assumed to be roughly Gaussian with pulse lengths�t ∼ 100 ns [16].

Figures 9(a) and 9(b) show the efficiency ε of thelaser pulses depleting the population in the parent stateof the two-photon transition (36,34) → (34,32) of p4He

+

[i.e., ε = 1 if the laser induces all the antiprotons oc-cupying state (n,L) to annihilate and ε = 0 when nosuch annihilations occur]. The virtual intermediate stateis offset ωd/2π = −12 GHz away from state (35,33),so that ω1 + ω2 coincides with the hyperfine component(Se,Sp) = (↑↑) → (↑↑) of the resonance line. As the laserintensity is increased between p = 0 and 0.2 mJ/cm2,the ε value increases quadratically as expected for a two-photon process. It begins to saturate at p > 1 mJ/cm2 cor-responding to ε ∼ 0.3. Monte Carlo simulations indicate thatthe two-photon resonance signal of ε ∼ 0.3 would be strongenough for clear detection against the background caused byspontaneously annihilating antiprotons [83] with a signal-to-noise ratio of >5. Higher laser intensities would, of course,provide an even stronger signal, but power broadening effectsthen deteriorate the spectral resolution to several hundred MHzand so this should be avoided for high-precision spectroscopy.

The resonance profiles of the two-photon transitions(40,36) → (38,34) of the v = 3 cascade, (38,35) → (36,33)of v = 2, (36,34) → (34,32) of v = 1, and (33,32) → (31,30)of v = 0 in p4He

+at temperature T ∼ 10 K are shown in

Figs. 10(a)–10(d). These resonances have among the largesttransition amplitudes, and the Auger decay rates of thedaughter states are large γA = (2.5 − 4) × 108 s−1, which is anecessary condition to obtain a strong annihilation signal [83].The intensities of the two lasers are around p ∼ 1 mJ/cm2.The laser frequency ω1 is fixed to an offset �ωd/2π = −12GHz from the intermediate state, whereas ω2 was scannedbetween −0.9 and 0.9 GHz around the two-photon resonancedefined by ω1 + ω2. In each simulated profile, the positions ofthe four hyperfine lines are indicated with arrows together withthe corresponding spin orientations (Se,Sp). The ∼200 MHzlinewidth of these profiles are primarily caused by the largeAuger width of the daughter states and the residual Dopplerand power broadening.

The resonance (40,36) → (38,34) shows a two-peak struc-ture [Fig. 10(a)] with a frequency interval of ∼1.1 GHz, whicharises from the dominant spin-orbit interaction between Se andL. Each peak is a superposition of two hyperfine lines witha few tens of MHz spacing caused by a further interactionbetween the antiproton and electron spins. The asymmetricstructure of the profile of Fig. 10(a) is due to the fact thatthe 25-MHz spacing between the hyperfine lines (Se,Sp) =(↑↑) → (↑↑) and (↑↓) → (↑↓) are small compared to the75-MHz spacing between (↓↑) → (↓↑) and (↓↓) → (↓↓).The spacings between the hyperfine lines becomes graduallysmaller for lower-lying transitions involving states of smallern and v values [e.g., 0.8 and 0.5 GHz for (38,35) → (36,33)and (36,34) →(34,32)]. The hyperfine lines can no longerbe resolved for the lowest−n transition (33,32) → (31,30)[Fig. 12(d)]. The low transition probability (Table I) of thisresonance causes the small depopulation efficiency ε < 0.1seen here; laser intensities of p � 2 mJ/cm2 would be neededto produce a sufficient experimental signal.

We expect the two UV transitions (n,L) = (36,34) →(34,32) and (33,32) → (31,30) in p4He

+to yield the highest

signal-to-noise ratios in laser spectroscopy experiments. Thisis because the parent states (36,34) and (33,32) retain largeantiproton populations for long periods t = 3–10 µs followingpHe+ formation [40]. By comparison, cascade processesrapidly deplete the populations in higher n > 37 states within1–2 µs, and so the associated two-photon spectroscopysignals contain a large background due to the spontaneouslyannihilating pHe+ atoms [40,41].

Higher experimental precisions on νexp may be achievedby cooling the atoms to lower temperature and by inducingtwo-photon transitions between pairs of pHe+ states withmicrosecond-scale lifetimes. Figure 10(e) shows the resonance(37,35) → (35,33) of p4He

+at temperature T = 1.5 K. Both

parent and daughter states have lifetimes of τ ∼ 1 µs, andso its natural linewidth ∼200 kHz is two orders of magnitudesmaller than in the other resonances [Figs. 10(a)–10(d) studiedhere]. It is unfortunately difficult to measure this transitionexperimentally, as the present detection method requires thedaughter state to proceed rapidly to antiproton annihilation.Cooling the pHe+ atoms to such low temperatures may betechnically challenging.

The profiles of two p3He+

resonances which are expectedto yield the highest signal-to-noise ratios [40] (35,33) →(33,31) and (33,32) → (31,30) at temperature T ∼ 10 Kare shown in Figs. 11(a) and 11(b). The positions of the eight

062508-10

Page 11: Calculation of transition probabilities and ac Stark ...theor.jinr.ru/~korobov/papers/Two_photon_Stark_PRA10.pdf · MASAKI HORI AND VLADIMIR I. KOROBOV PHYSICAL REVIEW A 81, 062508

CALCULATION OF TRANSITION PROBABILITIES AND . . . PHYSICAL REVIEW A 81, 062508 (2010)

0

0.2

0.4

0.6

-0.5 0 0.50

0.2

0.4

0.6

-0.5 0 0.50

0.1

0.2

0.3

0.4

-0.5 0 0.5

0

0.02

0.04

0.06

0.08

0.1

0.12

-0.5 0 0.5

(a) (40,36)→(38,34)

(↓↑)

to

(↓↑

)(↓

↓) t

o (

↓↓)

(↑↑)

to

(↑↑

)(↑

↓) t

o (

↑↓)

Laser detuning (GHz)

Dep

op

ula

tio

n e

ffic

ien

cy(b) (38,35)

→(36,33)

(↓↑)

to

(↓↑

)(↓

↓) t

o (

↓↓)

(↑↑)

to

(↑↑

)

(↑↓)

to

(↑↓

)

Laser detuning (GHz)D

epo

pu

lati

on

eff

icie

ncy

(c) (36,34)→(34,32)

(↓↑)

to

(↓↑

)(↓

↓) t

o (

↓↓)

(↑↑)

to

(↑↑

)(↑

↓) t

o (

↑↓)

Laser detuning (GHz)

Dep

op

ula

tio

n e

ffic

ien

cy

(d)(33,32)→(31,30)

(↓↑)

to

(↓↑

)

(↓↓)

to

(↓↓

)

(↑↑)

to

(↑↑

)(↑

↓) t

o (

↑↓)

Laser detuning (GHz)

Dep

op

ula

tio

n e

ffic

ien

cy

(e) (37,35)→(35,33)

(↓↑)

to

(↓↑

)

(↓↓)

to

(↓↓

)

(↑↑)

to

(↑↑

)(↑

↓) t

o (

↑↓)

Laser detuning (GHz)

Dep

op

ula

tio

n e

ffic

ien

cy

0

0.05

0.1

0.15

-0.5 0 0.5

FIG. 10. Simulated resonance profiles of five two-photon transitions in p4He+ excited at laser intensities p ∼ 1 mJ/cm2. The virtualintermediate state was tuned �ωd/2π ∼ −12 GHz from a real state. Four profiles (a)–(d) involve a resonance daughter state with lifetimesτ < 10 ns against Auger decay. They are simulated assuming a temperature T ∼ 10 K of the atom. The narrow resonance (e) involves adaughter state of much longer (τ ∼ 1 µs) lifetime, and was simulated at T ∼ 1.5 K. The positions of the four hyperfine lines are indicated byarrows, together with the principal and angular momentum quantum numbers (n,L) and spin orientations (Se,Sp) → (S ′

e,S′p) of the parent and

daughter states.

0

0.1

0.2

0.3

0.4

-0.5 0 0.5

(a) (35,33)→(33,31)

(↑↑↓

) to

(↑↑

↓) (↓↑↓

) to

(↓↑

↓)

(↑↓↓

) to

(↑↓

↓)

(↓↓↓

) to

(↓↓

↓)

(↑↓↑

) to

(↑↓

↑)(↓

↓↑)

to (

↓↓↑)

(↑↑↑

) to

(↑↑

↑)

(↓↑↑

) to

(↓↑

↑)

Laser detuning (GHz)

Dep

op

ula

tio

n e

ffic

ien

cy

(b) (33,32)→(31,30)

(↑↑↓

) to

(↑↑

↓)

(↓↑↓

) to

(↓↑

↓)

(↑↓↓

) to

(↑↓

↓)

(↓↓↓

) to

(↓↓

↓)

(↑↓↑

) to

(↑↓

↑)(↓

↓↑)

to (

↓↓↑)

(↑↑↑

) to

(↑↑

↑)

(↓↑↑

) to

(↓↑

↑)

Laser detuning (GHz)

Dep

op

ula

tio

n e

ffic

ien

cy

0

0.05

0.1

0.15

-0.5 0 0.5

FIG. 11. Simulated resonance profiles of two-photon transitions (a) (n,L) = (35,33) → (33,31) and (b) (33,32) → (31,30) in p3He+

excited at laser intensities p ∼ 1 mJ/cm2 and temperature T ∼ 10 K. The virtual intermediate state was tuned �ωd/2π ∼ −12 GHz froma real state. The positions of the eight hyperfine lines are indicated by arrows, together with the principal and angular momentum quantumnumbers (n,L) and spin orientations (Se,Sh,Sp) → (S ′

e,S′h,S

′p) of the resonance parent and daughter states.

062508-11

Page 12: Calculation of transition probabilities and ac Stark ...theor.jinr.ru/~korobov/papers/Two_photon_Stark_PRA10.pdf · MASAKI HORI AND VLADIMIR I. KOROBOV PHYSICAL REVIEW A 81, 062508

MASAKI HORI AND VLADIMIR I. KOROBOV PHYSICAL REVIEW A 81, 062508 (2010)

(a)

Laser detuning (GHz)

An

tip

roto

n d

epo

pu

lati

on

eff

icie

ncy

0

0.05

0.1

0.15

0.2

0.25

-0.4 -0.3 -0.2 -0.1 -0 0.1 0.2 0.3 0.4

Laser intensity ratio I2 / I1

ac S

tark

sh

ift

∆ω

ac /

(M

Hz)

(b)

12 GHz

-12 GHz

-15

-10

-5

0

5

10

15

0.5 0.6 0.7 0.8 0.9 1

FIG. 12. Simulated profiles (a) of the resonance (n,L) = (36,34) → (34,32) in p4He+, with the frequency offset �ωd/2π ∼ −12 GHz andintensity ratio between the two lasers I2/I1 = 0.5 (dashed lines) and 1 (solid lines) for constant I1 ∼ 5 × 104 W/cm2. The atom is thermalizedat T ∼ 10 K. The ac Stark shift in the simulated profiles (b) for offsets �ωd/2π ∼ −12 GHz (solid circles) and 12 GHz (squares), as a functionof I2/I1.

hyperfine lines and their spin configurations (Se,Sh,Sp) areindicated by arrows. Due to the large number of partiallyoverlapping lines, it may be difficult to determine the νexp

values for p3He+

with a similar level of precision as inp4He

+. The problem would be especially acute in the case of

(33,32) → (31,30) [Fig. 11(b)] which contains eight sublineswithin a relatively small 0.6-GHz interval.

We finally use these numerical simulations to determinethe ac Stark shift under realistic experimental conditions.Figure 12(a) shows the profiles of the resonance (36,34) →(34,32) of p4He

+at temperature T ∼ 10 K and laser offsets

�ωd/2π = −12 GHz. They were calculated at two com-binations of the laser intensities I1 = 5 × 104 W/cm2 andI2 = 2.5 × 104 W/cm2 (broken lines) and I1 = I2 = 5 × 104

W/cm2 (solid lines). As I2/I1 is increased, the transitionfrequency shifts to larger values. In Fig. 12(b), the ac Starkshifts �ωac/2π determined from the simulated profiles ofFig. 12(a) at laser offsets �ωd/2π = −12 GHz are plottedusing filled circles. It increases linearly from −2 MHz atI2/I1 = 0.5, to 5 MHz at I2/I1 = 1. A similar plot for offset�ωd/2π = 12 GHz is shown using filled squares. The twocalculated sets of ac Stark shifts are of equal magnitude andopposite sign, the minimum occurring around I2/I1 ∼ 0.65.

IV. CONCLUSION

We conclude that two-photon transitions in pHe+ of thetype (n,L) → (n − 2,L − 2) can indeed be induced using

two counterpropagating nanosecond laser pulses of intensity∼1 mJ/cm2, for cases where the virtual intermediate state istuned within |�ωd/2π | = 10–20 GHz of the real state (n −1,L − 1). The spectral resolution of the measured resonancesshould increase by an order of magnitude or more compared toconventional single-photon spectroscopy. The ac Stark shiftsat these experimental conditions can reach several MHz ormore, but this can be minimized by carefully adjusting therelative intensities of the two laser beams. Any remainingshift can be canceled by comparing the resonance profilesmeasured at positive and negative offsets ±�ωd of the virtualintermediate state from the real state. In practice, this can bedone by using a frequency comb [84] to accurately controlthe frequencies ω1 and ω2 of the counterpropagating laserbeams. The UV two-photon transitions (36,34) → (34,32)and (33,32) → (31,30) in p4He

+, and (35,33) → (33,31) in

p3He+

are expected to yield particularly strong resonancesignals that can be precisely measured.

ACKNOWLEDGMENTS

We are indebted to R.S. Hayano. This work was supportedby the European Science Foundation and the DeutscheForschungsgemeinschaft (DFG), the Munich Advanced Pho-tonics (MAP) cluster of DFG, the Research Grants in theNatural Sciences of the Mitsubishi Foundation, and theInitiative Grant No. 08-02-00341 of the Russian Foundationfor Basic Research.

[1] R. S. Hayano, M. Hori, D. Horvath, and E. Widmann, Rep. Prog.Phys. 70, 1995 (2007).

[2] T. Yamazaki, N. Morita, R. S. Hayano, E. Widmann, andJ. Eades, Phys. Rep. 366, 183 (2002).

[3] N. Morita et al., Phys. Rev. Lett. 72, 1180 (1994).[4] F. E. Maas et al., Phys. Rev. A 52, 4266 (1995).[5] H. A. Torii et al., Phys. Rev. A 59, 223 (1999).[6] M. Hori et al., Phys. Rev. Lett. 87, 093401 (2001).

062508-12

Page 13: Calculation of transition probabilities and ac Stark ...theor.jinr.ru/~korobov/papers/Two_photon_Stark_PRA10.pdf · MASAKI HORI AND VLADIMIR I. KOROBOV PHYSICAL REVIEW A 81, 062508

CALCULATION OF TRANSITION PROBABILITIES AND . . . PHYSICAL REVIEW A 81, 062508 (2010)

[7] M. Hori et al., Phys. Rev. Lett. 91, 123401 (2003).[8] M. Hori et al., Phys. Rev. Lett. 96, 243401 (2006).[9] V. I. Korobov, Phys. Rev. A 61, 064503 (2000).

[10] V. I. Korobov, Phys. Rev. A 67, 062501 (2003).[11] V. I. Korobov, Phys. Rev. A 77, 042506 (2008).[12] Y. Kino, M. Kamimura, and H. Kudo, Nucl. Instrum. Methods

Phys. Res. B 214, 84 (2004).[13] S. Andersson, N. Elander, and E. Yarevsky, J. Phys. B 31, 625

(1998).[14] P. J. Mohr, B. N. Taylor, and D. B. Newell, Rev. Mod. Phys. 80,

633 (2008).[15] M. Hori, in Proceedings of Hydrogen Atom II: Precision Physics

of Simple Atomic Systems, Castiglione della Pescaia, Italy, 1–3June 2000 (Springer, Berlin, 2001).

[16] M. Hori and A. Dax, Opt. Lett. 34, 1273 (2009).[17] R. G. Beausoleil and T. W. Hansch, Phys. Rev. A 33, 1661

(1986).[18] J. C. Garreau, M. Allegrini, L. Julien, and F. Biraben, J. Phys.

(France) 51, 2263 (1990).[19] M. Fischer et al., Phys. Rev. Lett. 92, 230802 (2004).[20] M. Haas et al., Phys. Rev. A 73, 052501 (2006).[21] M. Haas, U. D. Jentschura, and C. H. Keitel, Am. J. Phys. 74,

77 (2006).[22] N. Kolachevsky, A. Matveev, J. Alnis, C. G. Parthey, S. G.

Karshenboim, and T. W. Hansch, Phys. Rev. Lett. 102, 213002(2009).

[23] G. Gabrielse et al., Phys. Rev. Lett. 100, 113001 (2008).[24] G. Andresen et al., Phys. Rev. Lett. 98, 023402 (2007).[25] S. Hannemann, E. J. Salumibides, S. Witte, R. T. Zinkstok, E.-J.

van Duijn, K. S. E. Eikema, and W. Ubachs, Phys. Rev. A 74,062514 (2006).

[26] L. Hilico, N. Billy, B. Gremaud, and D. Delande, J. Phys. B 34,491 (2001).

[27] K. S. E. Eikema, W. Ubachs, W. Vassen, and W. Hogervorst,Phys. Rev. Lett. 76, 1216 (1996).

[28] K. S. E. Eikema, W. Ubachs, W. Vassen, and W. Hogervorst,Phys. Rev. A 55, 1866 (1997).

[29] F. Minardi, G. Bianchini, P. C. Pastor, G. Giusfredi, F. S. Pavone,and M. Inguscio, Phys. Rev. Lett. 82, 1112 (1999).

[30] S. D. Bergeson, K. G. H. Baldwin, T. B. Lucatorto, T. J. McIlrath,C. H. Cheng, and E. E. Eyler, J. Opt. Soc. Am. B 17, 1599 (2000).

[31] V. Yakhontov and K. Jungmann, Z. Phys. D 38, 141 (1996).[32] V. Yakhontov, R. Santra, and K. Jungmann, J. Phys. B 32, 1615

(1999).[33] V. Meyer et al., Phys. Rev. Lett. 84, 1136 (2000).[34] H. Knudsen et al., Phys. Rev. Lett. 101, 043201 (2008).[35] M. Foster, J. Colgan, and M. S. Pindzola, Phys. Rev. Lett. 100,

033201 (2008).[36] A. Luhr and A. Saenz, Phys. Rev. A 79, 042901 (2009).[37] N. Henkel, M. Keim, H. J. Ludde, and T. Kirchner, Phys. Rev.

A 80, 032704 (2009).[38] M. McGovern, D. Assafrao, J. R. Mohallem, C. T. Whelan, and

H. R. J. Walters, Phys. Rev. A 79, 042707 (2009).[39] I. F. Barna, K. Tokesi, L. Gulyas, and J. Burgdorfer, Radiat.

Phys. Chem. 76, 495 (2007).[40] M. Hori et al., Phys. Rev. Lett. 89, 093401 (2002).[41] M. Hori et al., Phys. Rev. A 70, 012504 (2004).[42] J. S. Briggs, P. T. Greenland, and E. A. Solov’ev, Hyperfine

Interact. 119, 235 (1999).

[43] J. S. Cohen, Rep. Prog. Phys. 67, 1769 (2004).[44] M. Hesse, A. T. Le, and C. D. Lin, Phys. Rev. A 69, 052712

(2004).[45] K. Tokesi, B. Juhasz, and J. Burgdorfer, J. Phys. B 38, S401

(2005).[46] S. Yu. Ovchinnikov and J. H. Macek, Phys. Rev. A 71, 052717

(2005).[47] J. Revai and N. Shevchenko, Eur. Phys. J. D 37, 83 (2006).[48] X. M. Tong, K. Hino, and N. Toshima, Phys. Rev. Lett. 101,

163201 (2008).[49] M. Genkin and E. Lindroth, Eur. Phys. J. D 51, 205 (2009).[50] M. Hori et al., Phys. Rev. A 57, 1698 (1998); 58, 1612 (1998).[51] J. E. Russell, Phys. Rev. A 65, 032509 (2002).[52] B. D. Obreshkov, D. D. Bakalov, B. Lepetit, and K. Szalewicz,

Phys. Rev. A 69, 042701 (2004).[53] M. Hori, R. S. Hayano, E. Widmann, and H. A. Torii, Opt. Lett.

28, 2479 (2003).[54] H. Yamaguchi et al., Phys. Rev. A 66, 022504 (2002).[55] H. Yamaguchi et al., Phys. Rev. A 70, 012501 (2004).[56] V. I. Korobov and I. Shimamura, Phys. Rev. A 56, 4587 (1997).[57] J. Revai and A. T. Kruppa, Phys. Rev. A 57, 174 (1998).[58] O. I. Kartavtsev, D. E. Monakhov, and S. I. Fedotov, Phys. Rev.

A 61, 062507 (2000).[59] M. Hori et al., Phys. Rev. Lett. 94, 063401 (2005).[60] K. Sakimoto, Phys. Rev. A 76, 042513 (2007).[61] G. Ya. Korenman and S. N. Yudin, J. Phys. Conf. Series 88,

012060 (2007).[62] K. Sakimoto, Phys. Rev. A 79, 042508 (2009).[63] J. E. Bjorkholm and P. F. Liao, Phys. Rev. Lett. 33, 128

(1974).[64] R. Salomaa and S. Stenholm, J. Phys. B 8, 1795 (1975).[65] R. Salomaa and S. Stenholm, J. Phys. B 9, 1221 (1976).[66] R. Salomaa, J. Phys. B 10, 3005 (1977).[67] V. G. Bordo and H. G. Rubahn, Phys. Rev. A 60, 1538 (1999).[68] C. Fort, M. Inguscio, P. Raspollini, F. Baldes, and A. Sasso,

Appl. Phys. B 61, 467 (1995).[69] C. Wei, D. Suter, A. S. M. Windsor, and N. B. Manson, Phys.

Rev. A 58, 2310 (1998).[70] T. Pask et al., Phys. Lett. B 678, 55 (2009).[71] D. Bakalov and V. I. Korobov, Phys. Rev. A 57, 1662 (1998).[72] N. Yamanaka, Y. Kino, H. Kudo, and M. Kamimura, Phys. Rev.

A 63, 012518 (2000).[73] Y. Kino, N. Yamanaka, M. Kamimura, and H. Kudo, Hyperfine

Interact. 146/147, 331 (2003).[74] V. I. Korobov, Phys. Rev. A 73, 022509 (2006).[75] D. Bakalov and E. Widmann, Phys. Rev. A 76, 012512 (2007).[76] A. Khadjavi, A. Lurio, and W. Happer, Phys. Rev. 167, 128

(1968).[77] K. T. Chung, J. Phys. B 25, 4711 (1992).[78] K. D. Bonin and M. A. Kadar-Kallen, Int. J. Mod. Phys. B 8,

3313 (1994).[79] A. K. Bhatia and R. J. Drachman, J. Phys. B 27, 1299 (1994).[80] M. Rerat and C. Pouchan, Phys. Rev. A 49, 829 (1994).[81] M. Masili and A. F. Starace, Phys. Rev. A 68, 012508 (2003).[82] V. I. Korobov, Phys. Rev. A 54, R1749 (1996).[83] M. Hori, K. Yamashita, R. S. Hayano, and T. Yamazaki, Nucl.

Instrum. Methods Phys. Res. A 496, 102 (2003).[84] Th. Udem, R. Holzwarth, and T. W. Hansch, Nature (London)

416, 233 (2002).

062508-13