6
CHEMISTRY Multidimensional photon correlation spectroscopy of cavity polaritons Konstantin E. Dorfman a,1 and Shaul Mukamel b,c,1 a State Key Laboratory of Precision Spectroscopy, East China Normal University, Shanghai 200062, China; b Department of Chemistry, University of California, Irvine, CA 92697; and c Department of Physics and Astronomy, University of California, Irvine, CA 92697 Contributed by Shaul Mukamel, December 27, 2017 (sent for review November 7, 2017; reviewed by Steven Cundiff and Angel Rubio) The strong coupling of atoms and molecules to radiation field modes in optical cavities creates dressed matter/field states known as polaritons with controllable dynamical and energy transfer properties. We propose a multidimensional optical spectroscopy technique for monitoring polariton dynamics. The response of a two-level atom to the time-dependent coupling to a single- cavity mode is monitored through time-and-frequency–resolved single-photon coincidence measurements of spontaneous emis- sion. Polariton population and coherence dynamics and its varia- tion with cavity photon number and controlled by gating parame- ters are predicted by solving the Jaynes–Cummings model. multidimensional spectroscopy | photon correlation | cavity polaritons C avity quantum electrodynamics (QED) provides a power- ful tool for studying quantum effects in matter (1–6). Due to strong coupling between electronic and nuclear degrees of freedom, molecular systems can undergo nonadiabatic dynam- ics, which is hard to detect. The nonadiabatic dynamics can be manipulated (7) when a molecule is coupled to a localized cav- ity mode. Earlier studies in atomic systems showed that the cavity photons can enhance cooperative signals, such as super- radiance and subradiance (8, 9). The description of these phe- nomena is based on the joint photon–matter states known as polaritons (10, 11). Cavity polaritons have been applied to trap- ping and cooling of atoms (12) and prescribe a new recipe for cooling molecules (13, 14). Cavity effects can provide a tool to probe larger molecules (15, 16), where various many-body quan- tum effects play an important role. For instance, the polariton– polariton interaction strength can be directly probed in a high- quality microcavity (17). Strong coupling in cavity QED has been recently demonstrated for organic molecules (18–21) and pho- tosynthetic light harvesting (22). Polaritons have been further investigated in chromophore aggregates (11) arising from elec- tronic transitions (10) as well as from vibrational transitions (23), which in turn, allow the manipulation of chemical reaction rates and outcomes (24–27). Cavity polariton dynamics can be investigated by nonlinear spectroscopy. IR and Raman spectroscopies have been recently used to show the enhancement of the spectra of vibrational polaritons in molecular aggregates (11, 28–31). More elaborate two-dimensional spectroscopic measurements have further pro- vided experimental demonstration for multiexciton correlation effects (32). Coherent multidimensional spectroscopy can reveal correlations of matter dynamics during several time intervals controlled by sequences of short pulses to reveal material infor- mation (33, 34) by a coherent measurement of a signal optical field. Such correlation plots carry qualitatively higher levels of information than single-interval (1D) techniques. A recent theo- retical study of vibrational polaritons using coherent 2D IR spec- troscopy (35) has been reported. In this paper, we propose to study polariton dynamics using a different class of incoherent multidimensional signals. Unlike coherent multidimensional techniques, which is based on care- fully timed laser pulse sequences, incoherent techniques detect spontaneously emitted light, and the control knobs of such sig- nals are based on single-photon gated detection. Time-and- frequency (TF)–gated N -photon measurement provides a 2N - dimensional parameter ωj tj space. An adequate microscopic description where joint matter and field information could be retrieved by a proper description of the detection process is required for, e.g., single-photon spectroscopy of single molecules (36–38). These photon counting techniques performed on bulk ensembles or at the single-molecule level offer unique windows into molecular events and relaxation processes that are comple- mentary to coherent multidimensional techniques (34). The pro- posed incoherent photon coincidence counting measurements (5, 39) can monitor the joint system–cavity mode state as well as the cavity mode statistics. The independent control of the TF photon gating parameters can capture detailed features of polariton dynamics. We consider a two-level system strongly coupled to the quantized cavity modes. The strong coupling is attributed to the enhanced density of radiation states inside the cavity governed by the Purcell effect (1), which results in the strong enhancement of the photon emission into the cavity modes. The joint atom plus photon states (polaritons) can be described by the Jaynes–Cummings (JC) model (2), which is a pillar of quantum optics. In addition, strong coupling requires large absorption oscillator strength, narrow exciton absorption line width, and small cavity losses into leaky modes. It is, there- fore, expected that dissipation does not affect the ladder dynam- ics. Furthermore, for typical molecular polaritons, the dynamics occurs on a picosecond timescale, whereas the cavity linewidth is usually in the nanosecond range. The quantum nature of the field has been shown to result in revival of damped Rabi oscillations, Significance We propose a spectroscopic technique that can track the time- dependent state of a dressed molecule in an optical cavity (polariton) by measuring coincidence of two emitted photons. The proposed technique offers an independent control of the spectral and temporal resolution; single-photon detection allows for low-intensity measurements, which do not disturb the state of the cavity field, and time-dependent atom/cavity coupling provides a control tool. Tracking the evolution of the polariton states with time-dependent atom/cavity coupling should be of interest in photochemistry and photobiology and could improve fundamental understanding of many physical processes in strongly coupled atom/radiation states. Possible applications include chemical sensors and quantum informa- tion processing. Author contributions: K.E.D. and S.M. designed research; K.E.D. performed calculations and analyzed the data; and K.E.D. and S.M. wrote the paper. Reviewers: S.C., University of Michigan; and A.R., Max Planck Institute for the Structure & Dynamics. The authors declare no conflict of interest. Published under the PNAS license. 1 To whom correspondence may be addressed. Email: [email protected] or [email protected]. This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10. 1073/pnas.1719443115/-/DCSupplemental. www.pnas.org/cgi/doi/10.1073/pnas.1719443115 PNAS | February 13, 2018 | vol. 115 | no. 7 | 1451–1456 Downloaded by guest on March 10, 2020

} of cavity polaritons · cavity mode is monitored through time-and-frequency–resolved single-photon coincidence measurements of spontaneous emis-sion. Polariton population and

  • Upload
    others

  • View
    7

  • Download
    0

Embed Size (px)

Citation preview

Page 1: } of cavity polaritons · cavity mode is monitored through time-and-frequency–resolved single-photon coincidence measurements of spontaneous emis-sion. Polariton population and

CHEM

ISTR

Y

Multidimensional photon correlation spectroscopyof cavity polaritonsKonstantin E. Dorfmana,1 and Shaul Mukamelb,c,1

aState Key Laboratory of Precision Spectroscopy, East China Normal University, Shanghai 200062, China; bDepartment of Chemistry, University of California,Irvine, CA 92697; and cDepartment of Physics and Astronomy, University of California, Irvine, CA 92697

Contributed by Shaul Mukamel, December 27, 2017 (sent for review November 7, 2017; reviewed by Steven Cundiff and Angel Rubio)

The strong coupling of atoms and molecules to radiation fieldmodes in optical cavities creates dressed matter/field states knownas polaritons with controllable dynamical and energy transferproperties. We propose a multidimensional optical spectroscopytechnique for monitoring polariton dynamics. The response ofa two-level atom to the time-dependent coupling to a single-cavity mode is monitored through time-and-frequency–resolvedsingle-photon coincidence measurements of spontaneous emis-sion. Polariton population and coherence dynamics and its varia-tion with cavity photon number and controlled by gating parame-ters are predicted by solving the Jaynes–Cummings model.

multidimensional spectroscopy | photon correlation | cavity polaritons

Cavity quantum electrodynamics (QED) provides a power-ful tool for studying quantum effects in matter (1–6). Due

to strong coupling between electronic and nuclear degrees offreedom, molecular systems can undergo nonadiabatic dynam-ics, which is hard to detect. The nonadiabatic dynamics can bemanipulated (7) when a molecule is coupled to a localized cav-ity mode. Earlier studies in atomic systems showed that thecavity photons can enhance cooperative signals, such as super-radiance and subradiance (8, 9). The description of these phe-nomena is based on the joint photon–matter states known aspolaritons (10, 11). Cavity polaritons have been applied to trap-ping and cooling of atoms (12) and prescribe a new recipe forcooling molecules (13, 14). Cavity effects can provide a tool toprobe larger molecules (15, 16), where various many-body quan-tum effects play an important role. For instance, the polariton–polariton interaction strength can be directly probed in a high-quality microcavity (17). Strong coupling in cavity QED has beenrecently demonstrated for organic molecules (18–21) and pho-tosynthetic light harvesting (22). Polaritons have been furtherinvestigated in chromophore aggregates (11) arising from elec-tronic transitions (10) as well as from vibrational transitions (23),which in turn, allow the manipulation of chemical reaction ratesand outcomes (24–27).

Cavity polariton dynamics can be investigated by nonlinearspectroscopy. IR and Raman spectroscopies have been recentlyused to show the enhancement of the spectra of vibrationalpolaritons in molecular aggregates (11, 28–31). More elaboratetwo-dimensional spectroscopic measurements have further pro-vided experimental demonstration for multiexciton correlationeffects (32). Coherent multidimensional spectroscopy can revealcorrelations of matter dynamics during several time intervalscontrolled by sequences of short pulses to reveal material infor-mation (33, 34) by a coherent measurement of a signal opticalfield. Such correlation plots carry qualitatively higher levels ofinformation than single-interval (1D) techniques. A recent theo-retical study of vibrational polaritons using coherent 2D IR spec-troscopy (35) has been reported.

In this paper, we propose to study polariton dynamics usinga different class of incoherent multidimensional signals. Unlikecoherent multidimensional techniques, which is based on care-fully timed laser pulse sequences, incoherent techniques detectspontaneously emitted light, and the control knobs of such sig-

nals are based on single-photon gated detection. Time-and-frequency (TF)–gated N -photon measurement provides a 2N -dimensional parameter ωj tj space. An adequate microscopicdescription where joint matter and field information could beretrieved by a proper description of the detection process isrequired for, e.g., single-photon spectroscopy of single molecules(36–38). These photon counting techniques performed on bulkensembles or at the single-molecule level offer unique windowsinto molecular events and relaxation processes that are comple-mentary to coherent multidimensional techniques (34). The pro-posed incoherent photon coincidence counting measurements(5, 39) can monitor the joint system–cavity mode state as wellas the cavity mode statistics. The independent control of theTF photon gating parameters can capture detailed features ofpolariton dynamics. We consider a two-level system stronglycoupled to the quantized cavity modes. The strong couplingis attributed to the enhanced density of radiation states insidethe cavity governed by the Purcell effect (1), which results inthe strong enhancement of the photon emission into the cavitymodes. The joint atom plus photon states (polaritons) can bedescribed by the Jaynes–Cummings (JC) model (2), which is apillar of quantum optics. In addition, strong coupling requireslarge absorption oscillator strength, narrow exciton absorptionline width, and small cavity losses into leaky modes. It is, there-fore, expected that dissipation does not affect the ladder dynam-ics. Furthermore, for typical molecular polaritons, the dynamicsoccurs on a picosecond timescale, whereas the cavity linewidth isusually in the nanosecond range. The quantum nature of the fieldhas been shown to result in revival of damped Rabi oscillations,

Significance

We propose a spectroscopic technique that can track the time-dependent state of a dressed molecule in an optical cavity(polariton) by measuring coincidence of two emitted photons.The proposed technique offers an independent control ofthe spectral and temporal resolution; single-photon detectionallows for low-intensity measurements, which do not disturbthe state of the cavity field, and time-dependent atom/cavitycoupling provides a control tool. Tracking the evolution of thepolariton states with time-dependent atom/cavity couplingshould be of interest in photochemistry and photobiology andcould improve fundamental understanding of many physicalprocesses in strongly coupled atom/radiation states. Possibleapplications include chemical sensors and quantum informa-tion processing.

Author contributions: K.E.D. and S.M. designed research; K.E.D. performed calculationsand analyzed the data; and K.E.D. and S.M. wrote the paper.

Reviewers: S.C., University of Michigan; and A.R., Max Planck Institute for the Structure& Dynamics.

The authors declare no conflict of interest.

Published under the PNAS license.1 To whom correspondence may be addressed. Email: [email protected] [email protected].

This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1719443115/-/DCSupplemental.

www.pnas.org/cgi/doi/10.1073/pnas.1719443115 PNAS | February 13, 2018 | vol. 115 | no. 7 | 1451–1456

Dow

nloa

ded

by g

uest

on

Mar

ch 1

0, 2

020

Page 2: } of cavity polaritons · cavity mode is monitored through time-and-frequency–resolved single-photon coincidence measurements of spontaneous emis-sion. Polariton population and

which is missed by a classical field (40). We shall use a gener-alized JC model, where the system–cavity field coupling is timedependent (41–43). To achieve temporal modulation, cavity mir-ror must be manipulated on a similar timescale to the molecu-lar dynamics that needs to be controlled. It is hard to achievethis by mechanical devices. However, terahertz scanning tunnel-ing microscope (THz-STM) techniques may provide necessarymodulation speed down to picosecond timescale (44).

Analytical solutions for the time evolution of the populationinversion for the time-dependent JC model can be derived forspecific time profiles of the system–cavity coupling (41, 42, 45).The gated number of photons spontaneously emitted into non-cavity modes and photon coincidence signals for this analyticallysolvable time-dependent JC model are given by two-point andfour-point correlation functions of the dipole operator, respec-tively. We shall use a compact time ordered superoperator for-malism in these calculations (46–49). These correlation functionsand the corresponding Green’s functions that describe dynamicsin the dressed field–matter (polariton) space will be calculated.The proposed signals provide a unique observation window forthe population and coherence polariton dynamics.

The spontaneously emitted photons will be detected by a TF–resolved photon gating obtained by consecutive spectral andtemporal gates with bandwidths Γω and ΓT , respectively. Opti-mization of the temporal and spectral resolutions is requiredwhen the field–matter coupling variation is faster than theinverse magnitude of the change of the coupling. We comparethis gating with a simpler commonly used “physical spectrum”(PS) (50), which only uses a spectral gate characterized by a sin-gle parameter Γps that controls both spectral Γps and temporalresolution 1/Γps . We find that the TF gating can resolve a varietyof dynamical features that are missed by the PS.

Evolution of Polaritons in Time-Modulated CavitiesThe two-level molecular system can be described by the Paulimatrices, which satisfy σiσj = δij + iεijkσk , where i , j , k = x , y , z .We denote the eigenstates of σz by | ↑〉 and | ↓〉 with eigenval-ues 1 and −1; the operators σ±= 1

2(σx ± iσy) represent dipole

raising V †=µ∗σ+ and lowering V =µσ− operators, where µis the transition dipole. Using this notation, the time-dependentJC model Hamiltonian reads

HJC =ωA+A−+ω0

2σz +λ(t)(A+σ−+A−σ+), [1]

where ω0 is the transition frequency, λ(t) is the time-modulatedsystem–cavity coupling, and A+ (A−) is a creation (annihila-tion) operator for the cavity mode. The first two terms repre-sent the cavity field and the molecule, respectively, while thethird term corresponds to the interaction between molecule andcavity field. The matter plus field dressed (polariton) state ener-gies that form the JC ladder obtained by diagonalizing Eq. 1

En±=ω(n + 1/2)±√

(ω0−ω)2/4 +λ2(t)(n + 1) are shown inFig. 1. We adopt the following coupling profile:

λ(t) =λ0sech[(t −T )/(2τ)]. [2]

This represents pulsed modulation of duration 2τ centered attime T with magnitude λ2

0 =ω2

0 |µ|2

2ωε0V~, where V is the cavity vol-

ume. The sech profile allows for exact analytical solution for thetime evolution using special functions (SI Text). Our approach,however, is general and can treat an arbitrary time profile.

In addition to the strong coupling to the single-longitudinalcavity mode, the atom is weakly coupled to the vacuum modesdescribed by Hvac =V

∑k gka

†k +H .c., where ak is a field oper-

ator for vacuum modes that is responsible for the spontaneousemission into noncavity modes. The spontaneous emission rate

noncavity modes

cavity mode

t1, ω1 t2, ω2

photon numberdetector

Fig. 1. (Top) Atom in the cavity emitting photons into the longitudinal cav-ity mode as well as noncavity vertical modes. Two TF-resolved detectors out-side of the cavity register noncavity modes photon coincidence events. Onephoton number-resolving detector on the main optical axis measures pho-ton number of the cavity mode photons. (Middle Left) Fock states. (Mid-dle Right) Time evolution of the dressed atom plus cavity (polariton) statesof the JC subject to the coupling Eq. 2. Vertical black dashed lines repre-sent transition between different ladder states captured by spectroscopicmeasurements. (Bottom) Time evolution of the atom–cavity mode couplingλ(t) (black line) and population inversion 〈σz(t)〉′ Eq. 3 (red line) for anatom initially in the excited state with cavity mode being initially in thevacuum state.

into the cavity mode (vertical lines in Fig. 1) at resonance ω=

ωc =ω0 is γcav = 2Q|µ|2~ε0V

(3), where Q =ω0/2κ is the quality fac-tor of a mode at frequency ωc with cavity loss rate κ. The typi-cal solid angle extended by the cavity mirrors is very small, andtherefore, the atomic decay rate governed by Hvac into noncav-ity modes is then comparable with the rate of atomic decay ratein free space γ0 =

ω30 |µ|

2

3~ε0πc3 . The Purcell factor f = γcavγ0

= 3QV

2πc3

ω30

1452 | www.pnas.org/cgi/doi/10.1073/pnas.1719443115 Dorfman and Mukamel

Dow

nloa

ded

by g

uest

on

Mar

ch 1

0, 2

020

Page 3: } of cavity polaritons · cavity mode is monitored through time-and-frequency–resolved single-photon coincidence measurements of spontaneous emis-sion. Polariton population and

CHEM

ISTR

Y

(1, 3) is much greater than one, and the cavity leads to a strongenhancement of the atom’s ability to emit photons into the cavitymode. Details are in SI Text. We use perturbation theory in Hvac

and trace over the vacuum field. We shall calculate various pho-ton counting signals by placing the detectors off the cavity axis.These signals are governed by multipoint correlation functions ofthe dipole operators, and the corresponding Green’s functionsare determined by an exact solution of the equation of motiongoverned by HJC alone.

Gated Photon Counting SignalsAtomic Inversion. We first present the atomic inversion (popula-tion difference of the two-level system) 〈σz (t)〉′= 〈ψ0|U (t)F0

U †(t)|ψ0〉 given in Eq. S51. Assuming that, initially, the atom isin the excited state (wn = 1, vn = 0) and the cavity mode in thevacuum state n = 0, the inversion

〈σz (t)〉′= 1− 2|F1(t)|2 [3]

undergoes oscillation between ±1 as seen in Fig. 1. F dependson the atom–field interaction time τ . We note that, althoughinitially, the cavity is empty (no photons), the atom will Rabiflip between the ground and excited states during the interactiontime due to the fact that continuum of modes of the free spacemodes is replaced by a single cavity mode. This results in Rabioscillations rather than dissipative spontaneous emission.

Gated Photon Counting Signals. In the following, we denote ageneral Nth-order correlation measurement as photon count-ing. Gated photon number and gated photon coincidence cor-respond to g(1) and g(2) measurements, respectively, where eachphoton has its own detector. We thus calculated the gated pho-ton number and photon coincidence signals using the formalismof ref. 48. TF-resolved photon number is detected by placing asequence of temporal and spectral filters in front of the bucketphoton detector, whereas the photon coincidence is detected bysimultaneous monitoring of a pair of the gated photons, whichis a single-photon version of the intensity–intensity correlationmeasurement. Assuming Lorentzian gating (Eq. S32), we obtain(for the TF-resolved photon number signal)

n(t ,ω) =

∫ ∞0

dt ′1

∫ ∞0

dt ′′1 D(ω; t ′1, t ′′1 )〈V †(t − t ′′1 )V (t − t ′1)〉′.

[4]

The coincidence counting signal is similarly given by

g(2)(t1,ω1; t2,ω2) =

∫ ∞0

dt ′1

∫ ∞0

dt ′′1

∫ ∞0

dt ′2

∫ ∞0

dt ′′2

×D(ω1; t ′1, t ′′1 )D(ω2; t ′2, t ′′2 )〈V †(t − t ′′1 )

×V †(t − t ′′2 )V (t − t ′2)V (t − t ′1)〉′. [5]

For the TF gated signals DTF (ω, t ′, t ′′) = θ(t ′′− t ′)F ∗Γ+(t ′′,ω)

FΓ−(t ′,ω), where exponential gate FΓ(t ′,ω) = e [i(ω−ω0)−Γ]t′.The gating bandwidth in DTF is given by a combination of tem-poral and spectral gates Γ±= ΓT ±Γω . Details of the derivationof the signals via matter correlation functions defined by Eqs.S45 and S48 are given in SI Text.

Results and DiscussionUsing Eqs. 4 and 5, we have simulated the photon number n

and the photon coincidence rate g(2); the time-dependent cou-pling profile given by Eq. 2 is shown in Fig. 1 for the param-eters given in Materials and Methods. Fig. 2 depicts the time–frequency dependence of the TF gated photon number. Considerfirst a fast variation τ , such that ∆t∆ω' τ2Ω1 = 1, where 2Ω1 is

0.0

0.5

1.0

0.0

0.5

1.0

A B

C D

-1/4

0

-1/8

1/8

1/4

(t-T)/τ

-4 0 4(ω - ω Ω

2-2-6 6 -4 0 4(ω - ω Ω

2-2-6 6

-4 0 4(ω - ω Ω

2-2-6 6-4 0 4(ω - ω Ω

2-2-6 6-1/4

0

-1/8

1/8

1/4

(t-T)/τ

Fig. 2. (A and B) A 2D depiction of TF gated photon number nTF (t,ω) [S60]using the gating parameters Γω = 2λ0 and ΓT = 2.5λ0 for the fast modu-lation with τ−1 = 2.2λ0 (A) and Γω = 0.8λ0 and ΓT = 0.95λ0 along withslow modulation τ−1 = 0.6λ0 (B). (C and D) Same coupling duration as in Aand B but for PS gated photon number nPS(t,ω) [S59] for gated bandwidthsΓps = 2λ0 (C) and Γps = 0.8λ0 (D).

a spectral interval between neighboring lines of the first-orderJC ladder (e.g., at ω=ω0 and ω0± 2Ω1, where Ωn =παn/4τ).The photon number signal shown in Fig. 2A now shows a domi-nant peak at ω=ω0 and two strong side peaks at ω=ω0± 4Ω1,which evolve for the entire range of t . We also see weak peaksat ω=ω0± 2Ω1. For a slow modulation ∆t∆ω= 3.7 due to theextra time gate, the TF signal yields no time resolution and veryweak side peaks at ω=ω0± 2Ω1 as shown in Fig. 2B. The sidepeaks signify the coherence origin of the photon coming fromthe superposition of the dressed atom–cavity states, which canonly be observed during the time modulation of the coupling.

The photon coincidence signals are shown in Fig. 3 A–D.For rapid coupling variation corresponding to Fig. 2A, the coin-cidence counting depicted in Fig. 3A contains side peaks atω1,2 =ω0± 2Ω1, ω0± 4Ω1. These are well-resolved and form agrid similar to that reported by del Valle and coworkers (51–53). For gating times t1 =T , t2 =T − 20τ , the signal shows highspectral resolution with well-pronounced peaks at ω2 =ω0± 2Ω1

and a single peak for ω1 =ω0. In addition, it contains peaks forω1 =ω0± 2Ω1 and ω0± 4Ω1 as depicted in Fig. 3B. At t1 =T +20τ , t2 =T − 20τ , Fig. 3C is similar to Fig. 3A. Finally, for t1 =T + 20τ , t2 =T , Fig. 3D shows a full grid of well-pronouncedresonances with three prominent peaks for ω1 =ω0,ω0± 2Ω1

and ω2 =ω0 and a set of weaker peaks for ω2 =ω0± 2Ω1 andω0± 4Ω1.

We next compare the commonly used TF gated resultsof Figs. 2 A and B and 3 A–D with the simpler PS gat-ing (50, 51), which uses the gating function DPS (ω; t ′, t ′′) =F ∗Γps

(t ′′,ω)FΓps (t′,ω). The PS photon number signal (4) shown

in Fig. 2C has a single dominant peak at ω=ω0 and twovery weak peaks at ω=ω0± 4Ω1 visible only at t 'T . This isdue to the limited resolution allowed by the model (∆t∆ω=2). Higher-order resonances at ω=ω0± 2Ω1 and ω=ω0±6Ω1 are not resolved by this gate. For the slow process

Dorfman and Mukamel PNAS | February 13, 2018 | vol. 115 | no. 7 | 1453

Dow

nloa

ded

by g

uest

on

Mar

ch 1

0, 2

020

Page 4: } of cavity polaritons · cavity mode is monitored through time-and-frequency–resolved single-photon coincidence measurements of spontaneous emis-sion. Polariton population and

A B C D

E F G H

0.0

0.5

1.0

0.0

0.5

1.0

-4-22

4

0(ω-ω)/Ω

-4 0 2(ω - ω )/Ω-2 4 -4 0 2(ω - ω )/Ω-2 4 -4 0 2(ω - ω )/Ω-2 4 -4 0 2(ω - ω )/Ω-2 4

- 0 2ω - ω Ω-2 - 0 2

ω - ω Ω-2 - 0 2

ω - ω Ω-2 - 0 2

ω - ω Ω-2

-4-22

4

0(ω-ω)/Ω

Fig. 3. (A–D) TF gated photon coincidence counting g(2)TF (t1,ω1; t2,ω2) [S62] vs. ω1 and ω2 using gating parameters Γω1 = Γω2 = 2λ0, ΓT1 = ΓT2 = 2.5λ0

for t1 = T − 20τ , t2 = T + 20τ (A); t1 = T , t2 = T − 20τ (B); t1 = T + 20τ , t2 = T − 20τ (C); and t1 = T + 20τ , t2 = T (D). (E–H) Same but for PS gatingg(2)

PS (t1,ω1; t2,ω2) [S61] using gated bandwidth Γps1 = Γps2 = 2λ0 cm−1.

depicted in Fig. 2D, the stationary PS gate might have higherspectral resolution, allowing us to visualize the strong side peaksat ω=ω0± 2Ω1. Therefore, for slow cavity modulation, the tem-poral gate is not necessary, and the PS is adequate. However,rapid modulation requires separately controlled time gate when∆t∆ω≤ 1, where PS fails and TF is required. Note that the keyparameter is not an absolute value of the modulation depth λ0

and duration τ but is their product for a given cavity photon num-ber via τΩ1∼ τλ0

√n + 1, which governs the temporal and spec-

tral resolution via the product ∆t∆ω.Turning to coincidence counting, we note that the PS Eq. 5

has a low temporal and spectral resolution and misses some spec-tral features. For instance, the JC ladder of various peaks cannotbe observed in Fig. 3 E and G for n = 1, t1 =T − 20τ , t2 =T +20τ and t1 =T + 20τ , t2 =T − 20τ , respectively, other than thezero-order peak at ω1 =ω2 =ω0. For t1 =T , t2 =T − 20τ , thePS signal shows two weak peaks at ω2 =ω0± 2Ω1 and a singlepeak for ω1 =ω0 as depicted in Fig. 3F, with much less contrastcompared with the case of TF. Finally, for t1 =T + 20τ , t2 =T ,the PS signal has three overlapping peaks for ω1 =ω0,ω0± 2Ω1

but a single peak for ω2 =ω0 as depicted in Fig. 3H vs. well-resolved peaks shown in Fig. 3D for TF. Therefore, the TF coin-cidence counting signal shows how the response to the timemodulation of the coupling changes the configuration of the JCstates participating in the photon emission for a given cavity pho-ton number, which is missed by the PS. The resonances thatmanifest as cross-peaks in Fig. 3 A, C, and D mark the role ofcoherence between different dressed atom–cavity states that aregoverned by the coupling temporal profile, since the pair ofphotons can be generated from the superposition of two JCladder states.

Note on the Joint Temporal/Spectral ResolutionWe identify two main differences between the PS and TF sig-nals. In PS, both frequency and time resolutions are deter-

mined by a single parameter Γps . The two time integrationsin Eq. S59 are independent, so that each field operator isgated separately. In contrast, the TF signal in Eq. S60 dependson two gating parameters Γ±. In addition, there is an extraθ(t ′′1 − t ′1) factor, which couples the two time integrations inEq. S60. To examine how these parameters determine the tem-poral and spectral resolutions of the signal, we performed anasymptotic expansion of Eq. 7, such that, in zeroth order, onecan approximate it at t0 = 0 as Eq. S63. This contains terms ofthe form (details are in SI Text) e(iωn±−1/(2τ)−Γps/2)(t−T)/(ωn±− iΓps/2), where ωn±=ω−ω0± 2Ωn . This term yields two sidepeaks due to Rabi splitting. Clearly, the spectral resolution isgiven by ∆ωPS = Γps/2, whereas the temporal resolution is givenby ∆tPS = 2/Γps . The product of the two gives the Fourier limit∆tPS ·∆ωPS = 1. The corresponding product of uncertaintiesfor the TF gated signal reads e(iωn±−1/(2τ)−ΓT/2)(t−T)/(ωn±− iΓ+/2). The spectral resolution is now given by Γ+/2, thecombined spectral and temporal gate bandwidth. The tempo-ral resolution is (Γ+/2 + Γ−/2)−1 = 2/ΓT , which is a temporalgate alone. Therefore, ∆tTF ·∆ωPS = 1 + Γω/ΓT . An importantpoint is related to the fact that we achieve the sum Γ+ + Γ− forthe temporal resolution. This is due to the nested time integra-tions and the θ(t ′′1 − t ′1) factor in Eq. S33. To illustrate the role ofthe time gate, we can consider a high spectral resolution, so thatΓω + ΓT = Γps < 2Ωn . We also take 2/ΓT <τ < 2/Γps , such thatthe PS temporal resolution is low, whereas the TF gated signalgives high temporal resolution. Ultrafast temporal resolution ofsingle photons can be readily achieved in strongly coupled quan-tum dot–cavity systems (54) or up-conversion schemes (55). Thepure time-resolved photon counting can be also used to probetemporal evolution of the cavity polaritons. However, similar tocoherent spectroscopy, where time-resolved measurement alonecan only provide partial information about the system vs. multi-dimensional four wave mixing spectroscopy that can distinguishsignal generated from the molecular populations and from the

1454 | www.pnas.org/cgi/doi/10.1073/pnas.1719443115 Dorfman and Mukamel

Dow

nloa

ded

by g

uest

on

Mar

ch 1

0, 2

020

Page 5: } of cavity polaritons · cavity mode is monitored through time-and-frequency–resolved single-photon coincidence measurements of spontaneous emis-sion. Polariton population and

CHEM

ISTR

Y

molecular coherences (due to intermolecular coupling), simulta-neous TF gating allows us to distinguish signals originated frompolariton coherences and populations as follows from Fig. 3.

In summary, we have demonstrated how multidimensionalphoton counting can be used to reveal polariton dynamics ina cavity. The TF gated photon number and photon coinci-dence detection can capture subtle time-evolving features, suchas dressed JC ladder polariton states and their correlations viacross-peaks in the 2D photon coincidence spectra. Note that,in the strong dissipation limit, the oscillations shown in Fig.1 and contributing to the polariton dynamics would be signif-icantly damped. This means that, for systems with the dynam-ics of the timescale similar to the cavity damping rate, one hasto take into account the cavity leakage when evaluating the sig-nals. Recent experiments in light harvesting molecules placed ina microcavity (22) observe large amounts of scattering into themicrocavity system; however, the observed damping is not strongenough to destroy the strong coupling. The technique describedin this paper can be useful for laser stabilization, where the cavityphoton distribution and gain medium dynamics are monitoredsimultaneously. In this scenario, cavity photon statistics acts asan input into the polariton configuration captured by the pho-ton counting signals, which consequently monitors the stabilityof the cavity radiation. The time-dependent coupling offers aversatile coherent control tool. Similarly to the pulse shapingtechnique used in ordinary spectroscopy, one can optimize thecoupling temporal profile in a generic algorithm setup to opti-mize the control. To implement this control scheme, one has tomatch the coupling duration and profile to the timescale of thenuclear motion. In addition to the suppression of the dephas-ing and changing of the dynamical rates that are observed inthe systems with the stationary cavity coupling, time-dependentcoupling provides a unique control mechanism for tracking thepolariton–polariton and other many-body correlation effects viaoptimizing the corresponding coherences visualized as the cross-peaks in 2D spectra in photon coincidence signals.

Materials and MethodsThe Hamiltonian [1] can be recast as H = H0 + Hi , where H0 =ω(∆) and∆ = n + 1+σz

2 is the field energy function given by Eq. S2, while Hi =

δ(∆)σz +λ(t)(A+σ− + A−σ+), and the δ(∆) is given by Eq. S3. For the one-photon JC ladder, ω(∆) = (∆− 1/2)ω and δ(∆) = δ/2, where δ=ω0−ω. Amore general m-photon JC ladder, which can be used for describing mul-tiphoton processes, is presented in SI Text. The subspace in which n, theeigenvalue of ∆, satisfies for n 6=0 the special unitary group of the sec-ond degree [SU(2)] symmetry of ∆. In many molecular polariton applica-tions that involve, for example, nonadiabatic dynamics, the inclusion of thecounterrotating terms is required, which results in band-diagonal structureof the Hamiltonian. While purely analytic, a solution without the rotatingwave approximation is possible only in certain cases (56), and the problemcan be treated in Fock space to allow for a numerically exact solution (57).The time evolution operator may be recast as U(t) = e−iω(∆)tUi(t), whereUi(t) is the evolution operator corresponding to Hi . Using the SU(2) algebraand the Wei–Norman formalism (42, 58, 59), we can recast the evolutionoperator in the form Ui(t) = eh(t)F0 eg(t)F+ ef(t)F− , where F± =±A∓σ±/

√∆,

F0 =σz, and f(t), g(t), and h(t) are generally complex functions, such thatx∗(t) =−x(t), x = f , g, h. Introducing G(t) = g(t)eh(t), F (t) = f(t)e−h(t), andH(t) = e−h(t) using unitary condition, we obtain G(t)=F∗(t). For simplicity

in the following, we assume that the atom is driven on resonance, such thatδ= 0. The functions X =F ,H satisfy the differential equation:

X−λ

λX + 4(n + 1)λ2X = 0. [6]

The initial conditions for F ,H are H(t0) = 1, H(t0) = 0, F (t0) = 0, andF (t0) = i

√n + 1λ(t0) =−iα/(2τ ), where α= 2λ0τ

√n + 1. Normalization

implies |H(t)|2 + |F (t)|2 = 1. Eq. 6 can be solved analytically for thesech function coupling λ(t). Changing the variable to z(t) = e(t−T)/τ/[1 +

e(t−T)/τ ], we obtain a hypergeometric equation with β=−α= 1/2. Thesolution of Eq. 6 subject to initial conditions above is given by a simplifiedhypergeometric function

H(z) = cosh

[2α log

( √z +√

z− 1√

z0 +√

z0− 1

)],

F (z) = sinh

[2α log

( √z +√

z− 1√

z0 +√

z0− 1

)], [7]

where z0≡ z(t0).The matter correlation functions in Eqs. 4 and 5 are traced over the

noncavity modes. Assuming initial wave function |ψ0〉=∑

n[wn|n, ↑〉+vn|n, ↓〉], where the coefficients wn (vn) represent the probability ampli-tude of the atom to be in the excited (ground) state, the field is in a Fockstate with n quanta. These amplitudes satisfy the normalization

∑n(|wn|2 +

|vn|2) = 1. The atomic inversion 〈σz(t)〉′ has been calculated previously (41)and is given by Eq. S51.

To get the photon counting signal, we have calculated two-point [4] andfour-point [5] correlation functions:

〈V†(t1)V(t2)〉′ = |µ|2e−iω0(t1−t2)∑

n

G∗in(t1)Gpn(t1, t2)Gin(t2), [8]

〈V†(t1)V†(t2)V(t3)V(t4)〉′ = |µ|4e−iω0(t1+t2−t3−t4)∑

n

G∗in(t1)G∗cn(t2, t1)

×Gpn(t2, t3)Gcn(t3, t4)Gin(t4), [9]

where the “initial state” Green’s function Gi , “coherence” Gc, and “popula-tion” Gp Green’s functions are defined, respectively, as

Gin(t) = wnHn+1(t)− vn+1F∗n+1(t),

Gcn(ti , tj) =Hn+1(ti)F∗n+1(tj)−F∗n+1(ti)Hn+1(tj),

Gpn(ti , tj) =H∗n+1(ti)Hn+1(tj) +Fn+1(ti)F∗n+1(tj). [10]

Making use of the closed form expressions [7], we have made use of Eqs.8–10 to calculate the gated emission [4] and the coincidence counting [5].Note that Eqs. 8–10 reduce to the simple Rabi oscillations cos(Ωt) in theτ→ 0 limit given by Eqs. S57 and S58, respectively. Additional details aresummarized in SI Text.

The simulations shown in Figs. 1–3 use typical parameters related to vibra-tional spectroscopy: vibrational frequency ω0 = 12600 cm−1, coupling mod-ulation λ0 = 100 cm−1, and centered at T = 111.5 ps. The coupling timescalein Figs. 1 and 3 is τ = 150 fs.

ACKNOWLEDGMENTS. K.E.D. is supported by the Zijiang Endowed YoungScholar Fund and Overseas Expertise Introduction Project for DisciplineInnovation (“111 Project,” B12024). S.M. is supported by National ScienceFoundation Grant CHE-1663822 and Chemical Sciences, Geosciences andBiosciences Division, Office of Basic Energy Sciences, Office of Science, USDepartment of Energy Award DE-FG02-04ER15571.

1. Purcell EM, Torrey H, Pound RV (1946) Resonance absorption by nuclear magneticmoments in a solid. Phys Rev 69:37–38.

2. Jaynes ET, Cummings FW (1963) Comparison of quantum and semiclassical radiationtheories with application to the beam maser. Proc IEEE 51:89–109.

3. Haroche S, Kleppner D (1989) Cavity quantum electrodynamics. Phys Today 42:24–30.4. Cui G, Raymer MG (2005) Quantum efficiency of single-photon sources in the cavity-

qed strong-coupling regime. Opt Express 13:9660–9665.5. Schneebeli L, Kira M, Koch S (2008) Characterization of strong light-matter coupling

in semiconductor quantum-dot microcavities via photon-statistics spectroscopy. PhysRev Lett 101:097401.

6. Kasprzak J, et al. (2010) Up on the Jaynes-Cummings ladder of a quantum-dot/microcavity system. Nat Mater 9:304–308.

7. Galego J, Garcia-Vidal FJ, Feist J (2015) Cavity-induced modifications of molecularstructure in the strong-coupling regime. Phys Rev X 5:041022.

8. Dicke RH (1954) Coherence in spontaneous radiation processes. Phys Rev 93:99–110.

9. Gross M, Haroche S (1982) Superradiance: An essay on the theory of collective spon-taneous emission. Phys Rep 93:301–396.

10. Hutchison JA, Schwartz T, Genet C, Devaux E, Ebbesen TW (2012) Modifying chemicallandscapes by coupling to vacuum fields. Angew Chem Int Ed 51:1592–1596.

11. Shalabney A, et al. (2015) Enhanced Raman scattering from vibro-polariton hybridstates. Angew Chem Int Ed 54:7971–7975.

12. Nussmann S, et al. (2005) Vacuum-stimulated cooling of single atoms in three dimen-sions. Nat Phys 1:122–125.

Dorfman and Mukamel PNAS | February 13, 2018 | vol. 115 | no. 7 | 1455

Dow

nloa

ded

by g

uest

on

Mar

ch 1

0, 2

020

Page 6: } of cavity polaritons · cavity mode is monitored through time-and-frequency–resolved single-photon coincidence measurements of spontaneous emis-sion. Polariton population and

13. Kowalewski M, Morigi G, Pinkse PWH, de Vivie-Riedle R (2011) Cavity sideband cool-ing of trapped molecules. Phys Rev A 84:033408.

14. Lev BL, et al. (2008) Prospects for the cavity-assisted laser cooling of molecules. PhysRev A 77:023402.

15. Spano FC (2015) Optical microcavities enhance the exciton coherence length and elim-inate vibronic coupling in j-aggregates. J Chem Phys 142:184707.

16. Caruso F, et al. (2012) Probing biological light-harvesting phenomena by optical cav-ities. Phys Rev B 85:125424.

17. Sun Y, et al. (2017) Direct measurement of polariton-polariton interaction strength.Nat Phys 13:870–875.

18. Kena-Cohen S, Forrest SR (2010) Room-temperature polariton lasing in an organicsingle-crystal microcavity. Nat Photon 4:371–375.

19. Bellessa J, et al. (2014) Strong coupling between plasmons and organic semiconduc-tors. Electronics 3:303–313.

20. Cacciola A, Di Stefano O, Stassi R, Saija R, Savasta S (2014) Ultrastrong coupling ofplasmons and excitons in a nanoshell. ACS Nano 8:11483–11492.

21. Tsuchimoto Y, Nagai H, Amano M, Bando K, Kondo H (2014) Cavity polaritons in anorganic single-crystalline rubrene microcavity. Appl Phys Lett 104:233307.

22. Coles DM, et al. (2014) Strong coupling between chlorosomes of photosynthetic bac-teria and a confined optical cavity mode. Nat Commun 5:5561.

23. Muallem M, Palatnik A, Nessim GD, Tischler YR (2016) Strong light-matter couplingbetween a molecular vibrational mode in a pmma film and a low-loss mid-ir micro-cavity. Annalen der Physik 528:313–320.

24. Herrera F, Spano FC (2016) Cavity-controlled chemistry in molecular ensembles. PhysRev Lett 116:238301.

25. Kowalewski M, Bennett K, Mukamel S (2016) Cavity femtochemistry: Manipulatingnonadiabatic dynamics at avoided crossings. J Phys Chem Lett 7:2050–2054.

26. Ebbesen TW (2016) Hybrid light–matter states in a molecular and material scienceperspective. Acc Chem Res 49:2403–2412.

27. Flick J, Ruggenthaler M, Appel H, Rubio A (2017) Atoms and molecules in cavities,from weak to strong coupling in quantum-electrodynamics (qed) chemistry. Proc NatlAcad Sci USA 114:3026–3034.

28. Herrera F, Peropadre B, A. Pachon L, Saikin S, Aspuru-Guzik A (2014) Quantum non-linear optics with polar j-aggregates in microcavities. J Phys Chem Lett 5:3708–3715.

29. Shalabney A, et al. (2014) Coherent coupling of molecular resonators with a micro-cavity mode. Nat Commun 6:5981.

30. del Pino J, Feist J, Garcia-Vidal FJ (2015) Quantum theory of collective strong couplingof molecular vibrations with a microcavity mode. New J Phys 17:053040.

31. Dressick W, et al. (2015) Spanning strong to weak normal mode coupling betweenvibrational and Fabry-Perot cavity modes through tuning of vibrational absorptionstrength. 2:1460–1467.

32. Wen P, Christmann G, Baumberg JJ, Nelson KA (2013) Influence of multi-exciton cor-relations on nonlinear polariton dynamics in semiconductor microcavities. New J Phys15:025005.

33. Ernst RR, Bodenhausen G, Wokaun A, et al. (1987) Principles of Nuclear MagneticResonance in One and Two Dimensions (Clarendon, Oxford), Vol 14.

34. Cundiff ST, Mukamel S (2013) Optical multidimensional coherent spectroscopy. PhysToday 66:44–49.

35. Saurabh P, Mukamel S (2016) Two-dimensional infrared spectroscopy of vibrationalpolaritons of molecules in an optical cavity. J Chem Phys 144:124115.

36. Fleury L, Segura JM, Zumofen G, Hecht B, Wild UP (2000) Nonclassical photon statisticsin single-molecule fluorescence at room temperature. Phys Rev Lett 84:1148–1151.

37. Lettow R, et al. (2010) Quantum interference of tunably indistinguishable photonsfrom remote organic molecules. Phys Rev Lett 104:123605.

38. Rezus YLA, et al. (2012) Single-photon spectroscopy of a single molecule. Phys RevLett 108:093601.

39. Dorfman KE, Mukamel S (2014) Indistinguishability and correlations of pho-tons generated by quantum emitters undergoing spectral diffusion. Sci Rep 4:3996.

40. Rempe G, Walther H, Klein N (1987) Observation of quantum collapse and revival ina one-atom maser. Phys Rev Lett 58:353–356.

41. Nasreen T, Razmi M (1993) Atomic emission and cavity field spectra for a two-photon Jaynes–Cummings model in the presence of the stark shift. JOSA B 10:1292–1300.

42. Joshi A, Lawande S (1993) Generalized Jaynes-Cummings models with a time-dependent atom-field coupling. Phys Rev A 48:2276–2284.

43. Deppe F, et al. (2008) Two-photon probe of the Jaynes–Cummings model and con-trolled symmetry breaking in circuit qed. Nat Phys 4:686–691.

44. Cocker TL, Peller D, Yu P, Repp J, Huber R (2016) Tracking the ultrafast motion of asingle molecule by femtosecond orbital imaging. Nature 539:263–267.

45. Cordero S, Recamier J (2012) Algebraic treatment of the time-dependent Jaynes–Cummings Hamiltonian including nonlinear terms. J Phys A Math Theor 45:385303.

46. Harbola U, Mukamel S (2008) Superoperator nonequilibrium green’s function theoryof many-body systems; applications to charge transfer and transport in open junc-tions. Phys Rep 465:191–222.

47. Rahav S, Mukamel S (2010) Ultrafast nonlinear optical signals viewed from themolecules perspective: Kramers-Heisenberg transition amplitudes vs. susceptibilities.Adv At Mol Opt Phys 59:223.

48. Dorfman KE, Mukamel S (2016) Time-and-frequency-gated photon coincidencecounting; a novel multidimensional spectroscopy tool. Phys Scr 91:083004.

49. Dorfman KE, Schlawin F, Mukamel S (2016) Nonlinear optical signals and spectroscopywith quantum light. Rev Mod Phys 88:045008.

50. Eberly J, Wodkiewicz K (1977) The time-dependent physical spectrum of light. JOSA67:1252–1261.

51. del Valle E, Gonzalez-Tudela A, Laussy FP, Tejedor C, Hartmann MJ (2012) The-ory of frequency-filtered and time-resolved n-photon correlations. Phys Rev Lett109:183601.

52. del Valle E (2013) Distilling one, two and entangled pairs of photons from a quantumdot with cavity qed effects and spectral filtering. New J Phys 15:025019.

53. Gonzalez-Tudela A, Del Valle E, Laussy FP (2015) Optimization of photon correlationsby frequency filtering. Phys Rev A 91:043807.

54. Volz T, et al. (2012) Ultrafast all-optical switching by single photons. Nat Photon6:605–609.

55. Ma L, Bienfang JC, Slattery O, Tang X (2011) Up-conversion single-photon detectorusing multi-wavelength sampling techniques. Opt Express 19:5470–5479.

56. Judd BR (1979) Exact solutions to a class of jahn-teller systems. J Phys C Solid StatePhys 12:1685.

57. Kus M, Lewenstein M (1986) Exact isolated solutions for the class of quantum opticalsystems, optical systems. J Phys A Math Gen 19:305–318.

58. Wei J, Norman E (1963) Lie algebraic solution of linear differential equations. J MathPhys 4:575–581.

59. Dattoli G, Di Lazzaro P, Torre A (1987) Su(1,1), su(2), and su(3) coherence-preservingHamiltonians and time-ordering techniques. Phys Rev A 35:1582–1589.

1456 | www.pnas.org/cgi/doi/10.1073/pnas.1719443115 Dorfman and Mukamel

Dow

nloa

ded

by g

uest

on

Mar

ch 1

0, 2

020