This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Resonator-assisted single molecule quantum state detection

Ming Zhu Department of Physics and Astronomy, Purdue University, West Lafayette, IN 47907, USA    Yan-Cheng Wei Department of Physics and Astronomy, Purdue University, West Lafayette, IN 47907, USA Department of Physics, National Taiwan University, Taipei 10617, Taiwan    Chen-Lung Hung Email: [email protected] Department of Physics and Astronomy, Purdue University, West Lafayette, IN 47907, USA Purdue Quantum Science and Engineering Institute, Purdue University, West Lafayette, IN 47907, USA
Abstract

We propose a state-sensitive scheme to optically detect a single molecule without a closed transition, through strong coupling to a high-Q whispering-gallery mode high-Q resonator. A background-free signal can be obtained by detecting a molecule-induced transparency in a photon bus waveguide that is critically coupled to the resonator, with a suppressed depumping rate to other molecular states by the cooperativity parameter CC. We numerically calculate the dynamics of the molecule-resonator coupled system using Lindblad master equations, and develop analytical solutions through the evolution of quasi-steady states in the weak-driving regime. Using Rb2 triplet ground state molecules as an example, we show that high fidelity state readout can be achieved using realistic resonator parameters. We further discuss the case of multiple molecules collectively coupled to a resonator, demonstrating near-unity detection fidelity and negligible population loss.

I Introduction

The ability to determine the state of a single quantum emitter is essential for quantum information processing. Resonance fluorescence imaging is a convenient and powerful method. During imaging, exciting closed optical transitions ensures that a quantum emitter scatters a large number of photons without leaving a specific ground state, thus making it possible to achieve state-sensitive detection with a high signal-to-noise ratio, even with low photon collection efficiency. This approach has been widely employed in various quantum systems, such as cold atoms Bochmann et al. (2010); Gehr et al. (2010); Fuhrmanek et al. (2011), trapped ions Olmschenk et al. (2007); Myerson et al. (2008), nitrogen vacancy centers Neumann et al. (2010); Robledo et al. (2011) and quantum dots Vamivakas et al. (2010).

Many promising quantum systems are not suited for fluorescent imaging due to their energy structure. For instance, cold molecules have a wide range of applications in quantum chemistry Bohn et al. (2017) and quantum computation DeMille (2002). However, most of the molecules have no real optical cycling transitions due to a myriad of rovibrational levels accessible in a radiative decay process. For specific kinds of molecules, one may find a transition with near-unity Franck-Condon factor for a target ground state, and use multiple lasers to drive an approximately closed transition in a manageable collection of ground and excited states. For instance, SrF Shuman et al. (2010) or CaF Anderegg et al. (2017) molecules can be laser cooled and trapped in a magneto-optical trap.

In general, alternative state detection method is needed for molecules without optically closed transitions. Resonance-enhanced multiphoton ionization is adopted for molecule detection Gabbanini et al. (2000), which ionizes the molecules and detects the subsequent ions. This method is nevertheless completely destructive. Direct absorption imaging of molecules is also reported Wang et al. (2010), when a molecular ensemble has a high optical density.

Cavity quantum electrodynamics opens a new way to implement state detection of a single molecule. Cavity-controlled light-matter interaction enables the manipulation of molecular photon emission properties. When a molecule is located inside a cavity or near a resonator which is tuned to the molecular resonance of interest, the branching ratio of decay into irrelevant states may be greatly suppressed Lev et al. (2008), allowing the interaction with resonator photons for an extended period of time. In addition, considering the molecules directly emit photons into the resonator mode(s), the signal photon collection efficiency may be significantly enhanced compared to the case of emission in freespace.

In this paper, we consider a scheme to detect the quantum state of molecules without a closed transition by utilizing high-finesse Fabry-Perot cavities Hood et al. (2001) or high-Q whispering-gallery mode (WGM) resonators Vernooy et al. (1998); Spillane et al. (2005); Pöllinger et al. (2009); Tien et al. (2011); Chang et al. (2019). Our scheme is inspired by the pioneering experiments for detecting single atoms falling through a microtoroidal resonator Aoki et al. (2006); Shomroni et al. (2014), and probing trapped single atoms inside a mirror cavity Boozer et al. (2006); Khudaverdyan et al. (2009), a fiber-based cavity Gehr et al. (2010); Kato and Aoki (2015), as well as in the vicinity of a photonic crystal cavity Thompson et al. (2013); Goban et al. (2014). In particular, we consider the transmission of a bus waveguide critically coupled to WGMs in a micro-ring resonator [Fig. 1(a)], displaying molecule-induced transparency on resonance due to strong light-molecule interaction. The proposed detection scheme is background-free, as there is no waveguide transmission unless a molecule in the target state couples to the resonator. The maximum scattered photon number is enhanced by the cooperativity parameter CC before the molecule is completely pumped away from its initial state. We numerically simulate this open quantum system with a Lindblad master equation and derive analytical solutions in the weak-driving regime. We then extend the single-molecule model to a multi-molecule case considering collective effect. Similar scheme can be realized in monitoring the off-resonant transmissivity in the case of a Febry-Perot cavity (Appendix A). Our scheme can also be applied to single-shot state readout for other quantum emitters Sun and Waks (2016).

For illustration purposes, we take Rb2 molecule as an example. The relevant internuclear potentials are plotted in Fig. 1(b), in which a single-photon, short-range photoassociation has been utilized for ground state molecule synthesis directly from cold atoms Bellos et al. (2011). Using an optical cavity or a photonic crystal waveguide to radiatively enhance the synthesis efficiency has recently been discussed Perez-Rios et al. (2017); Kampschulte and Denschlag (2018). These recent developments makes Rb2 a good candidate to demonstrate resonator-enhanced single molecule detection.

The paper is organized as follows. Section II reviews the optical setup of the resonator as well as a simplified energy structure of the Rb2 molecule. In Sec. III, we derive a formalism to calculate single molecule dynamics with one resonator mode. In Sec. IV, the discussion is extended to the coupling between a single molecule and two resonator modes. In Sec. V, we discuss the multi-molecule dynamics by taking into account collective effects.

Refer to caption
Figure 1: (color online) Schematic illustration of the investigated system. (a) Optical setup marked by basic rates. A resonator supports both clockwise (red solid arrows) and counter-clockwise (blue dashed arrows) circulating whispering-gallery modes (WGMs), and is excited by a bus waveguide. (b) Sample transition in Rb2 molecule between the rovibrational ground state |g\ket{g} in the a3Σu+a^{3}\Sigma_{u}^{+} potential and a molecular excited state |e\ket{e} in the 13Πg1^{3}\Pi_{g} potential. State |s\ket{s} represents a collection of all other rovibrational levels in the a3Σu+a^{3}\Sigma_{u}^{+} potential. A molecule in state |g\ket{g} (shaded circle in (a)) interacts with the WGM(s) at a coupling rate gcg_{c}.

II The Optical Setup and Molecular Model

We first introduce the optical system, a bus waveguide coupled to an empty micro-resonator as shown in Fig. 1(a), which supports WGMs that circulate either in the clockwise (CW) or the counter-clockwise (CCW) orientation along the resonator. We assume an input field of power PIP_{I} is injected from one end of the bus waveguide, and analyze the transmission power PTP_{T} and reflection power PRP_{R}. In the following cases, we assume that back-scattering in the resonator, which couples the CW and CCW modes, occurs at a rate much smaller than the total resonator loss rate κ\kappa, and thus there is negligible mode mixing. Here, κ=κe+κi\kappa=\kappa_{e}+\kappa_{i}, where κi\kappa_{i} is the intrinsic photon loss rate and κe\kappa_{e} is the waveguide-resonator coupling rate. Due to the phase matching condition, when the resonator couples light from single end of the bus waveguide shown in Fig. 1(a), only one mode (CW WGM, illustrated as solid line) can be excited and the resonator can be treated like a single mode cavity. Reflection stays at zero (PR=0P_{R}=0) due to the absence of CCW WGM excitation.

To achieve background-free molecule detection, we consider a critically coupled resonator (κi=κe\kappa_{i}=\kappa_{e}) for zero waveguide transmissivity on resonance (Δcl=0\Delta_{\text{cl}}=0). The bus waveguide transmissivity is evaluated by solving the standard Heisenburg-Langevin equation with the single-mode Hamiltonian of an empty resonator

H^0=Δcla^a^+i(εa^εa^),\hat{H}_{0}=\Delta_{\text{cl}}\hat{a}^{\dagger}\hat{a}+i\left(\varepsilon\hat{a}^{\dagger}-\varepsilon^{*}\hat{a}\right), (1)

where Δcl=ωcωl\Delta_{\text{cl}}=\omega_{c}-\omega_{l} is the detuning between the resonant mode frequency ωc\omega_{c} and the external driving frequency ωl\omega_{l}, a^()\hat{a}^{({\dagger})} represents the annihilation (creation) operator of the (CW) resonator mode, excited from the bus waveguide at a rate coefficient ε=i2κe\varepsilon=i\sqrt{2\kappa_{e}\mathcal{I}} and =PI/(ωl)\mathcal{I}=P_{I}/\left(\hbar\omega_{l}\right) is the waveguide photon input rate. At steady state, the expectation value of intra-resonator field is found to be a^=ε/(iΔcl+κ)\expectationvalue{\hat{a}}=\varepsilon/\left(i\Delta_{\textrm{{cl}}}+\kappa\right). The bus waveguide transmission is the interference between the input field \sqrt{\mathcal{I}} and the out-coupled field from the resonator i2κeai\sqrt{2\kappa_{e}}\expectationvalue{a}, resulting in a transmissivity Haus (1984)

T=PTPI=|1+i2κea^|2.T=\frac{P_{T}}{P_{I}}=\left|1+i\sqrt{\frac{2\kappa_{e}}{\mathcal{I}}}\langle\hat{a}\rangle\right|^{2}. (2)

As shown in Fig. 2(c), at critical coupling the waveguide transmissivity, T=|Δcl/(κ+iΔcl)|2T=\absolutevalue{\Delta_{\text{cl}}/(\kappa+i\Delta_{\text{cl}})}^{2}, drops to zero on resonance. As we discuss in the following, the waveguide transmissivity will be greatly modified when a molecule is present and couples to the resonator. This establishes our scheme to realize background-free molecule detection.

Refer to caption
Figure 2: (a) Simplified energy level structure for one molecule coupled to single resonator mode in the weak-driving regime for ωm=ωc\omega_{m}=\omega_{c}. (b) Transmission spectrum of a molecule-coupled resonator with Γs=0\Gamma_{s}=0. (c) Transmission spectrum of an empty resonator. Background-free measurement can be performed on resonance δΔcl=0\delta\equiv\Delta_{\textrm{cl}}=0.

To model the radiative dynamics of a resonator-coupled molecule, we treat it as an effective three-level system, as illustrated in Fig. 1(b). While radiative decay processes can couple a molecular excited state |e\ket{e} to a collection of (electronic) ground states of different rovibrational energy levels, the molecule-resonator coupling can in principle involve only one designated rovibrational state |g\ket{g} when the resonator frequency ωc\omega_{c} is aligned with the transition frequency ωm\omega_{m} and the resonator linewidth (O(1)\lesssim O(1)~{}GHz) is smaller than the relevant rovibrational energy level spacing (O(10)\gtrsim O(10)~{}GHz) by over an order of magnitude. A third state |s\ket{s} denotes all other uncoupled states that can accumulate population from spontaneous decay. Given a Franck-Condon factor fFCf_{\mathrm{FC}} between states |e\ket{e} and |g\ket{g}, the spontaneous decay rate to |g\ket{g} is Γg=fFCΓ\Gamma_{g}=f_{\mathrm{FC}}\Gamma and the decay rate to |s\ket{s} is Γs=(1fFC)Γ\Gamma_{s}=(1-f_{\mathrm{FC}})\Gamma, where Γ\Gamma is the total decay rate of the excited state |e\ket{e}.

We denote the coherent coupling rate between the molecule and the resonator mode as gc=3Γgc3/2Vωm2\displaystyle g_{c}=\sqrt{3\Gamma_{g}c^{3}/2V\omega_{m}^{2}}, where cc is the speed of light, V=ϵ(𝐫)|(𝐫)|2𝑑𝐫/|(𝐫mol)|2\displaystyle V=\int\epsilon(\mathbf{r})|\mathcal{E}(\mathbf{r})|^{2}d\mathbf{r}/\left|\mathcal{E}\left(\mathbf{r}_{\mathrm{mol}}\right)\right|^{2} is the mode volume, ϵ(𝐫)\epsilon(\mathbf{r}) is the dielectric function, (𝐫)\mathcal{E}\left(\mathbf{r}\right) is the mode field strength, and 𝐫mol\mathbf{r}_{\mathrm{mol}} is the molecular position. gcg_{c} is position dependent due to the mode field intensity variation near the resonator dielectric surface. We assume that molecules are trapped in close proximity of a resonator and hence gcg_{c} is a constant.

Table 1: Sample parameters adopted for the investigated platform, using Rb2 (|g\ket{g}: ν=0\nu^{\prime}=0 in a3Σu+a^{3}\Sigma^{+}_{u}, |e\ket{e}: ν=8\nu=8 in 13Πg+1^{3}\Pi_{g}^{+})Bellos et al. (2011)
Parameter Symbol Value
Total spontaneous decay rate (|e\ket{e}) Γ\Gamma 2π×12MHz2\pi\times 12\mathrm{MHz}
Franck-Condon factor (|e|g\ket{e}\Longleftrightarrow\ket{g}) fFCf_{FC} 0.37
Photon input rate \mathcal{I} 1MHz1\mathrm{MHz}
Resonator intrinsic loss rate κi\kappa_{i} 2π×50MHz2\pi\times 50\mathrm{MHz}
External coupling rate κe\kappa_{e} 2π×50MHz2\pi\times 50\mathrm{MHz}
Sample cooperativity parameter CC 50
Resonator coupling rate (C=50C=50) gcg_{c} 2π×245MHz2\pi\times 245\mathrm{MHz}

Table 1 lists typical resonator and molecule parameters used in the numerical and analytical calculations throughout this paper. We take Rb2 molecules coupled to a micro-ring resonator as an example. The relevant internuclear potentials are plotted in Fig. 1(b), where |g\ket{g} stands for the rovibrational ground state of interest in the a3Σu+a^{3}\Sigma^{+}_{u} triplet potential, and |e\ket{e} represents an excited state in the 13Πg1^{3}\Pi_{g} potential (vibrational level ν=8\nu=8, and angular momentum quantum number J=1J=1) Bellos et al. (2011). Due to the finite Frank-Condon overlap between |g\ket{g} and |e\ket{e}, this transition has been utilized for ground state molecule synthesis directly from cold atoms via a single-photon short-range photoassociation to state |e\ket{e}, followed by spontaneous decay into |g\ket{g} Bellos et al. (2011). On the other hand, our model optical system is adapted from a recent report on high-Q micro-ring resonators Chang et al. (2019), where we assume that high quality factor Q>105Q>10^{5} and large single-photon vacuum Rabi frequency 2gc2π×5002g_{c}\sim 2\pi\times 500~{}MHz can be simultaneously realized to achieve a large coorperativity parameter Cgc2/κΓC\equiv g_{c}^{2}/\kappa\Gamma, which is the key parameter to achieve high single-molecule detection sensitivity.

III Single Molecule Dynamics coupled with one resonator mode

We now analyze the recovery of bus waveguide transmissivity on resonance with the presence of a single molecule as shown in Fig. 2(b). We begin with the first scenario where a molecule couples to a single resonator mode. This applies to the case when a molecule, spin-polarized in a stretched state, couples only to a circularly polarized WGM, and cannot emit photons into the other WGM because of its opposite circular polarization state 111The WGMs are circularly polarized since they are traveling wave modes with strong evanescent field outside the waveguide. A ±π/2\pm\pi/2 degree out-of-phase axial field component, dictated by the transversality of the Maxwell’s equation and time reversal symmetry, gives the CW and CCW WGMs opposite circular polarization states.. The single-mode light-molecule interaction Hamiltonian H^1\hat{H}_{1} can be written as

H^1=Δmlσ^+σ^+igc(a^σ^a^σ^+),\hat{H}_{1}=\Delta_{\text{ml}}\hat{\sigma}_{+}\hat{\sigma}_{-}+ig_{c}\left(\hat{a}^{{\dagger}}\hat{\sigma}_{-}-\hat{a}\hat{\sigma}_{+}\right), (3)

where Δml=ωmωl\Delta_{\text{ml}}=\omega_{m}-\omega_{l}, σ^=|ge|\hat{\sigma}_{-}=\ket{g}\bra{e}, and σ^+=|eg|\hat{\sigma}_{+}=\ket{e}\bra{g}.

Taking into account of the resonator loss and the molecule spontaneous emission, the master equation of the full system is written as

dρdt=\displaystyle\frac{d\rho}{dt}= i[H0^+H1^,ρ]+\displaystyle-i\left[\hat{H_{0}}+\hat{H_{1}},\rho\right]+ (4)
2κ[a^]ρ+Γg[σ^]ρ+Γs[σ^]ρ,\displaystyle 2\kappa\mathcal{L}[\hat{a}]\rho+\Gamma_{g}\mathcal{L}[\hat{\sigma}_{-}]\rho+\Gamma_{s}\mathcal{L}[\hat{\sigma}_{-}^{\prime}]\rho,

where ρ\rho is the density matrix of the molecule and photon system and the Lindblad operators take the form of [b^]ρ=b^ρb^12b^b^ρ12ρb^b^\mathcal{L}[\hat{b}]\rho=\hat{b}\rho\hat{b}^{\dagger}-\frac{1}{2}\hat{b}^{\dagger}\hat{b}\rho-\frac{1}{2}\rho\hat{b}^{\dagger}\hat{b} and σ^=|se|\hat{\sigma}_{-}^{\prime}=\ket{s}\bra{e}. As shown in Table 1, both the coherent coupling rate gcg_{c} and the total resonator loss rate κ=κi+κe=2κe\kappa=\kappa_{i}+\kappa_{e}=2\kappa_{e} are at least an order of magnitude larger than the molecule spontaneous decay rates Γg(s)\Gamma_{g(s)}, thus allowing us to utilize fast resonator-molecule dynamics for state detection before losing the population into the uncoupled states |s\ket{s}.

In the limit of single excitation, the resonator and the molecule form an effective five-level system with states denoted by |m,n\ket{m,n} as shown in Fig. 2(a), where mm represents the molecular quantum state and n=0,1n=0,1 is the photon number in the WGM. Three coupled states |e,0\ket{e,0} and |g,0(1)\ket{g,0(1)} on the left of Fig. 2(a) evolve under the cavity QED Hamiltonian H0^+H1^\hat{H_{0}}+\hat{H_{1}} that can quickly establish a quasi-steady state under external driving. Population in this subsystem then slowly decays into the uncoupled states illustrated in the right part of Fig. 2(a), evolving as an empty resonator.

Refer to caption
Figure 3: (color online) Time evolution of (a) state populations and (b) resonator photon number aa\expectationvalue{a^{{\dagger}}a} under a resonant weak-driving (δ\delta=0), evaluated analytically (lines) and numerically (symbols) using cooperativity parameters C=10C=10 (dotted lines), and 50 (solid lines), respectively. For other system parameters used in this and the remaining figures, see Table 1.

Figure 2(b) shows the steady-state transmissivity in an ideal cavity QED system for Γs=0\Gamma_{s}=0. The transmission spectrum is evaluated by substituting the expectation value of the field amplitude a^\langle\hat{a}\rangle in Eq. (2) with the steady-state solution

a^=iΔml+Γ/2gc2+(iΔcl+κ)(iΔml+Γ/2)ε.\expectationvalue{\hat{a}}=\frac{i\Delta_{\text{ml}}+\Gamma/2}{g_{c}^{2}+(i\Delta_{\text{cl}}+\kappa)(i\Delta_{\text{ml}}+\Gamma/2)}\varepsilon. (5)

Recalling that the WGM frequency ωc\omega_{c} is aligned to the molecular transition frequency ωm\omega_{m}, we define detuning δΔclΔml\delta\equiv\Delta_{\text{cl}}\equiv\Delta_{\text{ml}}. When the cooperativity C1C\gg 1, the transmissivity at zero detuning (δ=0\delta=0),

T0=|1κΓ/2gc2+κΓ/2|211C,T_{0}=\left|1-\frac{\kappa\Gamma/2}{g_{c}^{2}+\kappa\Gamma/2}\right|^{2}\approx 1-\frac{1}{C}, (6)

nearly recovers to unity. This effect can be understood as the interference of two molecule-photon dressed states that results in a molecule-induced transparency window, similar to an electromagnetically-induced transparency (EIT). The EIT-like effect contrasts the vanishing transmissivity of an empty resonator critically coupled to the bus waveguide as shown in Fig. 2(c). This forms a highly sensitive scheme for quantum state detection similarly found in Aoki et al. (2006).

In the realistic case of Γs0\Gamma_{s}\neq 0, transmission can only recover for a finite period of time. One expects that the transmission spectrum evolves transiently from an EIT-like curve of Fig. 2(b) to the empty resonator case as shown in Fig. 2(c). Nevertheless, strong resonator coupling suppresses the excited state |e\ket{e} population, resulting in a much reduced decay rate into the uncoupled states |s\ket{s} compared to the free-space decay rate Γs\Gamma_{s}. Finite transmission through the bus waveguide can thus be collected for a finite time period for molecular state detection.

Using the separation of time scales, we find the analytical solution for the quasi-steady density matrix ρss\rho^{ss} of the system, as detailed in Appendix B, and evaluate the population transfer rate DD to the empty resonator. In the weak-driving regime (|ϵ|κ,gc2/κ\displaystyle\absolutevalue{\epsilon}\ll\kappa,g_{c}^{2}/\kappa), we find the population of the molecule-resonator dressed state primarily resides in |g,0\ket{g,0}. The transfer rate DD is greatly suppressed due to a small population in |e,0\ket{e,0}. Based on Eq. (36), We find

D(δ)=ρe0,e0ssρg0,g0ssΓs=gc2κ|gc2+(iδ+κ)(iδ+Γ2)|2Γs,D(\delta)=\frac{\rho_{e0,e0}^{ss}}{\rho_{g0,g0}^{ss}}\Gamma_{s}=\frac{g_{c}^{2}\kappa\mathcal{I}}{\absolutevalue{g_{c}^{2}+(i\delta+\kappa)\left(i\delta+\frac{\Gamma}{2}\right)}^{2}}\Gamma_{s}, (7)

where ρe0,e0(ρg0,g0)\rho_{e0,e0}~{}(\rho_{g0,g0}) is the population in the excited state |e,0\ket{e,0} (ground state |g,0\ket{g,0}), and the superscript ssss stands for steady state. D/D/\mathcal{I} also represents the probability for the dressed state to decay into |s,0\ket{s,0}.

At zero detuning, we find the decay rate

Dres=C(1/2+C)2ΓsΓ.D_{\mathrm{res}}=\frac{C}{(1/2+C)^{2}}\frac{\Gamma_{s}}{\Gamma}\mathcal{I}. (8)

Comparing Dres/D_{\mathrm{res}}/\mathcal{I} with the depumping probability in freespace Γs/Γ=(1fFC)\Gamma_{s}/\Gamma=(1-f_{FC}), the resonator-enhanced |e\ket{e}-|g\ket{g} transition enjoys a large suppression factor 1/C\sim 1/C for depumping into the uncoupled states when C1C\gg 1. As we will show in Eq. (14), this factor suggests that around C/(1fFC)\sim C/(1-f_{FC}) photons may be transmitted through the bus waveguide before the system is converted into an empty resonator and again blocks all resonant input photons.

We validated the slow-transfer model with full numerical calculations Johansson et al. (2013) using the master equation Eq. (4) and the parameters listed in Table 1; see Fig. 3 for an example. The numerical result shows negligible differences from the analytical model in the mean resonator photon number a^a^\langle\hat{a}^{\dagger}\hat{a}\rangle [Eq. (40)], as well as the state populations

Pg(t)\displaystyle P_{g}(t) ρg0,g0(t)=exp(Drest),\displaystyle\approx\rho_{g0,g0}(t)=\exp(-D_{\mathrm{res}}t), (9)
Ps(t)\displaystyle P_{s}(t) ρs0,s0(t)=1exp(Drest),\displaystyle\approx\rho_{s0,s0}(t)=1-\exp(-D_{\mathrm{res}}t),
Pe(t)\displaystyle P_{e}(t) =Pg(t)Dres/Γs.\displaystyle=P_{g}(t)D_{\mathrm{res}}/\Gamma_{s}.

In the following discussions, we will mainly present our analytical analysis.

Now we derive the transmission spectrum by evaluating the time evolution of the quasi-steady resonator field a^=ρg1,g0ss+ρs1,s0ss\expectationvalue{\hat{a}}=\rho^{ss}_{g1,g0}+\rho^{ss}_{s1,s0} as detailed in Eq. (39), where ρg1,g0\rho_{g1,g0} and ρs1,s0\rho_{s1,s0} are the off-diagonal density matrix elements between states (|g,1(\ket{g,1}, |g,0)\ket{g,0}) and (|s,1(\ket{s,1}, |s,0)\ket{s,0}), respectively. Substituting a^\expectationvalue{\hat{a}} in Eq. (2), we find

T(δ)=|iδ+κeDtiδ+κ(iδ+Γ/2)κeDtgc2+(iδ+κ)(iδ+Γ/2)|2.T(\delta)=\absolutevalue{\frac{i\delta+\kappa e^{-Dt}}{i\delta+\kappa}-\frac{(i\delta+\Gamma/2)\kappa e^{-Dt}}{g_{c}^{2}+(i\delta+\kappa)(i\delta+\Gamma/2)}}^{2}. (10)

The instantaneous transmission resembles that of a cavity QED system [black curve in Fig. 4(b)] with an EIT window near δ=0\delta=0 and two absorption dips at δ=±g\delta=\pm g separated by the vacuum Rabi frequency; T(δ)T(\delta) eventually evolves to be δ2/(δ2+κ2)\delta^{2}/(\delta^{2}+\kappa^{2}), the transmissivity of an empty resonator.

At zero detuning, the transmissivity T(0)T(0) decays exponentially with increasing input photon number t\mathcal{I}t as

T(0)=T0e2DrestC1exp(2(1fFC)tC),T(0)=T_{0}e^{-2D_{\text{res}}t}\overset{C\gg 1}{\approx}\exp(-\frac{2(1-f_{FC})\mathcal{I}t}{C}), (11)

which is robust against decay when C1C\gg 1. Figure 4(a) illustrates sample transmission curves at different cooperativity parameters CC.

Refer to caption
Figure 4: (a) Time evolution of resonant transmissivity T(0)T(0). (b) Time-averaged transmission spectra T¯(δ,τ)\overline{T}(\delta,\tau) evaluated under various time intervals (0, τ\tau) with C=50C=50.
Refer to caption
Figure 5: (color online) (a) Transmitted photon number NT(τ)N_{T}(\tau) versus input photon number τ\mathcal{I}\tau (solid curves from bottom to top) with cooperativity parameters as indicated in the legend of (b). Dashed line represents the upper bound NT=τN_{T}=\mathcal{I}\tau. Inset shows NT,maxNT()N_{T,\text{max}}\equiv N_{T}(\infty). (b) Ground state molecule population PgP_{g} versus NTN_{T}. (c) State detection fidelity \mathcal{F} versus depumping probability PsP_{s} for an overall photon counting efficiency η=0.3\eta=0.3 and a dark-count rate dark=100\mathcal{I}_{dark}=100 Hz.

As transmission measurement typically involves finite integration time, we calculate the time-averaged transmission spectra

T¯(δ,τ)=1τ0τT(δ)𝑑t\overline{T}(\delta,\tau)=\displaystyle\frac{1}{\tau}\int_{0}^{\tau}T(\delta)dt (12)

under various time intervals (0,τ)(0,\tau). Figure 4(b) shows how the transmitted signal at various laser detuning δ\delta evolves with the integration time τ\tau. These spectra demonstrate the transition from an EIT-like behavior in a molecule-coupled resonator to the resonant absorption spectrum in an empty resonator – Two initial transmission dips and a transparency window near δ=0\delta=0, formed by the destructive interference of two molecule-photon dressed states, gradually fade away to be overtaken by the single resonance of an empty resonator. Apparently the transmission signal at zero detuning or near the two dips corresponding to the dress-state resonances particularly provides sensitive transient signal for the detection of a molecule in the coupled ground state |g\ket{g}.

We consider background-free transmission at zero detuning since it allows us to take the interrogation time τ\tau\rightarrow\infty and maximize the number of transmitted photons for detection. Figure 5(a) displays the relationship between transmitted photon number NT(τ)N_{T}(\tau) and input photon number τ\mathcal{I}\tau, where

NT(τ)=T¯(0,τ)τC1C2(1fFC)(1e2Dresτ).N_{T}(\tau)=\overline{T}(0,\tau)\mathcal{I}\tau\overset{C\gg 1}{\approx}\frac{C}{2(1-f_{FC})}\left(1-e^{-2D_{\text{res}}\tau}\right). (13)

The diagonal dashed line in Fig. 5(a) represents the optimal case of unity transmissivity for CC\rightarrow\infty and for a closed transition fFC=1f_{FC}=1. All other cases of finite CC and fFC<1f_{FC}<1 fall short of the optimal case and gradually saturates at a maximum photon number

NT,maxNT(τ)C1C2(1fFC),N_{T,\text{max}}\equiv N_{T}(\tau\rightarrow\infty)\overset{C\gg 1}{\approx}\frac{C}{2(1-f_{FC})}, (14)

indicating that the molecule eventually decouples from the resonator due to pumping to |s\ket{s}. Figure 5(a) inset shows the approximate linear dependence of NT,maxN_{T,\text{max}} on the cooperativity CC when C1C\gg 1.

On the other hand, for a background-free setup with total single photon counting efficiency η\eta, the threshold photon number NthN_{\mathrm{th}} for molecule detection can be made far less than NT,maxN_{T,\text{max}}. It is possible to conduct a nearly non-destructive state measurement, that is, preserving the initial molecular state following state detection.

Refer to caption
Figure 6: Maximum transmitted photon number NT,maxN_{T,max} for transitions with different Franck-Condon factors fFCf_{FC} in the cases of cooperativity C=50C=50 (solid line). Collectable photon number for fluorescence scattering in freespace is plotted for comparison (dashed line), assuming perfect collection efficiency.

In Fig. 5(b-c), we calculate the relationship between the transmitted photon number NT(τ)N_{T}(\tau), the depumping probability Ps1PgP_{s}\approx 1-P_{g}, and the estimated detection fidelity. We consider approximate Poisson distributions in both the background counts (mean number n¯s=darkτ\bar{n}_{s}=\mathcal{I}_{\mathrm{dark}}\tau), when a molecule is uncoupled, and the signal photon counts (mean number n¯g=ηNT(τ)+n¯s\bar{n}_{g}=\eta N_{T}(\tau)+\bar{n}_{s}), when a molecule is in state |g\ket{g}. Here, dark\mathcal{I}_{\mathrm{dark}} is the dark-count rate of a single-photon detector. We define the state detection fidelity for successfully detecting the molecular state as

min{pg(nth),ps(nth)},\mathcal{F}\equiv\mathrm{min}\left\{p_{g}(n_{\mathrm{th}}),p_{s}(n_{\mathrm{th}})\right\}, (15)

where nthn_{\mathrm{th}} is the threshold photon count that maximizes the fidelity. Here, pg(n)=k=nn¯gken¯g/k!p_{g}(n)=\sum_{k=n}^{\infty}\bar{n}_{g}^{k}e^{-\bar{n}_{g}}/k! and ps(n)=k=0nn¯sken¯s/k!p_{s}(n)=\sum_{k=0}^{n}\bar{n}_{s}^{k}e^{-\bar{n}_{s}}/k!. As shown in Fig. 5(c), with typical experiment parameters (C=50C=50 and η=0.3\eta=0.3), near-unity fidelity of 95% can be achieved with 15\sim 15% depumping probability and NT10N_{T}\approx 10 as in Fig. 5(b).

Lastly, we comment that as large cooperativity C>1C>1 guarantees higher transmitted photon counts, Eq. (14) suggests that it is also possible to detect a molecular state using a transition of a small Frank-Condon factor. Figure 6 shows that the transmitted photon number NT,maxC/2N_{T,\text{max}}\sim C/2 even for fFC1f_{FC}\ll 1. In comparison, the collectable photon number in direct fluorescence imaging is 1/(1fFC)1/(1-f_{FC}) without repumping, subject to finite collection efficiency due to limited solid angle span of imaging instrument.

To sum up, in this section we propose a background-free state detection method, using single WGM coupled to a molecule without a closed transition. An alternative scheme using a single-mode Fabry-Perot cavity has also been investigated. To achieve near background-free detection in a cavity, one should instead monitor the transmissivity at the resonance of a molecule-cavity dressed state. A fiber Fabry-Perot cavity, for example, has a record C=145C=145 Colombe et al. (2007) that would serve as an excellent candidate for the proposed scheme. Details of the adaptation are described in Appendix A.

Refer to caption
Figure 7: (a)Simplified energy level structure for one molecule coupled to two WGMs in the weak-driving regime. (b) Time-averaged transmission and (c) reflection spectra for probing the molecule under various time intervals with C=50C=50.

IV Single Molecule Dynamics Coupled with Two Resoantor Modes

In the previous section, we consider single-mode interaction with a spin-polarized molecule. In general, a spin unpolarized molecule can couple to both CW and CCW WGMs. We now consider a general case that a molecule can couple to both modes. We continue to assume negligible back scattering or mode mixing to simplify the discussion. Here, the two modes are degenerate and we model the empty resonator with a two-mode Hamiltonian

H^0=\displaystyle\hat{H}^{\prime}_{0}= Δcl(a^CWa^CW+a^CCWa^CCW)\displaystyle\Delta_{cl}(\hat{a}_{CW}^{\dagger}\hat{a}_{CW}+\hat{a}_{CCW}^{\dagger}\hat{a}_{CCW}) (16)
+i(εa^CWεa^CW)\displaystyle+i\left(\varepsilon\hat{a}^{\dagger}_{CW}-\varepsilon^{\ast}\hat{a}_{CW}\right)

where a^CW()\hat{a}_{CW}^{({\dagger})} and a^CCW()\hat{a}_{CCW}^{({\dagger})} are annihilation (creation) operators for the CW and CCW modes, respectively, and we consider the input field in the bus waveguide excites only the CW mode as in Fig. 1(a).

To illustrate the key signature of molecule-WGM-bus waveguide coupling while keeping the calculation tractable, we continue to use a simple two-level structure to effectively model an unpolarized molecule equally coupled to the CW and CCW modes 222While the hyperfine and magnetic sub-level structure should be taken into account in a realistic model, the example given here assumes a simple spherical dipole coupled to the WGMs as an approximation of an unpolarized molecule interacting with the electric field of WGMs.. We write down the two-mode light-molecule interaction Hamiltonian

H^1=\displaystyle\hat{H}^{\prime}_{1}= Δmlσ^+σ^+igCW(a^CWσ^a^CWσ^+)\displaystyle\Delta_{ml}\hat{\sigma}_{+}\hat{\sigma}_{-}+ig_{CW}\left(\hat{a}^{\dagger}_{CW}\hat{\sigma}_{-}-\hat{a}_{CW}\hat{\sigma}_{+}\right) (17)
+igCCW(a^CCWσ^a^CCWσ^+),\displaystyle\ +ig_{CCW}\left(\hat{a}^{\dagger}_{CCW}\hat{\sigma}_{-}-\hat{a}_{CCW}\hat{\sigma}_{+}\right),

and assume equal coupling strength with the two modes gCW=gCCW=gcg_{CW}=g_{CCW}=g_{c}. The master equation of the full system is then modified to be

dρdt=\displaystyle\frac{d\rho}{dt}= i[H^0+H^1,ρ]+2κ[a^CW]ρ\displaystyle-i\left[\hat{H}^{\prime}_{0}+\hat{H}^{\prime}_{1},\rho\right]+2\kappa\mathcal{L}[\hat{a}_{CW}]\rho (18)
+2κ[a^CCW]ρ+Γg[σ^]ρ+Γs[σ^]ρ,\displaystyle+2\kappa\mathcal{L}[\hat{a}_{CCW}]\rho+\Gamma_{g}\mathcal{L}[\hat{\sigma}_{-}]\rho+\Gamma_{s}\mathcal{L}[\hat{\sigma_{-}}^{\prime}]\rho,

where the two WGMs are assumed to have the same intrinsic loss rates κi\kappa_{i} and bus waveguide coupling rates κe\kappa_{e}.

In the limit of single excitation, the resonator and the molecule form an effective six-level system shown in Fig. 7(a). State |g,1CW(CCW)\ket{g,1_{CW(CCW)}} represents the degenerate level with one photon in the CW (CCW) mode and the molecule in |g\ket{g}. The four coupled states on the left of Fig. 7(a) form a cavity QED subsystem with quasi-steady equilibrium, whose population is gradually transferred, via spontaneous decay, to the right part of Fig. 7(a) that evolves like an empty resonator described in Sec. II. The decay rate is similarly described by Eq. (7) except now gc2g_{c}^{2} is replaced by gCW2+gCCW2=2gc2g^{2}_{CW}+g^{2}_{CCW}=2g_{c}^{2}, effectively giving a total cooperativity 2C2C. We now have

D(δ)=gc2κΓs|2gc2+(iδ+κ)(iδ+Γ2)|2,D^{\prime}(\delta)=\frac{g_{c}^{2}\kappa\mathcal{I}\Gamma_{s}}{\absolutevalue{2g_{c}^{2}+(i\delta+\kappa)\left(i\delta+\frac{\Gamma}{2}\right)}^{2}}, (19)

based on Eq. (41) and assuming δ=Δml=Δcl\delta=\Delta_{\text{ml}}=\Delta_{\text{cl}} and κ=2κe\kappa=2\kappa_{e}. The transfer rate at zero detuning is approximately four times slower than that expected in Eq. (8),

Dres=4C(4C+1)2(1fFC)ΓC1(1fFC)4C,D^{\prime}_{\text{res}}=\frac{4C}{(4C+1)^{2}}(1-f_{FC})\mathcal{I}\Gamma\overset{C\gg 1}{\approx}\frac{(1-f_{FC})\mathcal{I}}{4C}, (20)

due to the increased cooperativity 2C2C.

While waveguide transmission is modified with the presence of a single molecule, there is now also reflection in the bus waveguide due to the molecule-excited CCW resonator field which couples to the bus waveguide in the backward direction relative to the input field. As similarly discussed in Sec. III, we evaluate the time-dependent transmissivity and reflectivity using T=|1+i2κea^CW|2T^{\prime}=\left|1+i\sqrt{\frac{2\kappa_{e}}{\mathcal{I}}}\langle\hat{a}_{CW}\rangle\right|^{2} and R=|i2κea^CCW|2R^{\prime}=\left|i\sqrt{\frac{2\kappa_{e}}{\mathcal{I}}}\langle\hat{a}_{CCW}\rangle\right|^{2} with a^CW\expectationvalue{\hat{a}_{CW}} and a^CCW\expectationvalue{\hat{a}_{CCW}} calculated in Eqs. (42) and (43) respectively. We find

T(δ)=|1κiδ+κ+r|2andR=|r|2,T^{\prime}(\delta)=\absolutevalue{1-\frac{\kappa}{i\delta+\kappa}+r^{\prime}}^{2}~{}\mathrm{and}~{}R^{\prime}=\absolutevalue{r^{\prime}}^{2}, (21)

where

r=κgc2eDt(iδ+κ)[2gc2+(iδ+κ)(iδ+Γ2)].r^{\prime}=\frac{\kappa g_{c}^{2}e^{-D^{\prime}t}}{(i\delta+\kappa)\left[2g_{c}^{2}+(i\delta+\kappa)\left(i\delta+\frac{\Gamma}{2}\right)\right]}. (22)

In Fig. 7(b-c), we calculate the time-averaged transmission and reflection spectra at critical coupling, using the definitions similar to Eq. (12) for T¯(δ,τ)\overline{T}^{\prime}(\delta,\tau) and R¯(δ,τ)\overline{R}^{\prime}(\delta,\tau). There are now three absorption dips (reflection peaks) found in T¯(R¯)\overline{T}^{\prime}(\overline{R}^{\prime}), resulting from the resonances of three eigenstates within the single excitation subspace of the Hamiltonian H0+H1H_{0}^{\prime}+H_{1}^{\prime}. Compared to Fig. 4(b), the molecule-induced transparency and finite reflectivity near δ=0\delta=0, albeit with much reduced contrast, continue to provide ideal background-free signal.

At δ=0\delta=0, we find equal transmissivity and reflectivity,

T(0)=R(0)\displaystyle T^{\prime}(0)=R^{\prime}(0) =|C2C+12|2e2Drest\displaystyle=\absolutevalue{\frac{C}{2C+\frac{1}{2}}}^{2}e^{-2D^{\prime}_{\text{res}}t} (23)
C114exp((1fFC)t2C),\displaystyle\overset{C\gg 1}{\approx}\frac{1}{4}\exp(-\frac{(1-f_{FC})\mathcal{I}t}{2C}),

where the peak values at t=0t=0 is reduced to 25%\approx 25\% of the maximum transmission for the single mode case Eq. (11). The reduced total bus waveguide output, T(0)+R(0)0.5<1T^{\prime}(0)+R^{\prime}(0)\lesssim 0.5<1, is due to the excitation of a pure photonic eigenmode that dissipates through the intrinsic resonator loss. Nevertheless, the transfer rate DresD^{\prime}_{\textrm{res}} in T(0)(R(0))T^{\prime}(0)\ (R^{\prime}(0)) is also smaller by four times resulting from the collective coupling strength of two modes. As a result, at zero detuning the integrated transmitted (reflected) photon number NT(τ)N_{T^{\prime}}(\tau) (NR(τ)N_{R^{\prime}}(\tau)), defined as in Eq. (13), still leads to the same maximum counts NT,max=NR,max=NT,maxN_{T^{\prime},\text{max}}=N_{R^{\prime},\text{max}}=N_{T,\text{max}} as shown in Figs. 8 and 5(a).

Refer to caption
Figure 8: (color online) Transmitted and reflected photon number (NT(τ)=NR(τ)N_{T^{\prime}}(\tau)=N_{R^{\prime}}(\tau)) versus input photon number τ\mathcal{I}\tau, integrated over time interval (0,τ)(0,\tau), and calculated using C=1C=1 (black), 25 (red), 50 (blue), 75 (green), and 100 (purple), respectively (solid curves from bottom to top). Dashed line represents the upper bound NT=NR=τN_{T^{\prime}}=N_{R^{\prime}}=\mathcal{I}\tau. Inset displays NT,maxNT()N_{T^{\prime},\text{max}}\equiv N_{T^{\prime}}(\infty).

As suggested from the discussions above, the fidelity for molecular state measurement, using either tranmission or reflection signal alone, remains identical to the case of coupling to a single mode as shown in Fig. 5(c). Moreover, simultaneous detection of bus waveguide transmission and reflection can offer superior sensitivity, with one time more signal photons and with non-trivial temporal correlation between the transmitted and reflected photons when one exploits quantum nonlinearity in the molecule-WGM interactions.

Refer to caption
Figure 9: (color online)(a) Effective cascade model describing the population dynamics. When interacting with a weakly-driven resonator, ground state molecules (filled circles) are slowly depumped into the uncoupled states (open circles) one-by-one. pnp_{n} marks the probability of having nn remaining molecules coupled to the resonator and DnD_{n} is the transfer rate from nn to n1n-1 molecule manifold. (b) State detection fidelity \mathcal{F} versus ground state population loss PsP_{s}, calculated using C=10C=10 and the initial ground state molecule numbers N=1N=1, 2, 5, and 10, 25, respectively (solid curves from bottom to top).

V multiple molecule dynamics

In the previous sections, we assume large cooperativity (C1C\gg 1) for single molecule detection. In this section, we discuss the case when more than two molecules in the same state are present in the system. Even with a small cooperativity, one may take advantage of collective effects to achieve state detection with high-fidelity.

We begin with NN ground state molecules trapped on a micro-ring resonator, interacting with a single WGM with identical coupling strength gcg_{c}. We consider the resonant case δ=0\delta=0, and write down the light-molecule interaction Hamiltonian as

H^N=α=1Nigc(a^σ^αeiϕαa^σ^+αeiϕα),\displaystyle\hat{H}_{N}=\sum^{N}_{\alpha=1}ig_{c}\left(\hat{a}^{\dagger}\hat{\sigma}^{\alpha}_{-}e^{-i\phi_{\alpha}}-\hat{a}\hat{\sigma}^{\alpha}_{+}e^{i\phi_{\alpha}}\right), (24)

where the index α\alpha labels individual molecules and ϕα\phi_{\alpha} represents a position-dependent phase in the light-molecule coupling since a WGM is a traveling wave.

We incorporate a modified Dicke model to investigate the collective behavior of these resonator-coupled molecules. Let J^±N=1Nα=1Nσ^±αe±iϕα\displaystyle\hat{J}^{N}_{\pm}=\frac{1}{\sqrt{N}}\sum^{N}_{\alpha=1}\hat{\sigma}^{\alpha}_{\pm}e^{\pm i\phi_{\alpha}} be the collective spin lowering and raising operators, we can rewrite the equivalent Hamiltonian for Eq. (24) as

H^JN=iNgc(a^J^Na^J^+N),\hat{H}_{J_{N}}=i\sqrt{N}g_{c}\left(\hat{a}^{\dagger}\hat{J}^{N}_{-}-\hat{a}\hat{J}^{N}_{+}\right), (25)

where the molecule-resonator coupling strength is replaced by Ngc\sqrt{N}g_{c}.

In the limit of single excitation, the resonator-coupled NN molecules resemble an effective two-level system with a NN-molecule ground state |gNΠα=1N|gα\ket{g_{N}}\equiv\Pi^{N}_{\alpha=1}\ket{g}_{\alpha} and an excited state |eNJ^+N|gN\ket{e_{N}}\equiv\hat{J}^{N}_{+}\ket{g_{N}}, in which one molecule gets excited to |e\ket{e}. Spontaneous decay (via emitting single photon into freespace) can either bring the population in |eN\ket{e_{N}} back to |gN\ket{g_{N}} or bring one excited state molecule down to the uncoupled state |s\ket{s}. The full master equation for the resonator-coupled NN-molecule density matrix ρN\rho_{N} can be expressed as

dρNdt=\displaystyle\frac{d\rho_{N}}{dt}= i[H^0+H^JN,ρN]+Γg[J^N]ρN\displaystyle-i[\hat{H}_{0}+\hat{H}_{J_{N}},\rho_{N}]+\Gamma_{g}\mathcal{L}\left[\hat{J}^{N}_{-}\right]\rho_{N} (26)
+2κ[a^]ρN+Γsα=1N[σ^αeiϕα]ρN,\displaystyle+2\kappa\mathcal{L}[\hat{a}]\rho_{N}+\Gamma_{s}\sum_{\alpha=1}^{N}\mathcal{L}\left[\hat{\sigma}^{\prime\alpha}_{-}e^{-i\phi_{\alpha}}\right]\rho_{N},

where the first line of the equation describes the evolution of the NN-molecule collective states with coupling strength Ngc\sqrt{N}g_{c} and decay rate Γg\Gamma_{g}, and the second line adds resonator dissipation and single molecular decay into |s\ket{s}.

While full evolution dynamics of Eq. (26) can be evaluated numerically, calculation for large NN can be computationally expansive. Here, we develop an analytical approximation in the weak-driving limit. Starting with NN molecules in the ground state |gN\ket{g_{N}} weakly excited to |eN\ket{e_{N}} by the resonator mode, the system evolves collectively similar to the single molecule case in Fig. 2(a). Within this NN-coupled molecule manifold, the system can be described by a simple three level system consisting of |gN,0\ket{g_{N},0}, |gN,1\ket{g_{N},1} and |eN,0\ket{e_{N},0} until spontaneous decay into state |s\ket{s} occurs. If we trace out the molecule that decays into the uncoupled state, not knowing which one did, the system can be described again by a collective state |gN1\ket{g_{N-1}} with N1N-1 molecules in the ground state, weakly excited to |eN1\ket{e_{N-1}}, as detailed in Appendix D. The dynamics can cascade down as prescribed with N2,N3,,1N-2,N-3,...,1 molecule(s) left in the system until all molecules become uncoupled, as illustrated in Fig. 9 (a).

We can calculate the probability pnp_{n} of having nn coupled molecules in the system, using the simple cascade model. This leads to a system of equations,

{dpNdt=pNDNdpndt=pn+1Dn+1pnDn, 1<n<Ndp0dt=p1D1,\begin{cases}\displaystyle\frac{dp_{N}}{dt}=-p_{N}D_{N}\\ \displaystyle\frac{dp_{n}}{dt}=p_{n+1}D_{n+1}-p_{n}D_{n},\ 1<n<N\\ \displaystyle\frac{dp_{0}}{dt}=p_{1}D_{1},\end{cases} (27)

where p0p_{0} is for all molecules in |s\ket{s}. The effective transfer rate of the population from nn to n1n-1 coupled molecules can be calculated according to Eq. (49). We find

Dn=2ngc2κe|ngc2+κΓ2|2ΓsD_{n}=\frac{2ng_{c}^{2}\kappa_{e}\mathcal{I}}{\absolutevalue{ng_{c}^{2}+\kappa\frac{\Gamma}{2}}^{2}}\Gamma_{s} (28)

as detailed in Appendix D. For n1n\geq 1, the transfer rate is suppressed by the cooperativity 1/nC\sim 1/nC. We note that Eq. (7) (with δ=0\delta=0) is the single molecule case of Eq. (28).

We can also derive the bus waveguide transmission by finding the expectation value for the resonator field,

a^=n=0NΓ2ngc2+κΓ2pnε.\displaystyle\expectationvalue{\hat{a}}=\sum_{n=0}^{N}\frac{\frac{\Gamma}{2}}{ng_{c}^{2}+\kappa\frac{\Gamma}{2}}p_{n}\varepsilon. (29)

Bus waveguide transmissivity TNT_{N} for initially NN ground state molecules can then be evaluated using Eq. (2).

We have compared the analytical solutions with numerical calculations for mean populations and photon numbers, and found very good agreement for N=2,3,4N=2,3,4. Equations (27-29) allow us to evaluate the dynamics of waveguide transmission with arbitrarily large number of coupled molecules.

The major advantage for coupling more than one molecules to a resonator is that the collective coupling leads to many more signal photons and higher fidelity without losing significant fraction of ground state molecules to the uncoupled states. To illustrate this, in Fig. 9(b) we calculate the state detection fidelity \mathcal{F} as a function of the ground state molecule loss

Ps=11Nn=1Nnpn.P_{s}=1-\frac{1}{N}\sum_{n=1}^{N}np_{n}. (30)

It is shown that, under a moderate C=10C=10, N=10N=10 ground state molecules can be detected with over 99% fidelity with 1% ground state population loss. For even larger NN, non-destructive state detection with negligible loss can be realized with cooperavity parameter C<0.1C<0.1, as described in Ref. Sawant et al. (2018).

VI Conclusion

To conclude, we have proposed a background-free state detection scheme for single molecules without an optically closed transition. High-fidelity measurement can be realized in resonators with a sufficiently large cooperativity C>10C>10. A possible experiment with cold molecules could begin with an array of cold atoms trapped in optical tweezers Thompson et al. (2013); Kim et al. (2019) or in a lattice of evanescent field traps above the surface of a high-Q micro-ring resonator or a photonic crystal cavity of C25C\gtrsim 25 Chang et al. (2019); Samutpraphoot et al. (2020). Resonator-assisted photoassociation (PA) to a molecular ground state (also with high fidelity) can be performed by simply introducing PA light into the experimental setup Perez-Rios et al. (2017); Kampschulte and Denschlag (2018). Immediately following PA, one can detect the existence of ground state molecules using the proposed scheme with probe photons directly launched into the bus waveguide. This state detection technique could also be employed for atoms Kim et al. (2019) or quantum emitters coupled to a high-Q micro-ring resonator or other whispering-gallery mode resonators.

Acknowledgements.
We thank J. Pérez-Ríos for discussions. Funding is provided by the Office of Naval Research (N00014-17-1-2289). M. Zhu acknowledges support from the Rolf Scharenberg Graduate Fellowship. C.-L. Hung acknowledges support from the AFOSR YIP (FA9550-17-1-0298).

Appendix A Alternative setup using a Fabry-Perot cavity

Fabry-Perot cavity is widely used in the investigation of cavity QED. In this section we discuss a similar way to detect a molecule with no optically closed transitions. The system Hamiltonian is similarly described by Eq. (1), except now the driving field amplitude ε=2κl\varepsilon=\sqrt{2\kappa_{l}\mathcal{I}}, where κl(r)\kappa_{l(r)} is the effective loss rate due to transmission through the left (right) mirror, as illustrated in Fig. A1(a) and κ=κi+κr+κl\kappa=\kappa_{i}+\kappa_{r}+\kappa_{l} is the cavity total decay rate.

Refer to caption
Figure A1: (a) Schematic illustration of the coupling between a Fabry-Perot cavity and a molecule. (b) Transmission spectra with (solid line) and without a ground state molecule coupled to the cavity(dashed line).

Figure A1(b) displays the transmission spectra with and without a molecule, assuming a short interrogation time, and with the parameters listed in Table 1 where κr=κl=2π×25MHz\kappa_{r}=\kappa_{l}=2\pi\times 25\text{MHz}. In the cavity setup, transmissivity is T=|2κra^|2\displaystyle T=\absolutevalue{\sqrt{\frac{2\kappa_{r}}{\mathcal{I}}}\expectationvalue{\hat{a}}}^{2} , where the time evolution of a^\expectationvalue{\hat{a}} can be similarly evaluated as in Sec. III. When a molecule in the target state is coupled to the cavity, transmission around the resonance splits into two peaks due to the vacuum Rabi splitting. One could thus monitor the cavity transmission at δ=±gc\delta=\pm g_{c}, where the transmissivity increases by more than ten-fold, to perform nearly background-free measurement.

Appendix B Derivation for the quasi-steady state density matrix in the case of one resonator mode

In the weak-driving regime, we derive the effective decay rate from the cavity QED subsystem to the empty resonator states as shown in Fig. 3(a). Here, we determine the elements of the quasi-steady density matrix ρss\rho^{ss}. We introduce an arbitrarily slow artificial repump of rate ζ\zeta between |s,0\ket{s,0} (|s,1\ket{s,1}) and |g,0\ket{g,0} Kampschulte and Denschlag (2018). For convenience, we define |1|g,0\ket{1}\equiv\ket{g,0}, |2|e,0\ket{2}\equiv\ket{e,0}, |3|g,1\ket{3}\equiv\ket{g,1}, |4|s,0\ket{4}\equiv\ket{s,0}, |5|s,1\ket{5}\equiv\ket{s,1}, and σ^ij|ij|\hat{\sigma}_{ij}\equiv\ket{i}\bra{j}. Then the Hamiltonian of the full system with artificial repump is

H^R=H^0+H^1+ζ(σ^41+σ^14+σ^51+σ^15)\hat{H}_{R}=\hat{H}_{0}+\hat{H}_{1}+\zeta\left(\hat{\sigma}_{41}+\hat{\sigma}_{14}+\hat{\sigma}_{51}+\hat{\sigma}_{15}\right) (31)

where H^0\hat{H}_{0} is the Hamiltonian of empty resoantor as in Eq. (1) and H^1\hat{H}_{1} is the single-mode light-molecule interaction Hamiltonian as in Eq. (3). Taking into account the loss channels, we write down the master equation

dρdt=i[H^R,ρ]+2κ[a^]ρ+Γg[σ12^]ρ+Γs[σ42^]ρ.\frac{d\rho}{dt}=-i\left[\hat{H}_{R},\rho\right]+2\kappa\mathcal{L}[\hat{a}]\rho+\Gamma_{g}\mathcal{L}[\hat{\sigma_{12}}]\rho+\Gamma_{s}\mathcal{L}[\hat{\sigma_{42}}]\rho. (32)

We first focus on the evolution of the density matrix elements to the leading order of |ε|\absolutevalue{\varepsilon} (assuming ζε\zeta\ll\varepsilon). We define ρij=i|ρ^|j\rho_{ij}=\bra{i}\hat{\rho}\ket{j}, and find the evolutions of for ρ21\rho_{21} and ρ31\rho_{31} satisfy

dρ21dt\displaystyle\frac{d\rho_{21}}{dt} =(iΔml+Γ2)ρ21gcρ31,\displaystyle=-\ (i\Delta_{\text{ml}}+\frac{\Gamma}{2})\rho_{21}-g_{c}\rho_{31}, (33)
dρ31dt\displaystyle\frac{d\rho_{31}}{dt} =(iΔcl+κ)ρ31+ερ11+gcρ21,\displaystyle=-\ (i\Delta_{\text{cl}}+\kappa)\rho_{31}+\varepsilon\rho_{11}+g_{c}\rho_{21},

which is independent of ζ\zeta as well as the empty resonator states |4\ket{4} and |5\ket{5}. Solving for the quasi-steady state, dρ31dtdρ21dt0\displaystyle\frac{d\rho_{31}}{dt}\approx\frac{d\rho_{21}}{dt}\approx 0, we find

ρ21ssρ11ss\displaystyle\frac{\rho_{21}^{ss}}{\rho_{11}^{ss}} =gcgc2+(iΔcl+κ)(iΔml+Γ2)ε\displaystyle=\frac{-g_{c}}{g_{c}^{2}+(i\Delta_{\text{cl}}+\kappa)\left(i\Delta_{\text{ml}}+\frac{\Gamma}{2}\right)}\varepsilon (34)
ρ31ssρ11ss\displaystyle\frac{\rho_{31}^{ss}}{\rho_{11}^{ss}} =iΔml+Γ2gc2+(iΔcl+κ)(iΔml+Γ2)ε.\displaystyle=\frac{i\Delta_{\text{ml}}+\frac{\Gamma}{2}}{g_{c}^{2}+(i\Delta_{\text{cl}}+\kappa)\left(i\Delta_{\text{ml}}+\frac{\Gamma}{2}\right)}\varepsilon. (35)

We similarly derive the components of ρss\rho^{ss} to the next order of |ε|\absolutevalue{\varepsilon},

ρ22ssρ11ss=2gc2κe|gc2+(iΔcl+κ)(iΔml+Γ2)|2\frac{\rho_{22}^{ss}}{\rho_{11}^{ss}}=\frac{2g_{c}^{2}\kappa_{e}\mathcal{I}}{\absolutevalue{g_{c}^{2}+(i\Delta_{\text{cl}}+\kappa)\left(i\Delta_{\text{ml}}+\frac{\Gamma}{2}\right)}^{2}} (36)
ρ33ssρ11ss=2((Γ/2)2+Δml2)κe|gc2+(iΔcl+κ)(iΔml+Γ2)|2.\frac{\rho_{33}^{ss}}{\rho_{11}^{ss}}=\frac{2\left((\Gamma/2)^{2}+\Delta_{\text{ml}}^{2}\right)\kappa_{e}\mathcal{I}}{\absolutevalue{g_{c}^{2}+(i\Delta_{\text{cl}}+\kappa)\left(i\Delta_{\text{ml}}+\frac{\Gamma}{2}\right)}^{2}}. (37)

where we have used ε=i2κe\varepsilon=i\sqrt{2\kappa_{e}\mathcal{I}}.

Under weak-driving, the initial population in the ground state |g\ket{g} is gradually transferred to states |s\ket{s} via the spontaneous decay channel |2|4\ket{2}\rightarrow\ket{4}. From Eq. (36), we obtain the population ratio between states |e,0\ket{e,0} and |g,0\ket{g,0} and find the effective transfer rate DD

D=ρ22ssρ11ssΓs=2gc2κe|gc2+(iΔcl+κ)(iΔml+Γ2)|2Γs,D=\frac{\rho_{22}^{ss}}{\rho_{11}^{ss}}\Gamma_{s}=\frac{2g_{c}^{2}\kappa_{e}\mathcal{I}}{\absolutevalue{g_{c}^{2}+(i\Delta_{\text{cl}}+\kappa)\left(i\Delta_{\text{ml}}+\frac{\Gamma}{2}\right)}^{2}}\Gamma_{s}, (38)

and Eq. (7) for the case at critical coupling and Δcl=Δml\Delta_{\text{cl}}=\Delta_{\text{ml}}. Considering initially the system begins in |g,0\ket{g,0} and most of the population resides in either state |1\ket{1} (|g,0\ket{g,0}) or |4\ket{4} (|s,0\ket{s,0}), we find ρ11(t)eDt\rho_{11}(t)\approx e^{-Dt}, ρ44(t)1eDt\rho_{44}(t)\approx 1-e^{-Dt}, and hence Eq. (9). The dynamics of all other components in ρ\rho can be solved once the populations of |1\ket{1} and |4\ket{4} are known.

We can solve the expectation value a\expectationvalue{a} for calculating transmissivity in Eq. (2). We find

a^=\displaystyle\expectationvalue{\hat{a}}= ρ31+ρ54\displaystyle\rho_{31}+\rho_{54} (39)
=\displaystyle= iΔml+Γ2gc2+(iΔcl+κ)(iΔml+Γ2)εeDt\displaystyle\frac{i\Delta_{\text{ml}}+\frac{\Gamma}{2}}{g_{c}^{2}+(i\Delta_{\text{cl}}+\kappa)\left(i\Delta_{\text{ml}}+\frac{\Gamma}{2}\right)}\varepsilon e^{-Dt}
+εiΔcl+κ(1eDt).\displaystyle+\frac{\varepsilon}{i\Delta_{\text{cl}}+\kappa}(1-e^{-Dt}).

We can also calculate the dynamics of resonator photon number a^a^\expectationvalue{\hat{a}^{{\dagger}}\hat{a}}, using the populations of one photon states

a^a^=\displaystyle\expectationvalue{\hat{a}^{{\dagger}}\hat{a}}= ρ33+ρ55\displaystyle\rho_{33}+\rho_{55} (40)
=\displaystyle= 2((Γ/2)2+Δml2)κe|gc2+(iΔcl+κ)(iΔml+Γ2)|2eDt\displaystyle\frac{2\left((\Gamma/2)^{2}+\Delta_{\text{ml}}^{2}\right)\kappa_{e}\mathcal{I}}{\absolutevalue{g_{c}^{2}+(i\Delta_{\text{cl}}+\kappa)\left(i\Delta_{\text{ml}}+\frac{\Gamma}{2}\right)}^{2}}e^{-Dt}
+2κeΔcl2+κ2(1eDt).\displaystyle+\frac{2\kappa_{e}\mathcal{I}}{\Delta_{\text{cl}}^{2}+\kappa^{2}}(1-e^{-Dt}).

Figure 3 validates the analytical approximation with full numerical calculations in the weak driving regime.

Appendix C Derivation for the quasi-steady state density matrix in the case of two resonator modes

The same scenario used in Appendix B can be applied in the case of coupling to two resonator modes. We can evaluate the transfer rate to the empty resonator state DD^{\prime} and the expectation value of aCW\expectationvalue{a_{CW}} and aCCW\expectationvalue{a_{CCW}} in the weak-driving regime. Defining |1|g,0\ket{1}\equiv\ket{g,0}, |2|e,0\ket{2}\equiv\ket{e,0}, |3|g,1CW\ket{3}\equiv\ket{g,1_{CW}}, |4|g,1CCW\ket{4}\equiv\ket{g,1_{CCW}}, |5|s,0\ket{5}\equiv\ket{s,0} and |6=|s,1\ket{6}=\ket{s,1}, we obtain

ρ22ssρ11ss=2gc2κe|2gc2+(iΔcl+κ)(iΔml+Γ2)|2\frac{\rho_{22}^{ss}}{\rho_{11}^{ss}}=\frac{2g_{c}^{2}\kappa_{e}\mathcal{I}}{\absolutevalue{2g_{c}^{2}+(i\Delta_{\text{cl}}+\kappa)\left(i\Delta_{\text{ml}}+\frac{\Gamma}{2}\right)}^{2}} (41)

and thus the transfer rate DD^{\prime} in Eq. (19).

Similarly, we find ρ11(t)=eDt\rho_{11}(t)=e^{-D^{\prime}t}, ρ44(t)=1eDt\rho_{44}(t)=1-e^{-D^{\prime}t},

a^CW=\displaystyle\expectationvalue{\hat{a}_{CW}}= ρ31+ρ65\displaystyle\rho_{31}+\rho_{65} (42)
=\displaystyle= gc2+(iΔcl+κ)(iΔml+Γ2)(iΔcl+κ)[2gc2+(iΔcl+κ)(iΔml+Γ2)]εeDt\displaystyle\frac{g_{c}^{2}+(i\Delta_{\text{cl}}+\kappa)\left(i\Delta_{\text{ml}}+\frac{\Gamma}{2}\right)}{(i\Delta_{\text{cl}}+\kappa)\left[2g_{c}^{2}+(i\Delta_{\text{cl}}+\kappa)\left(i\Delta_{\text{ml}}+\frac{\Gamma}{2}\right)\right]}\varepsilon e^{-D^{\prime}t}
+εiΔcl+κ(1eDt),\displaystyle+\frac{\varepsilon}{i\Delta_{\text{cl}}+\kappa}(1-e^{-D^{\prime}t}),
a^CCW\displaystyle\expectationvalue{\hat{a}_{CCW}} =ρ41\displaystyle=\rho_{41} (43)
=gc2(iΔcl+κ)[2gc2+(iΔcl+κ)(iΔml+Γ2)]εeDt.\displaystyle=\frac{-g_{c}^{2}}{(i\Delta_{\text{cl}}+\kappa)\left[2g_{c}^{2}+(i\Delta_{\text{cl}}+\kappa)\left(i\Delta_{\text{ml}}+\frac{\Gamma}{2}\right)\right]}\varepsilon e^{-D^{\prime}t}.

Substituting a^CW\expectationvalue{\hat{a}_{CW}} and a^CCW\expectationvalue{\hat{a}_{CCW}} in the expressions of TT^{\prime} and RR^{\prime} leads to Eqs. (21) and (22).

Appendix D Cascade model for multiple molecules resonantly coupled to one resonator mode

In the main text, we consider multiple molecules collectively couple to one resonator mode by directly tracing out all the position dependence. Here, we explicitly carry out the derivation and arrive at the cascade model described by Eq. (27).

We assume NN trapped molecules randomly spread along the micro-ring resonator. We denote |gn,k\ket{g_{n,k}} as the kk-th configuration that satisfies NnN-n molecules in state |s\ket{s} and nn molecules in state |g\ket{g}, and we define GkG_{k} to be the set of positions labeling these nn molecules. We assume every molecule in the ground state can be equally excited by the WGM. Thus, with single excitation created in a given configuration |gn,k\ket{g_{n,k}}, the system forms a superposition state |en,k,01nαGkσ^+αeiϕα|gn,k\ket{e_{n,k},0}\equiv\frac{1}{\sqrt{n}}\sum_{\alpha\in G_{k}}\hat{\sigma}^{\alpha}_{+}e^{i\phi_{\alpha}}\ket{g_{n,k}}. Each excited state molecule might decay to |s\ket{s}, thus |en,k,0\ket{e_{n,k},0} can evolve into nn orthogonal configurations of n1n-1 molecules in the ground state, |gn1,k,0=|sαgα|gn,k,0\ket{g_{n-1,k^{\prime}},0}=\ket{s_{\alpha}}\bra{g_{\alpha}}\ket{g_{n,k},0}, and the superposition is destroyed. Here, αGk\alpha\in G_{k} labels the position of the molecule that decays into |s\ket{s}, and we denote |gn1,k\ket{g_{n-1,k^{\prime}}} as the kk^{\prime}-th configuration that n1n-1 molecules are in state |g\ket{g}.

We note that the evolution dynamics of each position configuration is identical, since the major differences between the states |en,k,0\ket{e_{n,k},0} are the position-dependent phases that can be absorbed as a part of the spin lowering and raising operators defined in Sec. V. If we ignore the position information, one can trace out all different configurations without losing relevant physical information. For simplicity, we denote the states in analogy to the states in Appendix B,

|1n,k\displaystyle\ket{1_{n,k}} |gn,k,0\displaystyle\equiv\ket{g_{n,k},0} (44)
|2n,k\displaystyle\ket{2_{n,k}} |en,k,0\displaystyle\equiv\ket{e_{n,k},0}
|3n,k\displaystyle\ket{3_{n,k}} |gn,k,1\displaystyle\equiv\ket{g_{n,k},1}
|4n,k\displaystyle\ket{4_{n,k^{\prime}}} |1n1,k\displaystyle\equiv\ket{1_{n-1,k^{\prime}}}

where kk and kk^{\prime} are the indices of configurations for nn and n1n-1 molecules coupled to the resonator, respectively. We trace out all the kk configurations in the density matrix elements by the following summation

ρin,jn\displaystyle\rho_{i_{n},j_{n}} kρin,k,jn,k,\displaystyle\equiv\sum_{k}\rho_{i_{n,k},j_{n,k}}, (45)

for i,j=1,2,3i,j=1,2,3 and ρin,k,jn,kin,k|ρ^|jn,k\rho_{i_{n,k},j_{n,k}}\equiv\bra{i_{n,k}}\hat{\rho}\ket{j_{n,k}}. When states |4n,k\ket{4_{n,k^{\prime}}} are involved, we additionally trace out all kk^{\prime} configurations

ρin,4n\displaystyle\rho_{i_{n},4_{n}} k,kρin,k,4n,k,\displaystyle\equiv\sum_{k,k^{\prime}}\rho_{i_{n,k},4_{n,k^{\prime}}}, (46)

and similar definitions for ρ4n,in\rho_{4_{n},i_{n}} and ρ4n,4n\rho_{4_{n},4_{n}} follow. We note that ρ4n,4n=ρ1n1,1n1\rho_{4_{n},4_{n}}=\rho_{1_{n-1},1_{n-1}} for n1n\geq 1.

These density matrix elements evolve similar to ρij\rho_{ij} of one molecule coupled to one WGM. For example, we find

dρ2n,1ndt\displaystyle\frac{d\rho_{2_{n},1_{n}}}{dt} =(iΔml+Γ2)ρ2n,1nngcρ3n,1n,\displaystyle=-\ (i\Delta_{\text{ml}}+\frac{\Gamma}{2})\rho_{2_{n},1_{n}}-\sqrt{n}g_{c}\rho_{3_{n},1_{n}}, (47)
dρ3n,1ndt\displaystyle\frac{d\rho_{3_{n},1_{n}}}{dt} =(iΔcl+κ)ρ3n,1n+ερ1n,1n+ngcρ2n,1n,\displaystyle=-\ (i\Delta_{\text{cl}}+\kappa)\rho_{3_{n},1_{n}}+\varepsilon\rho_{1_{n},1_{n}}+\sqrt{n}g_{c}\rho_{2_{n},1_{n}},

sharing the same form as those in Eq. (33) except that we replace gcg_{c} by ngc\sqrt{n}g_{c}. Following the procedures in Appendix B, we obtain the ratio of quasi-steady state population at zero detuning

ρgn1,gn0ρgn0,gn0ρ3n,1nρ1n,1n\displaystyle\frac{\rho_{g_{n}1,g_{n}0}}{\rho_{g_{n}0,g_{n}0}}\equiv\frac{\rho_{3_{n},1_{n}}}{\rho_{1_{n},1_{n}}} =Γ2εngc2+κΓ2\displaystyle=\frac{\frac{\Gamma}{2}\varepsilon}{ng_{c}^{2}+\kappa\frac{\Gamma}{2}} (48)
ρen0,en0ρgn0,gn0ρ2n,1nρ1n,1n\displaystyle\frac{\rho_{e_{n}0,e_{n}0}}{\rho_{g_{n}0,g_{n}0}}\equiv\frac{\rho_{2_{n},1_{n}}}{\rho_{1_{n},1_{n}}} =2ngc2κe|ngc2+κΓ2|2,\displaystyle=\frac{2ng_{c}^{2}\kappa_{e}\mathcal{I}}{\absolutevalue{ng_{c}^{2}+\kappa\frac{\Gamma}{2}}^{2}}, (49)

where we have converted the subscripts of the density matrix elements to those used in the main text. Given the ground state population ρgn0,gn0ρ1n,1n\rho_{g_{n}0,g_{n}0}\equiv\rho_{1_{n},1_{n}}, the above equations describe the dynamics within each nn-coupled molecule manifold.

Since ρ4n,4n=ρ1n1,1n1\rho_{4_{n},4_{n}}=\rho_{1_{n-1},1_{n-1}}, spontaneous decay to state |s\ket{s} connects the population in every nn-coupled molecule manifold with the populations in the n±1n\pm 1 manifolds. We thus arrive at the cascade decay model Eq. (27). Solving for the population ρgn0,gn0\rho_{g_{n}0,g_{n}0} in each nn-molecule manifold, we can also obtain the dynamics of transmission using the expectation value of the resonator mode field, Eq. (48),

a^=n=0Nρgn1,gn0.\expectationvalue{\hat{a}}=\sum_{n=0}^{N}\rho_{g_{n}1,g_{n}0}. (50)

References

  • Bochmann et al. (2010) J. Bochmann, M. Mücke, C. Guhl, S. Ritter, G. Rempe, and D. L. Moehring, Lossless state detection of single neutral atoms, Physical review letters 104, 203601 (2010).
  • Gehr et al. (2010) R. Gehr, J. Volz, G. Dubois, T. Steinmetz, Y. Colombe, B. L. Lev, R. Long, J. Esteve, and J. Reichel, Cavity-based single atom preparation and high-fidelity hyperfine state readout, Physical review letters 104, 203602 (2010).
  • Fuhrmanek et al. (2011) A. Fuhrmanek, R. Bourgain, Y. R. P. Sortais, and A. Browaeys, Free-space lossless state detection of a single trapped atom, Physical review letters 106, 133003 (2011).
  • Olmschenk et al. (2007) S. Olmschenk, K. C. Younge, D. L. Moehring, D. N. Matsukevich, P. Maunz, and C. Monroe, Manipulation and detection of a trapped yb+ hyperfine qubit, Physical Review A 76, 052314 (2007).
  • Myerson et al. (2008) A. H. Myerson, D. J. Szwer, S. C. Webster, D. T. C. Allcock, M. J. Curtis, G. Imreh, J. A. Sherman, D. N. Stacey, A. M. Steane, and D. M. Lucas, High-fidelity readout of trapped-ion qubits, Physical review letters 100, 200502 (2008).
  • Neumann et al. (2010) P. Neumann, J. Beck, M. Steiner, F. Rempp, H. Fedder, P. R. Hemmer, J. Wrachtrup, and F. Jelezko, Single-shot readout of a single nuclear spin, Science 329, 542 (2010).
  • Robledo et al. (2011) L. Robledo, L. Childress, H. Bernien, B. Hensen, P. F. Alkemade, and R. Hanson, High-fidelity projective read-out of a solid-state spin quantum register, Nature 477, 574 (2011).
  • Vamivakas et al. (2010) A. N. Vamivakas, C. Y. Lu, C. Matthiesen, Y. Zhao, S. Fält, A. Badolato, and M. Atatüre, Observation of spin-dependent quantum jumps via quantum dot resonance fluorescence, Nature 467, 297 (2010).
  • Bohn et al. (2017) J. L. Bohn, A. M. Rey, and J. Ye, Cold molecules: Progress in quantum engineering of chemistry and quantum matter, Science 357, 1002 (2017).
  • DeMille (2002) D. DeMille, Quantum computation with trapped polar molecules, Physical Review Letters 88, 067901 (2002).
  • Shuman et al. (2010) E. S. Shuman, J. F. Barry, and D. DeMille, Laser cooling of a diatomic molecule, Nature 467, 820 (2010).
  • Anderegg et al. (2017) L. Anderegg, B. L. Augenbraun, E. Chae, B. Hemmerling, N. R. Hutzler, A. Ravi, A. Collopy, J. Ye, W. Ketterle, and J. M. Doyle, Radio frequency magneto-optical trapping of caf with high density, Physical review letters 119, 103201 (2017).
  • Gabbanini et al. (2000) C. Gabbanini, A. Fioretti, A. Lucchesini, S. Gozzini, and M. Mazzoni, Cold rubidium molecules formed in a magneto-optical trap, Physical review letters 84, 2814 (2000).
  • Wang et al. (2010) D. Wang, B. Neyenhuis, M. H. G. de Miranda, K.-K. Ni, S. Ospelkaus, D. S. Jin, and J. Ye, Direct absorption imaging of ultracold polar molecules, Physical review A 81, 061404(R) (2010).
  • Lev et al. (2008) B. L. Lev, A. Vukics, E. R. Hudson, B. C. Sawyer, P. Domokos, H. Ritsch, and J. Ye, Prospects for the cavity-assisted laser cooling of molecules, Physical Review A 77, 023402 (2008).
  • Hood et al. (2001) C. J. Hood, H. J. Kimble, and J. Ye, Characterization of high-finesse mirrors: Loss, phase shifts, and mode structure in an optical cavity, Physical review A 64, 033804 (2001).
  • Vernooy et al. (1998) D. Vernooy, V. S. Ilchenko, H. Mabuchi, E. Streed, and H. Kimble, High-Q measurements of fused-silica microspheres in the near infrared, Optics letters 23, 247 (1998).
  • Spillane et al. (2005) S. M. Spillane, T. J. Kippenberg, K. J. Vahala, K. W. Goh, E. Wilcut, and H. J. Kimble, Ultrahigh-Q toroidal microresonators for cavity quantum electrodynamics, Physical review A 71, 013817 (2005).
  • Pöllinger et al. (2009) M. Pöllinger, D. O’Shea, F. Warken, and A. Rauschenbeutel, Ultrahigh-Q tunable whispering-gallery-mode microresonator, Physical review letters 103, 053901 (2009).
  • Tien et al. (2011) M.-C. Tien, J. F. Bauters, M. J. R. Heck, D. T. Spencer, D. J. Blumenthal, and J. E. Bowers, Ultra-high quality factor planar Si3N4 ring resonators on si substrates, Opt. Express 19, 13551 (2011).
  • Chang et al. (2019) T.-H. Chang, B. M. Fields, M. E. Kim, and C.-L. Hung, Microring resonators on a suspended membrane circuit for atom–light interactions, Optica 6, 1203 (2019).
  • Aoki et al. (2006) T. Aoki, B. Dayan, E. Wilcut, W. P. Bowen, A. S. Parkins, T. Kippenberg, K. Vahala, and H. Kimble, Observation of strong coupling between one atom and a monolithic microresonator, Nature 443, 671 (2006).
  • Shomroni et al. (2014) I. Shomroni, S. Rosenblum, Y. Lovsky, O. Bechler, G. Guendelman, and B. Dayan, All-optical routing of single photons by a one-atom switch controlled by a single photon, Science 345, 903 (2014).
  • Boozer et al. (2006) A. D. Boozer, A. Boca, R. Miller, T. E. Northup, and H. J. Kimble, Cooling to the ground state of axial motion for one atom strongly coupled to an optical cavity, Physical review letters 97, 083602 (2006).
  • Khudaverdyan et al. (2009) M. Khudaverdyan, W. Alt, T. Kampschulte, S. Reick, A. Thobe, A. Widera, and D. Meschede, Quantum jumps and spin dynamics of interacting atoms in a strongly coupled atom-cavity system, Physical review letters 103, 123006 (2009).
  • Kato and Aoki (2015) S. Kato and T. Aoki, Strong coupling between a trapped single atom and an all-fiber cavity, Physical review letters 115, 093603 (2015).
  • Thompson et al. (2013) J. D. Thompson, T. Tiecke, N. P. de Leon, J. Feist, A. Akimov, M. Gullans, A. S. Zibrov, V. Vuletić, and M. D. Lukin, Coupling a single trapped atom to a nanoscale optical cavity, Science 340, 1202 (2013).
  • Goban et al. (2014) A. Goban, C.-L. Hung, S.-P. Yu, J. Hood, J. Muniz, J. Lee, M. Martin, A. McClung, K. Choi, D. E. Chang, et al., Atom–light interactions in photonic crystals, Nature communications 5, 3808 (2014).
  • Sun and Waks (2016) S. Sun and E. Waks, Single-shot optical readout of a quantum bit using cavity quantum electrodynamics, Physical Review A 94, 012307 (2016).
  • Bellos et al. (2011) M. Bellos, D. Rahmlow, R. Carollo, J. Banerjee, O. Dulieu, A. Gerdes, E. Eyler, P. Gould, and W. Stwalley, Formation of ultracold Rb2 molecules in the v′′=0v^{\prime\prime}=0 level of the a3Σu+a^{3}\Sigma^{+}_{u} state via blue-detuned photoassociation to the 13Πg1^{3}\Pi_{g} state., Physical chemistry chemical physics: PCCP 13, 18880 (2011).
  • Perez-Rios et al. (2017) J. Perez-Rios, M. E. Kim, and C.-L. Hung, Ultracold molecule assembly with photonic crystals, New Journal of Physics 19, 123035 (2017).
  • Kampschulte and Denschlag (2018) T. Kampschulte and J. H. Denschlag, Cavity-controlled formation of ultracold molecules, New Journal of Physics 20, 123015 (2018).
  • Haus (1984) H. A. Haus, Waves and fields in optoelectronics (Prentice-Hall, 1984).
  • Note (1) The WGMs are circularly polarized since they are traveling wave modes with strong evanescent field outside the waveguide. A ±π/2\pm\pi/2 degree out-of-phase axial field component, dictated by the transversality of the Maxwell’s equation and time reversal symmetry, gives the CW and CCW WGMs opposite circular polarization states.
  • Johansson et al. (2013) J. R. Johansson, P. D. Nation, and F. Nori, Qutip 2: A python framework for the dynamics of open quantum systems, Computer Physics Communications 184, 1234 (2013).
  • Colombe et al. (2007) Y. Colombe, T. Steinmetz, G. Dubois, F. Linke, D. Hunger, and J. Reichel, Strong atom–field coupling for bose–einstein condensates in an optical cavity on a chip, Nature 450, 272 (2007).
  • Note (2) While the hyperfine and magnetic sub-level structure should be taken into account in a realistic model, the example given here assumes a simple spherical dipole coupled to the WGMs as an approximation of an unpolarized molecule interacting with the electric field of WGMs.
  • Sawant et al. (2018) R. Sawant, O. Dulieu, and S. A. Rangwala, Detection of ultracold molecules using an optical cavity, Physical Review A 97, 063405 (2018).
  • Kim et al. (2019) M. E. Kim, T.-H. Chang, B. M. Fields, C.-A. Chen, and C.-L. Hung, Trapping single atoms on a nanophotonic circuit with configurable tweezer lattices, Nature communications 10, 1 (2019).
  • Samutpraphoot et al. (2020) P. Samutpraphoot, T. Đorđević, P. L. Ocola, H. Bernien, C. Senko, V. Vuletić, and M. D. Lukin, Strong coupling of two individually controlled atoms via a nanophotonic cavity, Phys. Rev. Lett. 124, 063602 (2020).