Quantum electrodynamics is rapidly finding a set of new applications in thresholdless lasing, photochemistry, and quantum entanglement due to the development of sophisticated patterning techniques to couple nanoscale photonic emitters with photonic and plasmonic cavities. Colloidal and epitaxial semiconductor nanocrystals or quantum dots (QDs) are promising candidates for emitters within these architectures but are dramatically less explored in this role than are molecular emitters. This perspective reviews the basic physics of emitter-cavity coupling in the weak-to-strong coupling regimes, describes common architectures for these systems, and lists possible applications (in particular, photochemistry), with a focus on the advantages and issues associated with using QDs as the emitters.

The fields of photophysics and photochemistry have generally focused on the use of photons to promote molecules to a metastable excited state that accesses interesting relaxation pathways including intersystem crossing, charge transfers,1–3 bond making/breaking,2,4 or isomerizations.5 Quantum electrodynamics (QED), in which a quantized electric field generated either by an optical cavity or by a plasmonic excitation interacts with a photonic emitter such as a molecule or a quantum dot (QD), is emerging as a subfield of these areas. When this interaction is sufficiently strong, it generates a new potential energy surface for the hybridized system. QED is therefore a way to achieve unprecedented control over photochemical reaction pathways and to access new types of photochemical reactivities based on, for example, creation of conical intersections and enhancement of interemitter energy transfer pathways.6,7

There exists a range of coupling strengths between a cavity-mediated quantized electric field and a photonic emitter from “weak” to “ultrastrong.” In the weak coupling regime [Fig. 1(a)], two possible excited states are present in the system, one in which the emitter is in its excited electronic state (purple trace) and the cavity contains zero photons, and the other in which the emitter is in its ground state (blue trace) and the cavity contains a photon (orange trace). As the coupling between the cavity and the emitter increases, the two excited states hybridize to form polaritonic states [Figs. 1(b) and 1(c)] with potential energy surfaces dependent upon the strength of the coupling, which can be modulated by the properties and geometry of the field and emitter. Theoretical treatment of a molecular photonic emitter coupled strongly to a quantized electric field predicts unprecedented selectivity for a single stereochemical product of a photodriven reaction.8,9 The quantum yield for such a reaction may be greater than 1 due to the ability of multiple individual emitters to couple collectively to one photonic excitation.9 

FIG. 1.

The potential energy surfaces of a quantum emitter in the ground (blue) and excited (orange) states coupled to a photon (purple) in the (a) weak coupling, (b) intermediate coupling, and (c) strong coupling regime. It is only in the strong coupling regime that a hybrid polariton state forms, and this hybrid state shows anticrossing behavior. Reprinted with permission from J. Feist, J. Galego, and F. J. Garcia-Vidal, ACS Photonics 5(1), 205–216 (2018). Copyright 2018 American Chemical Society.

FIG. 1.

The potential energy surfaces of a quantum emitter in the ground (blue) and excited (orange) states coupled to a photon (purple) in the (a) weak coupling, (b) intermediate coupling, and (c) strong coupling regime. It is only in the strong coupling regime that a hybrid polariton state forms, and this hybrid state shows anticrossing behavior. Reprinted with permission from J. Feist, J. Galego, and F. J. Garcia-Vidal, ACS Photonics 5(1), 205–216 (2018). Copyright 2018 American Chemical Society.

Close modal

Zero-dimensional semiconductor nanocrystals, or QDs, which can be deposited from solution or grown epitaxially within a cavity or near plasmonic structures, have some advantages over molecules for use as emitters in QED systems. QDs are exceptionally photostable upon prolonged exposure to high energy and high intensity light sources.10 Spherical QDs have isotropic transition dipole moments11 such that their orientation in space does not affect their ability to couple to a quantized electric field. Anisotropic nanorods and nanoplatelets, which depending on their size are quantum-confined in between one and three dimensions,12–14 must be aligned within a cavity for maximum coupling but have oscillator strengths that are a factor of ten or more higher than analogous spherical particles.15 Previous work has compared QDs to atoms for light-matter coupled systems.16 

In this perspective, we explore the potential of epitaxial and colloidal QDs to serve as emitters within QED systems and highlight some advantages of QD-cavity systems for applications such as photochemistry. This perspective outlines the fundamental physics of these systems, provides a general overview of the field as it currently stands, and shares our outlook on new avenues for the field to explore but does not comprehensively review the field. We first discuss the characteristic spectral changes for an emitter in the weak, intermediate, and strong emitter-field coupling regimes and highlight applications in photochemistry and quantum information. We then discuss experimental realizations of strongly coupled polaritonic systems with QD emitters and discuss the application of such systems to biexciton extraction from QDs, generation of entangled photon pairs, and photochemistry. For the reader’s convenience, we have compiled a glossary of common terms in Table I.

TABLE I.

Glossary of terms.

AnticrossingThe observation that the hybridized states formed in the strong coupling regime remain nondegenerate as the emitter and cavity are brought in and out of resonance with each other.
Bowtie A metal nanogap cavity that is composed of two triangular nanoparticles oriented such that two vertices are in close proximity, resembling a bowtie. 
Bragg reflector A component of optical cavities that has a periodic variation of refractive indices and acts as a reflector. 
Cavity quality factor, Q The ratio of the energy of the cavity to the linewidth, or the rate of energy loss, of the same cavity. Higher energies and narrower linewidths result in improved quality. 
Exciton-photon coupling parameter, g Indicates coupling strength; mathematically defines the strong and weak coupling regimes. 
Fano interference Asymmetric line shapes resulting from the interference of two or more waves. 
Hot-spot The enhanced and localized electromagnetic field formed between plasmonic nanostructures. 
Metal nanogap cavity A plasmonic cavity structure between metallic nanoantennae. The electric field in the cavity is strongly enhanced in the “hot spot” generated from the nanoantennae. 
Micodisk device An optical QED architecture where emitters interact with the modes of a bent waveguide. 
Micropillar device An optical QED architecture where emitters are sandwiched between Bragg reflectors. 
Mode volume, Vm The volume of the quantized electromagnetic field generated within a cavity. 
Nanoantenna A plasmonic nanostructure that transfers electromagnetic energy from near-field to far-field due to its localized surface plasmon resonances. 
Photonic crystal cavity device An optical QED architecture where emitters interact with the modes of periodic dielectric structures. 
Plasmonic nanoresonator A metallic optical device with a periodic structure. 
Purcell effect The enhancement of the rate of spontaneous emission upon the coupling of an emitter to a cavity or a plasmon. 
Rabi splitting The observable splitting of the emission spectrum of an emitter resulting from the nondegenerate hybridized states accessible in the strong coupling regime. 
γC The full width at half maximum (FWHM) of the cavity mode. 
γX The full width at half maximum (FWHM) of the exciton mode. 
AnticrossingThe observation that the hybridized states formed in the strong coupling regime remain nondegenerate as the emitter and cavity are brought in and out of resonance with each other.
Bowtie A metal nanogap cavity that is composed of two triangular nanoparticles oriented such that two vertices are in close proximity, resembling a bowtie. 
Bragg reflector A component of optical cavities that has a periodic variation of refractive indices and acts as a reflector. 
Cavity quality factor, Q The ratio of the energy of the cavity to the linewidth, or the rate of energy loss, of the same cavity. Higher energies and narrower linewidths result in improved quality. 
Exciton-photon coupling parameter, g Indicates coupling strength; mathematically defines the strong and weak coupling regimes. 
Fano interference Asymmetric line shapes resulting from the interference of two or more waves. 
Hot-spot The enhanced and localized electromagnetic field formed between plasmonic nanostructures. 
Metal nanogap cavity A plasmonic cavity structure between metallic nanoantennae. The electric field in the cavity is strongly enhanced in the “hot spot” generated from the nanoantennae. 
Micodisk device An optical QED architecture where emitters interact with the modes of a bent waveguide. 
Micropillar device An optical QED architecture where emitters are sandwiched between Bragg reflectors. 
Mode volume, Vm The volume of the quantized electromagnetic field generated within a cavity. 
Nanoantenna A plasmonic nanostructure that transfers electromagnetic energy from near-field to far-field due to its localized surface plasmon resonances. 
Photonic crystal cavity device An optical QED architecture where emitters interact with the modes of periodic dielectric structures. 
Plasmonic nanoresonator A metallic optical device with a periodic structure. 
Purcell effect The enhancement of the rate of spontaneous emission upon the coupling of an emitter to a cavity or a plasmon. 
Rabi splitting The observable splitting of the emission spectrum of an emitter resulting from the nondegenerate hybridized states accessible in the strong coupling regime. 
γC The full width at half maximum (FWHM) of the cavity mode. 
γX The full width at half maximum (FWHM) of the exciton mode. 

The coupling of a QD (or any emitter) to a cavity can be described by a coupled-oscillator model.17 Consider the on-resonance (or zero detuning) case, where the energy of the cavity mode is equal to the energy of the exciton mode of the QD (EC = EX = E0). At resonance, the energy of the coupled modes, E1,2, is defined by the following equation:

E1,2=E0iγC+γX4±g2γCγX21612,
(1)

where γC is the full-width-at-half-maximum (FWHM) of the cavity mode, γX is the FWHM of the exciton mode, and g is the exciton-photon coupling parameter (Fig. 2).17,g is the scalar product of the transition matrix element of the exciton’s dipole moment and the value of the electric field at the QD’s position, either within a microcavity or with respect to a proximate plasmonic structure. The threshold for the strong coupling of a QD and a cavity is defined as g>γCγX4.17 

FIG. 2.

(a) An illustration of a two-level quantum emitter within a microcavity. The γ is the decay rate of the quantum emitter, and κ is the loss rate of the cavity photon. The emitter is coupled to the cavity with a coupling strength, g. (b) The vacuum Rabi splitting observed when the emitter is in resonance with the cavity. The Rabi splitting has a magnitude of 2g and a width of κ+γ2. Reprinted with permission from Khitrova et al., Nat. Phys. 2(2), 81–90 (2006). Copyright 2006 Nature Publishing Group.

FIG. 2.

(a) An illustration of a two-level quantum emitter within a microcavity. The γ is the decay rate of the quantum emitter, and κ is the loss rate of the cavity photon. The emitter is coupled to the cavity with a coupling strength, g. (b) The vacuum Rabi splitting observed when the emitter is in resonance with the cavity. The Rabi splitting has a magnitude of 2g and a width of κ+γ2. Reprinted with permission from Khitrova et al., Nat. Phys. 2(2), 81–90 (2006). Copyright 2006 Nature Publishing Group.

Close modal

A semiclassical model of a plasmonic metal nanoparticle interacting with an excitonic material suggests that the only type of coupled system in which the coupling constant is independent of the magnitude of the electric field at the position of the excitonic material is one in which the plasmonic metal nanoparticle is homogenously and infinitely coated with the excitonic material.18 Since a single QD cannot homogenously and infinitely encompass the electromagnetic field of a cavity, the QD should be placed in the antinode of the field in all three dimensions to maximize the available field strength. With this geometry, the coupling parameter is optimized by maximizing the product fVm12Q, where f is the QD oscillator strength, Vm is the mode volume, and Q is the quality of the cavity. Q is defined as EC/γC, the ratio of the energy of the cavity to the FWHM of the cavity mode.17 While higher Q values are desirable for increasing the coupling strength, the mode volume of the cavity must also be considered. Ultrahigh quality (Q > 108) cavities have been fabricated,19 but their microscale geometry results in larger cavity mode volumes, which limit the maximum possible coupling strength. In fact, Q factors of just 102–103 may result in higher coupling strengths if the cavity is of nanoscale.20,21 The strong coupling threshold therefore may be achieved by increasing the oscillator strength of the relevant QD excitonic transitions (by, for instance, using semiconductor nanoplatelets or nanorods), decreasing the mode volume for cavities of high quality, or increasing the quality factor, Q, for cavities with small mode volumes.

We proceed below by defining the various exciton-photon coupling regimes (summarized in Table II). We specify both the weak and intermediate regimes, which fall within the mathematically defined weak-coupling limit (gγCγX4), as well as the strong and ultrastrong regimes, which are within the mathematically defined strong-coupling limit (g>γCγX4). It is possible to observe multiple coupling regimes within the same system (Fig. 3).22,23

TABLE II.

Observable changes in the spectroscopic properties of QED systems in each coupling regime.

gγCγX4g>γCγX4
Weak couplingIntermediate couplingStrong coupling
Absorbance /Scattering Some broadening and shifting of peak Dip in peak and asymmetry (Fano-like interference) Two resolved peaks (Rabi splitting) 
Emission Modified decay rate (Purcell effect) Modified decay rate (Purcell effect) Two resolved peaks (Rabi splitting) 
gγCγX4g>γCγX4
Weak couplingIntermediate couplingStrong coupling
Absorbance /Scattering Some broadening and shifting of peak Dip in peak and asymmetry (Fano-like interference) Two resolved peaks (Rabi splitting) 
Emission Modified decay rate (Purcell effect) Modified decay rate (Purcell effect) Two resolved peaks (Rabi splitting) 
FIG. 3.

The predicted scattering (blue) and emission (green) spectra for a coupled QD-cavity system as a function of coupling strength, g, where weak coupling (a) produces a single peak in both the scattering and emission spectra, (b) intermediate coupling produces a doublet in the scattering spectrum and an asymmetric peak in the emission spectrum, and, (c) strong coupling yields a doublet with resolvable peaks in both the scattering and emission spectra. Reprinted with permission from Leng et al., Nat. Commun. 9(1), 4012 (2018). Copyright 2018 Nature Publishing Group.

FIG. 3.

The predicted scattering (blue) and emission (green) spectra for a coupled QD-cavity system as a function of coupling strength, g, where weak coupling (a) produces a single peak in both the scattering and emission spectra, (b) intermediate coupling produces a doublet in the scattering spectrum and an asymmetric peak in the emission spectrum, and, (c) strong coupling yields a doublet with resolvable peaks in both the scattering and emission spectra. Reprinted with permission from Leng et al., Nat. Commun. 9(1), 4012 (2018). Copyright 2018 Nature Publishing Group.

Close modal

In the weak coupling regime, the energies of the coupled modes, E1,2, are degenerate (i.e., zero splitting)17 and the QD and cavity spectral components generally resemble their isolated states.24,25 Weak coupling between a QD and a cavity results in a peak shift and line broadening of the absorption spectrum of the QD, although both the multipeak absorption spectrum and the single-peak emission spectrum retain the shape of those of an isolated QD. The rate of irreversible spontaneous emission from the QD is, however, enhanced (when on-resonance with the cavity mode) or suppressed (upon detuning) due to the increase of the density of states of radiation that occurs upon confinement within a cavity, as dictated by Fermi’s golden rule.24,26 The magnitude of the enhancement resulting from this phenomenon, known as the Purcell effect,27,28 can be calculated using cavity properties Q and Vm, and is quantified by the Purcell factor, FP,

FP=34π2λ3QVm,
(2)

where λ is the wavelength of the transition within the cavity. Equation (2) assumes that (i) the QD is located in the antinode of the cavity’s electromagnetic field in all three directions, (ii) the linewidth of the QD is narrower than the linewidth of the cavity mode, (iii) the dipole of the QD is aligned with the polarization of the cavity mode, and (iv) the cavity mode is in resonance with the frequency of the excitonic transition (i.e., zero detuning).29 Although it may be expected that the linewidth of a single QD would be larger than that of a cavity mode at 300 K, an epitaxially grown InAs QD located within a microcavitiy has been shown to have a smaller emitter linewidth compared to the cavity mode.30 Such a condition has also been fulfilled with epitaxially grown CdSe/ZnSe QDs located between SiO2/TiO2 Bragg reflectors.31 

As the coupling strength increases, but remains below the strong-coupling limit, an asymmetric shape in the absorption spectrum (or scattering spectrum) of the emitter develops due to quantum interference between the discrete exciton mode and the continuum of states to which it couples.22,23,32,33 The emission spectrum, however, retains a single peak. The appearance of the asymmetric shape in the optical spectrum of the emitter is known as the Fano effect.33 

In the strong coupling regime, the energies of the coupled modes (E1,2) formed upon hybridization result in two resolvable peaks with different energies but identical linewidths. The presence of clearly resolvable doublets in the spontaneous emission of a QED system is known as vacuum Rabi splitting,34 and the magnitude of the splitting increases with the coupling strength.35 In the strong coupling regime, coherent energy exchange is faster than energy dissipation due to decay and/or decoherence;8,26,28 therefore, coherent oscillations and quantum effects resulting from coupling are observed.25 The nondegeneracy of the eigenstates gives rise to avoided level crossings, a key indication of a strongly interacting quantum system.35 

The so-called “ultrastrong coupling” occurs when the coupling strength, g, is greater than the dipole transition energy of the emitter. In the ultrastrong coupling regime, unlike weaker coupling regimes, the ground state of the coupled system is modified, and previously neglected terms of the Hamiltonian, such as contributions from the carriers of the cavity mode and counter-rotating terms typically neglected under the rotating wave approximation, become significant.25,26,36,37 Accessing this regime may result in asymmetry in the vacuum Rabi splitting, photon generation, and new transition pathways.37 

Strong coupling has been achieved in a variety of QED structures where either colloidal or epitaxial QDs are the emitters (Table III). Initial QED systems used epitaxially grown QDs and optical microcavities; these cavities are diffraction-limited, where the mode volume, Vm, is dictated by the wavelength of incident light.38 More recently, QED devices were fabricated with plasmonic cavities, which overcome the diffraction limit.

TABLE III.

Experimental demonstrations of coupling of a QD to a cavity.

Type of QDEpitaxial or colloidal QDComposition of cavityYearReferences
GaAs Epitaxial Fabry-Pérot cavity with GaAlAs Bragg reflectors of GaAlAs/AlAs layers 1992 41  
In0.13Ga0.87As Epitaxial AlxGa1−xAs/InyGa1−xAs micropillar 1994 39 and 40  
InAs Epitaxial Epitaxially grown GaAs microcavity with AlAs/GaAs Bragg reflectors 2001 30  
InAs Epitaxial Photonic crystal slab on GaAs substrate 2004 44  
In0.3Ga0.7As Epitaxial GaAs micropillar 2004 17  
CdSe/ZnSe Epitaxial ZnSe cavity between two SiO2/TiO2 Bragg reflectors 2005 31  
GaAs Epitaxial Al0.33Ga0.67As microdisk 2005 42  
CdSe/ZnS nanorods Colloidal Polystyrene sphere 2006 47  
InAs Epitaxial GaAs photonic crystal slab 2007 72  
InGaAs Epitaxial Bragg reflectors on GaAs substrate 2007 69  
InGaAs Epitaxial GaAs photonic crystal 2010 90  
CdSe Colloidal Thermally evaporated Ag film 2010–2012 48 and 49  
CdSe/ZnSe Colloidal Slab waveguide on a glass substrate 2011 74  
CdSe/ZnS Colloidal Ag wire 2013 50  
InAs Epitaxial GaAs micropillar cavity with GaAs and AlAs layered Bragg reflectors 2013 16  
InAs Epitaxial GaAs photonic crystal 2013 16  
CdSe/ZnS Colloidal Ag bowtie plasmonic cavities 2016 52  
CdSe Colloidal Au nanohole array 2016 55  
CdSe/ZnS Colloidal Ag nanoparticles on Ag nanoshells 2016 91  
CdSeTe/ZnS Colloidal Slit-like plasmonic nanoresonator fabricated at corner of Ag flake 2018 56  
CdSe/CdS Colloidal Nanogap between Au nanoparticle and Ag film 2018 22  
CdSe/ZnS Colloidal Au TERS tip and Au template-stripped surface 2019 92  
Type of QDEpitaxial or colloidal QDComposition of cavityYearReferences
GaAs Epitaxial Fabry-Pérot cavity with GaAlAs Bragg reflectors of GaAlAs/AlAs layers 1992 41  
In0.13Ga0.87As Epitaxial AlxGa1−xAs/InyGa1−xAs micropillar 1994 39 and 40  
InAs Epitaxial Epitaxially grown GaAs microcavity with AlAs/GaAs Bragg reflectors 2001 30  
InAs Epitaxial Photonic crystal slab on GaAs substrate 2004 44  
In0.3Ga0.7As Epitaxial GaAs micropillar 2004 17  
CdSe/ZnSe Epitaxial ZnSe cavity between two SiO2/TiO2 Bragg reflectors 2005 31  
GaAs Epitaxial Al0.33Ga0.67As microdisk 2005 42  
CdSe/ZnS nanorods Colloidal Polystyrene sphere 2006 47  
InAs Epitaxial GaAs photonic crystal slab 2007 72  
InGaAs Epitaxial Bragg reflectors on GaAs substrate 2007 69  
InGaAs Epitaxial GaAs photonic crystal 2010 90  
CdSe Colloidal Thermally evaporated Ag film 2010–2012 48 and 49  
CdSe/ZnSe Colloidal Slab waveguide on a glass substrate 2011 74  
CdSe/ZnS Colloidal Ag wire 2013 50  
InAs Epitaxial GaAs micropillar cavity with GaAs and AlAs layered Bragg reflectors 2013 16  
InAs Epitaxial GaAs photonic crystal 2013 16  
CdSe/ZnS Colloidal Ag bowtie plasmonic cavities 2016 52  
CdSe Colloidal Au nanohole array 2016 55  
CdSe/ZnS Colloidal Ag nanoparticles on Ag nanoshells 2016 91  
CdSeTe/ZnS Colloidal Slit-like plasmonic nanoresonator fabricated at corner of Ag flake 2018 56  
CdSe/CdS Colloidal Nanogap between Au nanoparticle and Ag film 2018 22  
CdSe/ZnS Colloidal Au TERS tip and Au template-stripped surface 2019 92  

In early examples of strong coupling between quantum-confined semiconductors and optical microcavities, the cavity devices were fabricated using either molecular beam epitaxy39,40 or chemical vapor deposition.41 The emitters in these devices were typically InxGa1−xAs or GaAs quantum wells (QWs). The three most common geometries of these devices include micropillar devices, microdisk devices, and photonic crystal cavities, as shown in Fig. 4.

FIG. 4.

(a) Schematic of a micropillar QED structure with InAs QDs in a GaAs cavity that is sandwiched between mirrors or distributed Bragg reflectors. Reprinted with permission from Muller et al., Phys. Rev. Lett. 99(18), 187402 (2007). Copyright 2007 American Physical Society. (b) Scanning electron micrograph of a microdisk QED system. Reprinted with permission from Peter et al., Phys. Rev. Lett. 95(6), 067401 (2005). Copyright 2005 American Physical Society. (c) Schematic of a photonic crystal QED structure with InAs QDs. Reprinted with permission from Yoshie et al., Nature 432(7014), 200–203 (2004). Copyright 2014 Nature Publishing Group.

FIG. 4.

(a) Schematic of a micropillar QED structure with InAs QDs in a GaAs cavity that is sandwiched between mirrors or distributed Bragg reflectors. Reprinted with permission from Muller et al., Phys. Rev. Lett. 99(18), 187402 (2007). Copyright 2007 American Physical Society. (b) Scanning electron micrograph of a microdisk QED system. Reprinted with permission from Peter et al., Phys. Rev. Lett. 95(6), 067401 (2005). Copyright 2005 American Physical Society. (c) Schematic of a photonic crystal QED structure with InAs QDs. Reprinted with permission from Yoshie et al., Nature 432(7014), 200–203 (2004). Copyright 2014 Nature Publishing Group.

Close modal

Micropillar devices [Fig. 4(a)] consist of one or multiple layers of QWs sandwiched between Bragg reflectors, alternating layers of AlAs and AlGaAs.39–41 The Bragg reflectors act as mirrors and optically confine light in the axial direction such that light from micropillar devices is emitted in this direction. Q values for the cavity modes of micropillar devices can reach 5000–10 000.34 In microdisk devices [Fig. 4(b)], QWs are grown between slabs of a dielectric material.42 Electron beam lithography and chemical etching are then utilized to create the microdisk structures.42 Light is confined in all three dimensions by total internal reflection, which in some cases results in cavity modes with Q-factors greater than 10 000.34 In photonic crystal cavity devices [Fig. 4(c)], QWs are grown epitaxially in dielectric slabs, and lithography and etching are then utilized to cut holes into the dielectric slabs; these holes serve as the cavities.43 The geometry of the holes in the photonic crystal layer has a large impact on the Q value of the cavity mode; GaAs-based photonic crystal cavities are known to achieve Q factors of 20 000.44 

With the development of increasingly precise fabrication techniques, use of epitaxial QDs as emitters became possible. QDs offer two primary advantages over QWs: narrower linewidths under weak excitation conditions29 and an enhanced rate of spontaneous emission into the cavity mode.30,45

The main limitation of these epitaxially grown, optical microcavity devices is the Vm. The mode volumes are large because they are dictated by the wavelength of the incident light, which limits the Q-factor of the cavity and results in Rabi splitting energies of <1 meV.46 As a result of the large Vm of the cavities, the emitters in the cavities must have large dipole moments in order for the system to achieve strong coupling;34 this condition limits the types of quantum-confined semiconductors that can be used as emitters. One can straightforwardly increase the oscillator strength of semiconductor emitters by moving from spherical QDs to anisotropic nanocrystals such as nanorods, nanoplatelets, and nanosheets. This strategy was successfully employed in the development of a QED architecture that exhibited strong-coupling between the exciton mode of a CdSe/ZnS nanorod and a photon mode of a dielectric polystyrene microsphere.47 

Cavities composed of plasmonic materials can have cavity-mode volumes that are smaller than the diffraction limit because surface plasmon polaritons (SPPs) can localize light fields to nanoscale dimensions.24,46 In experimental realizations of plasmonic QED systems, the QD exciton interacts with (i) single isolated surface plasmons, (ii) multiple coupled plasmons, or (iii) the collective plasmon of the cavity.

The earliest experimental examples of plasmonic QED systems demonstrated coupling between QDs and the surface plasmons of planar Ag films.18,48,49 These devices were prepared by spin coating colloidal suspensions of CdSe QDs onto thermally evaporated Ag films on glass substrates. In one example, a QED device was fabricated by lithographically placing CdSeTe/ZnS QDs along a plasmonic silver wire [Fig. 5(a)].50 This device construction enables the QD excitons to interact with the localized surface plasmon resonances of the Ag nanowire.50 One strategy to increase the coupling strength in a QED system is to decease the size of the nanoantenna; e.g., in the case of spheroid nanoantenna, the size of the minor axis affects the coupling strength between the nanoantenna and QDs.51 Small nanoantennae, however, suffer from high absorption losses.51 

FIG. 5.

(a) A QED structure with QDs interacting with the surface plasmon polaritons of a Ag nanowire. Reprinted with permission from Gruber et al., Nano Lett. 13(9), 4257–4262 (2013). Copyright 2013 American Chemical Society. (b) Schematic of a QD in a metal nanogap cavity composed of Ag bowties. Reprinted with permission from Santhosh et al., Nat. Commun. 7, 11823 (2016). Copyright 2016 Nature Publishing Group. (c) A QED structure of CdSe QDs interacting with a plasmonic Au nanohole array. Reprinted with permission from Wang et al., J. Phys. Chem. Lett. 7(22), 4648–4654 (2016). Copyright 2016 American Chemical Society.

FIG. 5.

(a) A QED structure with QDs interacting with the surface plasmon polaritons of a Ag nanowire. Reprinted with permission from Gruber et al., Nano Lett. 13(9), 4257–4262 (2013). Copyright 2013 American Chemical Society. (b) Schematic of a QD in a metal nanogap cavity composed of Ag bowties. Reprinted with permission from Santhosh et al., Nat. Commun. 7, 11823 (2016). Copyright 2016 Nature Publishing Group. (c) A QED structure of CdSe QDs interacting with a plasmonic Au nanohole array. Reprinted with permission from Wang et al., J. Phys. Chem. Lett. 7(22), 4648–4654 (2016). Copyright 2016 American Chemical Society.

Close modal

One promising QED architecture for modulating the electric field generated by localized surface plasmons is the metal nanogap structure, where the electric field of the cavity is produced by coupled plasmons of proximate nanostructures. Such systems have been studied extensively since the theoretical prediction that strong coupling is achievable between a single quantum emitter, e.g., a QD, that is placed in a nanoscale gap between two metal nanoparticles and the electric field in that gap.38 A variety of methods have been utilized to create metal nanogap structures. For example, interfacial capillary forces were exploited to install CdSe/ZnS QDs in the gap between two Ag nanoparticles arranged in a “bowtie” shape [Fig. 5(b)].52 The resonance of the cavity was tuned by varying the side-to-side length of the Ag bowties and the distance between the bowties.52 In the same year, strong coupling was reported between colloidal suspensions of mixtures of Ag nanoshells on SiO2 cores and CdSe/ZnS QDs.53 Because the Ag nanoshells were synthesized using a seed-mediated method, their surfaces were notably roughened with Ag nanoparticles. The electric field of the nanoshell under laser excitation was enhanced, especially between neighboring nanoparticles, and, as such, strong coupling was achievable with QDs.53 

More recently, surface chemistry of colloidal QDs was used to covalently link Au nanoparticles to CdSe/CdS QDs through the ligands on both the Au particles and the QDs.22 These linked structures were then drop-cast onto Ag films, and an Au nanoparticle and the Ag film acted as the cavity.22 Weak, intermediate, and strong coupling were observed between plasmonic cavities and QDs in the system due to the nonuniformity of the local structure.22 Another method, explored theoretically, for achieving such devices used Ag nanotweezers to optically trap QDs in the cavity.54 A major limitation of the metal nanogap structure strategy for achieving strong coupling between a QD and a photon is that, as mentioned above, the QD emitter must be localized to the antinode of the electric field of the cavity.

There are several examples of QED devices where the QD interacts with the collective plasmon mode of the cavity. Such a QED device has been fabricated with a drop-cast layer of colloidal CdSe QDs and an Au nanohole array [Fig. 5(c)].55 The coupling strength between the QDs and the Au nanoarray was altered by changing the spacing of the array.55 Recently, a slit-like nanoresonator on the corner of an Ag flake served as a probe for nanometer-precise raster scanning of a film of isolated CdSeTe/ZnS QDs embedded in a poly(methylmethacrylate) film.56 The coupling strength of the QED system was easily modifiable by changing the distance between the QD and the nanoresonator.56 Although challenging to fabricate, plasmonic nanoresonators are attractive for cavities because their resonance wavelength is easily tunable. The electric field produced by the resonators, however, is weaker than the one produced in the “hot spot” of the nanogap structures.

As discussed in the Definitions of Coupling Regimes, the coupling of QDs to photonic cavities or surface plasmon polaritons (SPPs) can greatly affect the absorption and emission properties of the system. Even weakly coupled systems can increase the rate of spontaneous emission by the Purcell effect, as observed, for example, through time-resolved fluorescence spectroscopy [Fig. 6(a)].57 In addition, coupling a QD to a cavity can dramatically decrease the probability of multiphoton emission. The deterministic generation of single photons by QD-microcavity systems is a first step toward realization of photonic qubits, where the quantum properties of light such as polarization, momentum, and energy can be used for the storage and processing of quantum information.58 

FIG. 6.

(a) Change in emission lifetime for quantum dots (QDs) on a glass substrate (blue) and in a photonic nanocavity (red). Reprinted with permission from Hoang et al., Nano Lett. 16(1), 270–275 (2016). Copyright 2016 American Chemical Society. (b) A single QD placed in a Yagi-Uda antenna and the theoretical (red) and measured (black) radiation pattern of its fluorescence. Reprinted with permission from Curto et al., Science 329(5994), 930 (2010). Copyright 2010 AAAS.

FIG. 6.

(a) Change in emission lifetime for quantum dots (QDs) on a glass substrate (blue) and in a photonic nanocavity (red). Reprinted with permission from Hoang et al., Nano Lett. 16(1), 270–275 (2016). Copyright 2016 American Chemical Society. (b) A single QD placed in a Yagi-Uda antenna and the theoretical (red) and measured (black) radiation pattern of its fluorescence. Reprinted with permission from Curto et al., Science 329(5994), 930 (2010). Copyright 2010 AAAS.

Close modal

The directionality and polarization of the QD emission can also be controlled by an anisotropic electric field that encourages spontaneous emission in a preferred direction.24 This anisotropy is obtained with plasmonic materials by altering the shape of the metal nanostructure and the position of the QD.24,50,59–61 For example, Fig. 6(b) shows that placing a single QD emitter in a Yagi-Uda antenna array results in unidirectional and strongly polarized emission.62 Similarly, in 2009, it was reported that the rate of directional emission for CdS-ZnS core/shell QDs was enhanced when coupled to an Au plasmonic nanodisk system.59 Using time-resolved emission, the authors observed that the radiative decay of the system is sensitive to the polarization of the incident light and the collection angle, suggesting that the enhancement of the directional emission requires both spectral and spatial overlap. This directionality can potentially be applied in lighting, optical sensing, quantum emitters, and microscopy.

While QD-cavity systems in the weak coupling regime show perturbations of the QD emission, systems in the strong coupling regime demonstrate properties unique to either of the separate components. As discussed in the Definitions of Coupling Regimes, strong coupling occurs when the emitter is able to exchange energy with the cavity or SPP, resulting in the formation of hybrid exciton/electric field states.34 The observation of these new upper polariton (UP) and lower polariton (LP) states and the Rabi splitting between them is the definitive indication that the system is strongly coupled.63 Optical excitation from the QD ground state to the UP and LP states has been predicted theoretically64 and experimentally measured through steady-state and time-resolved absorbance spectroscopy.40,55 Similarly, Rabi splitting has been observed in scattering spectra.18,52 For example, in 2010, room-temperature coupling between CdSe QDs and a Ag plasmonic film was reported with a Rabi splitting of ∼112 meV.48 The authors proposed that, in the strong coupling regime, QD-based materials have the potential to be used in next-generation all-optical nonlinear devices, threshold-less laser operation, and single photon transistors. One consideration in identifying the strong coupling regime, however, is that similar splitting in the scattering spectra can be caused by Fano-like interference in the intermediate coupling regime.22,64 The distinction between Fano-splitting and Rabi splitting can be resolved by examining the photoluminescence of the QD, as shown in Figs. 3(a)–3(c). Photoluminescence, unlike scattering, is an incoherent process and thus is unaffected by Fano interferences. For these reasons, fluorescence spectroscopy is recognized as the preferred characterization technique for identifying and quantifying Rabi splitting.22,34

For more than a decade, it was predicted theoretically that strong coupling of a QD and a metal nanoparticle could result in splitting of the emission spectrum of the QD.65–67 More recently, such splitting has also been measured experimentally for QDs coupled with SPPs22,49,52,53,56,68 and microcavities.17,44,69 Particularly with SPPs, much work has been dedicated to understanding how the geometry of the metal nanoparticle can be optimized to maximize coupling while keeping the size of the system small. In 2010, it was predicted that Rabi splitting should be observed for a single QD emitter such as CdS or ZnS in the center of a Ag dimer nanoantenna, a system with dimensions smaller than 36 nm.38 As mentioned in the Experimental architectures section, “bowtie” structures like the example shown in Fig. 7(a) are known to create a “hotspot” region between the two prisms with an electric field that is enhanced compared to that created by the single nanostructure. In one study, placing CdSe/ZnS QDs in this hotspot resulted in a 140-meV Rabi splitting.52,68 The utilization of SPPs to produce focused electromagnetic fields provides an effective approach for quantum logic operations on a scale compatible with compact electronic devices.

FIG. 7.

(a) Distribution of the coupling rate for two nanoprisms in a bowtie configuration. The scale bar indicates 10 nm. Reprinted with permission from Santhosh et al., Nat. Commun. 7, 11823 (2016). Copyright 2016 Nature Publishing Group. (b) Reflectivity spectra obtained at varying angles of incidence for QDs on a silver film, and (c) the corresponding dispersion curve with fits for the upper (UP) and lower (LP) polaritons as well as the isolated QD (blue) and SPP (black) features. Reprinted with permission from Gómez et al., J. Phys. Chem. B 117(16), 4340–4346 (2012). Copyright 2012 American Chemical Society.

FIG. 7.

(a) Distribution of the coupling rate for two nanoprisms in a bowtie configuration. The scale bar indicates 10 nm. Reprinted with permission from Santhosh et al., Nat. Commun. 7, 11823 (2016). Copyright 2016 Nature Publishing Group. (b) Reflectivity spectra obtained at varying angles of incidence for QDs on a silver film, and (c) the corresponding dispersion curve with fits for the upper (UP) and lower (LP) polaritons as well as the isolated QD (blue) and SPP (black) features. Reprinted with permission from Gómez et al., J. Phys. Chem. B 117(16), 4340–4346 (2012). Copyright 2012 American Chemical Society.

Close modal

The transition between a polariton state and a lasing state has been observed in a self-assembled InAs QD-photonic crystal microcavity (Fig. 8).70 The emission spectrum of the system changes from double-peaked to single-peaked upon increasing the power of the light source, suggesting that lasing is possible even in the strongly coupled regime. Similarly, in 2014, it was determined computationally that the dark plasmon in a QD-SPP system can transfer its energy to the QD to achieve emission without population inversion. This gain-without-inversion effect can be used for QD lasing.71 

FIG. 8.

(a) Schematic of self-assembled InAs QDs in a photonic crystal nanocavity. (b) Emission spectrum of the system at different excitation power. The system changes from a polariton state in the strong-coupled regime to a lasing state upon increase in the excitation power. Reprinted with permission from Nomura et al., Nat. Phys. 6(4), 279 (2010). Copyright 2010 Nature Publishing Group.

FIG. 8.

(a) Schematic of self-assembled InAs QDs in a photonic crystal nanocavity. (b) Emission spectrum of the system at different excitation power. The system changes from a polariton state in the strong-coupled regime to a lasing state upon increase in the excitation power. Reprinted with permission from Nomura et al., Nat. Phys. 6(4), 279 (2010). Copyright 2010 Nature Publishing Group.

Close modal

A further method of identifying strong coupling is to observe the mixing of QD electronic excited states and photonic states to form the new UP and LP modes as a function of coupling strength. The strength of coupling can be changed by manipulating the physical location of the QD relative to the plasmonic material41,56,72 or by “detuning:” bringing the energies of the excitonic and optical modes in and out of resonance.24 Early experiments with photonic cavities demonstrated detuning by changing the temperature of the system.16,17,42,44

For planar systems, the angle of incidence can similarly be used to vary the energy of the plasmonic state and, therefore, the degree of detuning.39,48,49,73 As shown in Figs. 7(b) and 7(c), a dispersion curve can be generated by mapping the UP and LP states as a function of the angle of incidence. The characteristics of the UP and LP states switch from primarily excitonic to primarily plasmonic depending on the angle of incidence. For example, the SPP state is expected to have a shorter lifetime than the exciton state.74 Time-resolved reflectivity measurements revealed a decrease in the decay rate of the LP mode between 45° and 65°, which indicates that the relative composition of the hybrid state changes from primarily plasmonic to primarily excitonic.49 

QDs are good emitters for QED systems, as they have narrow photoluminescence linewidths (20–30 nm at 298 K) and are often more photostable than molecular emitters. QDs also offer an advantage over molecules when it comes to biexciton extraction and the generation of entangled photon pairs. A recent study showed that CdSe/ZnSe QDs in the strong coupling regime with an optical microcavity have exciton-photon couplings on the order of 20–30 meV, which is on the same order of the biexciton binding energy in these materials and therefore suggests that biexciton extraction and generation of entangled photon pairs is possible for colloidal QD-based structures.74 

QD-based systems still suffer from several limiting factors. First, QDs are not “artificial atoms” as they are often modeled. The quasicontinuum of transitions of the QD can overlap with the cavity resonance, increasing the difficulty of accurately predicting their light-matter interactions.75,76 QDs are, in general, photostable, but for a strongly coupled light-matter system, continuous incoherent pumping averages out Rabi oscillations and broadens emission lines.77,78 In addition to homogeneous broadening, there is also dispersity in QD sizes and, therefore, bandgaps.78 Even high-quality synthetic batches of QDs have particles with radii that vary by 5%–10% within the ensemble; such heterogeneity complicates the determination of Rabi splitting.73 Realizing strong coupling with single QD-cavity systems alleviates this issue, but it is also challenging to measure Rabi splitting, the magnitude of which is proportional to N, where N is the number of emitters, with isolated individual QDs .52,56,79 Multiemitter ensembles additionally have the advantage that they can give rise to the so-called “cavity protection,” where the polariton is decoupled from the nonemissive states that result in its relaxation.78 

Cavity-coupled photochemistry, in which coupling to a cavity modifies the potential energy surface of the reaction, has been demonstrated with organic dye molecules in a number of theoretical and experimental studies. For example, in 2016, theoretical prediction showed that the rate of electron transfer from a molecular excited state coupled to the cavity mode to a molecular excited state decoupled from the cavity mode can be significantly enhanced within an optical cavity. This increase was attributed to (i) a reduced energy gap between donor and acceptor orbitals due to Rabi splitting and (ii) the suppression of the rearrangement of nuclei (reduced reorganization energy) during charge transfer due to strong cavity-molecule coupling.80 In 2012, Rabi splitting was observed in the absorbance of spyropyran in a silver mirror cavity. This formation of hybridized light-matter states leads to a decrease in the rate of photoinduced ring-opening of spyropyran.7 A study from 2016 reported a decrease in the rate of the deprotection of 1-phenyl-2-trimethylsilylacetylene when the Si-C vibrational mode couples with the vacuum electromagnetic field.81 In 2017, a system was reported that used J-aggregate cyanine dyes as energy donors and acceptors spatially separated by poly(methylmethacrylate) in an optical cavity of two parallel silver mirrors. The energy transfer between dyes in the strongly coupled regime occurred with an efficiency of 37% when the distance between the donor and the acceptor was larger than 100 nm, which is beyond the characteristic Förster radius of the system, due to formation of hybrid polaritonic states.82 Thus, we see that cavity-coupled organic dye molecules may undergo photochemical processes that are highly sensitive to this coupling.

We propose that a next step in cavity-coupled photochemistry is to use QDs as the photosensitizer or photocatalyst for a substrate molecule and thereby investigate the emergent photochemical properties of cavity-coupled QD systems. In order to achieve such a system, we may take advantage of extensive previous work with (i) QD-coupled QED systems in both the weak and strong coupling regime, (ii) QED photochemistry achieved with molecular emitters, and (iii) studies of QDs as photocatalysts for a variety of excited state and redox reactions.83 

The confluence of these three research avenues provides a foundation for the future exploration of QD-based QED photochemistry. For example, the ability of QD photocatalysts to perform electron or energy transfer to molecular acceptors may be significantly enhanced when coupling with a plasmonic material or an optical cavity. Such architectures may also enable chiral photocatalysis, without the use of bulky substituents, through the incorporation of a chiral emitter that also serves as the photocatalyst. QDs can be transformed into chiral emitters through functionalization with chiral ligands84 or coupling with nanophotonic waveguides.85,86

The photochemical QED systems that have been realized use dye molecules as both the emitter and the reagent, such that reaction pathways access newly formed hybrid polaritonic states of the emitter-cavity system. With QD-based QED systems, the QD emitter is generally not the reagent in a photochemical reaction but rather the photosensitizer or the photocatalyst. The fact that the QDs in such a system would be intermediate chromophores may pose a challenge to the modification of the potential surface of a cavity-coupled photochemical reaction. We propose that a promising direction to explore to answer this challenge is to link the QDs to catalytic substrate molecules through so-called “exciton-delocalizing ligands” such as phenyldithiocarbamate and its derivatives.87–89 When capped with exciton-delocalizing ligands, the wavefunction of the QD extends into the ligand capping layer through the generation of interfacial orbitals at the QD surface, resulting in a relaxation of exciton confinement and strong coupling with the substrate molecule. The creation of such interfacial states should result in the entire QD-delocalizing linker-substrate system acting as essentially one species that is coupled to the cavity. We therefore suspect that the QD and its ligand shell can serve as an intermediary between the substrate and the cavity and provide enough coupling to change the potential energy surface for the photocatalytic reaction.89 

Experimental systems demonstrating strong-coupling between a single QD or multiple QDs and a quantized electric field are few in number compared to the theoretical studies on such systems. The few experimental QD-based polaritonic systems in the literature suggest that more extensive research is warranted to fully understand how to optimize QED systems with QDs for applications such as quantum information, lasing, and new photochemical reactivity.

The authors thank the National Science Foundation (Grant No. CHE-1664184) and the Air Force Office for Scientific Research (Grant No. FA9550-17-1-0271) for support of this work. K.P.M. acknowledges a graduate fellowship through the National Science Foundation (Award No. DGE-1324585). K.A.P. was supported by the Department of Defense (DoD) through the National Defense Science and Engineering Graduate Fellowship (NDSEG) Program. J.C.S. acknowledges a graduate fellowship through the National Science Foundation (Award No. DGE-1324585).

The authors declare no competing financial interests.

1.
M. A.
Ischay
,
Z.
Lu
, and
T. P.
Yoon
, “
[2+2] cycloadditions by oxidative visible light photocatalysis
,”
J. Am. Chem. Soc.
132
(
25
),
8572
8574
(
2010
).
2.
C. K.
Prier
,
D. A.
Rankic
, and
D. W. C.
MacMillan
, “
Visible light photoredox catalysis with transition metal complexes: Applications in organic synthesis
,”
Chem. Rev.
113
(
7
),
5322
5363
(
2013
).
3.
S.
Lian
,
M. S.
Kodaimati
, and
E. A.
Weiss
, “
Photocatalytically active superstructures of quantum dots and iron porphyrins for reduction of CO2 to CO in water
,”
ACS Nano
12
(
1
),
568
575
(
2018
).
4.
M. S.
Kodaimati
,
S.
Lian
,
G. C.
Schatz
, and
E. A.
Weiss
, “
Energy transfer-enhanced photocatalytic reduction of protons within quantum dot light-harvesting–catalyst assemblies
,”
Proc. Natl. Acad. Sci. U. S. A.
115
(
33
),
8290
8295
(
2018
).
5.
W.
Cai
,
H.
Fan
,
D.
Ding
,
Y.
Zhang
, and
W.
Wang
, “
Synthesis of Z-alkenes via visible light promoted photocatalytic E → Z isomerization under metal-free conditions
,”
Chem. Commun.
53
(
96
),
12918
12921
(
2017
).
6.
T.
Szidarovszky
,
G. J.
Halász
,
A. G.
Császár
,
L. S.
Cederbaum
, and
Á.
Vibók
, “
Conical intersections induced by quantum light: Field-dressed spectra from the weak to the ultrastrong coupling regimes
,”
J. Phys. Chem. Lett.
9
(
21
),
6215
6223
(
2018
).
7.
J. A.
Hutchison
,
T.
Schwartz
,
C.
Genet
,
E.
Devaux
, and
T. W.
Ebbesen
, “
Modifying chemical landscapes by coupling to vacuum fields
,”
Angew. Chem., Int. Ed.
51
(
7
),
1592
1596
(
2012
).
8.
J.
Feist
,
J.
Galego
, and
F. J.
Garcia-Vidal
, “
Polaritonic chemistry with organic molecules
,”
ACS Photonics
5
(
1
),
205
216
(
2018
).
9.
J.
Galego
,
F. J.
Garcia-Vidal
, and
J.
Feist
, “
Many-molecule reaction triggered by a single photon in polaritonic chemistry
,”
Phys. Rev. Lett.
119
(
13
),
136001
(
2017
).
10.
J. K.
Jaiswal
,
H.
Mattoussi
,
J. M.
Mauro
, and
S. M.
Simon
, “
Long-term multiple color imaging of live cells using quantum dot bioconjugates
,”
Nat. Biotechnol.
21
(
1
),
47
51
(
2003
).
11.
J.
Hu
,
L.-s.
Li
,
W.
Yang
,
L.
Manna
,
L.-w.
Wang
, and
A. P.
Alivisatos
, “
Linearly polarized emission from colloidal semiconductor quantum rods
,”
Science
292
(
5524
),
2060
2063
(
2001
).
12.
A.
Shabaev
and
A. L.
Efros
, “
1D exciton spectroscopy of semiconductor nanorods
,”
Nano Lett.
4
(
10
),
1821
1825
(
2004
).
13.
D.
Katz
,
T.
Wizansky
,
O.
Millo
,
E.
Rothenberg
,
T.
Mokari
, and
U.
Banin
, “
Size-dependent tunneling and optical spectroscopy of CdSe quantum rods
,”
Phys. Rev. Lett.
89
(
8
),
086801
(
2002
).
14.
S.
Ithurria
,
G.
Bousquet
, and
B.
Dubertret
, “
Continuous transition from 3D to 1D confinement observed during the formation of CdSe nanoplatelets
,”
J. Am. Chem. Soc.
133
(
9
),
3070
3077
(
2011
).
15.
A. W.
Achtstein
,
A.
Antanovich
,
A.
Prudnikau
,
R.
Scott
,
U.
Woggon
, and
M.
Artemyev
, “
Linear absorption in CdSe nanoplates: Thickness and lateral size dependency of the intrinsic absorption
,”
J. Phys. Chem. C
119
(
34
),
20156
20161
(
2015
).
16.
K. H.
Madsen
and
P.
Lodahl
, “
Quantitative analysis of quantum dot dynamics and emission spectra in cavity quantum electrodynamics
,”
New J. Phys.
15
(
2
),
025013
(
2013
).
17.
J. P.
Reithmaier
,
G.
Sęk
,
A.
Löffler
,
C.
Hofmann
,
S.
Kuhn
,
S.
Reitzenstein
,
L. V.
Keldysh
,
V. D.
Kulakovskii
,
T. L.
Reinecke
, and
A.
Forchel
, “
Strong coupling in a single quantum dot–semiconductor microcavity system
,”
Nature
432
(
7014
),
197
200
(
2004
).
18.
D. E.
Gómez
,
H.
Giessen
, and
T. J.
Davis
, “
Semiclassical plexcitonics: Simple approach for designing plexcitonic nanostructures
,”
J. Phys. Chem. C
118
(
41
),
23963
23969
(
2014
).
19.
A. M.
Armani
,
R. P.
Kulkarni
,
S. E.
Fraser
,
R. C.
Flagan
, and
K. J.
Vahala
, “
Label-free, single-molecule detection with optical microcavities
,”
Science
317
(
5839
),
783
(
2007
).
20.
B.
Min
,
E.
Ostby
,
V.
Sorger
,
E.
Ulin-Avila
,
L.
Yang
,
X.
Zhang
, and
K.
Vahala
, “
High-Q surface-plasmon-polariton whispering-gallery microcavity
,”
Nature
457
,
455
(
2009
).
21.
M.
Jiang
,
J.
Qi
,
M.
Zhang
,
Q.
Sun
,
J.
Chen
,
Z.
Chen
,
X.
Yu
,
Y.
Li
, and
J.
Tian
, “
Ultra-high quality factor metallic micro-cavity based on concentric double metal-insulator-metal rings
,”
Sci. Rep.
7
(
1
),
15663
(
2017
).
22.
H.
Leng
,
B.
Szychowski
,
M.-C.
Daniel
, and
M.
Pelton
, “
Strong coupling and induced transparency at room temperature with single quantum dots and gap plasmons
,”
Nat. Commun.
9
(
1
),
4012
(
2018
).
23.
R. D.
Artuso
and
G. W.
Bryant
, “
Optical response of strongly coupled quantum dot–metal nanoparticle systems: Double peaked Fano structure and bistability
,”
Nano Lett.
8
(
7
),
2106
2111
(
2008
).
24.
P.
Vasa
and
C.
Lienau
, “
Strong light–matter interaction in quantum emitter/metal hybrid nanostructures
,”
ACS Photonics
5
(
1
),
2
23
(
2018
).
25.
M.
Kowalewski
and
S.
Mukamel
, “
Manipulating molecules with quantum light
,”
Proc. Natl. Acad. Sci. U. S. A.
114
(
13
),
3278
3280
(
2017
).
26.
S.
D’Agostino
,
F.
Alpeggiani
, and
L. C.
Andreani
, “
Strong coupling between a dipole emitter and localized plasmons: Enhancement by sharp silver tips
,”
Opt. Express
21
(
23
),
27602
27610
(
2013
).
27.
Anonymous
, “
Proceedings of the American physical society
,”
Phys. Rev.
69
(
11-12
),
674
(
1946
).
28.
T.
Hümmer
,
F. J.
García-Vidal
,
L.
Martín-Moreno
, and
D.
Zueco
, “
Weak and strong coupling regimes in plasmonic QED
,”
Phys. Rev. B
87
(
11
),
115419
(
2013
).
29.
J. M.
Gérard
,
B.
Sermage
,
B.
Gayral
,
B.
Legrand
,
E.
Costard
, and
V.
Thierry-Mieg
, “
Enhanced spontaneous emission by quantum boxes in a monolithic optical microcavity
,”
Phys. Rev. Lett.
81
(
5
),
1110
1113
(
1998
).
30.
G. S.
Solomon
,
M.
Pelton
, and
Y.
Yamamoto
, “
Single-mode spontaneous emission from a single quantum dot in a three-dimensional microcavity
,”
Phys. Rev. Lett.
86
(
17
),
3903
3906
(
2001
).
31.
I. C.
Robin
,
R.
André
,
A.
Balocchi
,
S.
Carayon
,
S.
Moehl
,
J. M.
Gérard
, and
L.
Ferlazzo
, “
Purcell effect for CdSe/ZnSe quantum dots placed into hybrid micropillars
,”
Appl. Phys. Lett.
87
(
23
),
233114
(
2005
).
32.
X.
Wu
,
S. K.
Gray
, and
M.
Pelton
, “
Quantum-dot-induced transparency in a nanoscale plasmonic resonator
,”
Opt. Express
18
(
23
),
23633
23645
(
2010
).
33.
M.
Kroner
,
A. O.
Govorov
,
S.
Remi
,
B.
Biedermann
,
S.
Seidl
,
A.
Badolato
,
P. M.
Petroff
,
W.
Zhang
,
R.
Barbour
,
B. D.
Gerardot
,
R. J.
Warburton
, and
K.
Karrai
, “
The nonlinear Fano effect
,”
Nature
451
,
311
(
2008
).
34.
G.
Khitrova
,
H. M.
Gibbs
,
M.
Kira
,
S. W.
Koch
, and
A.
Scherer
, “
Vacuum Rabi splitting in semiconductors
,”
Nat. Phys.
2
(
2
),
81
90
(
2006
).
35.
L.
Novotny
, “
Strong coupling, energy splitting, and level crossings: A classical perspective
,”
Am. J. Phys.
78
(
11
),
1199
1202
(
2010
).
36.
R.
Stassi
,
A.
Ridolfo
,
O.
Di Stefano
,
M. J.
Hartmann
, and
S.
Savasta
, “
Spontaneous conversion from virtual to real photons in the ultrastrong-coupling regime
,”
Phys. Rev. Lett.
110
(
24
),
243601
(
2013
).
37.
F.
Beaudoin
,
J. M.
Gambetta
, and
A.
Blais
, “
Dissipation and ultrastrong coupling in circuit QED
,”
Phys. Rev. A
84
(
4
),
043832
(
2011
).
38.
S.
Savasta
,
R.
Saija
,
A.
Ridolfo
,
O.
Di Stefano
,
P.
Denti
, and
F.
Borghese
, “
Nanopolaritons: Vacuum Rabi splitting with a single quantum dot in the center of a dimer nanoantenna
,”
ACS Nano
4
(
11
),
6369
6376
(
2010
).
39.
R.
Houdré
,
C.
Weisbuch
,
R. P.
Stanley
,
U.
Oesterle
,
P.
Pellandini
, and
M.
Ilegems
, “
Measurement of cavity-polariton dispersion curve from angle-resolved photoluminescence experiments
,”
Phys. Rev. Lett.
73
(
15
),
2043
2046
(
1994
).
40.
R.
Houdré
,
R. P.
Stanley
,
U.
Oesterle
,
M.
Ilegems
, and
C.
Weisbuch
, “
Room-temperature cavity polaritons in a semiconductor microcavity
,”
Phys. Rev. B
49
(
23
),
16761
16764
(
1994
).
41.
C.
Weisbuch
,
M.
Nishioka
,
A.
Ishikawa
, and
Y.
Arakawa
, “
Observation of the coupled exciton-photon mode splitting in a semiconductor quantum microcavity
,”
Phys. Rev. Lett.
69
(
23
),
3314
3317
(
1992
).
42.
E.
Peter
,
P.
Senellart
,
D.
Martrou
,
A.
Lemaître
,
J.
Hours
,
J. M.
Gérard
, and
J.
Bloch
, “
Exciton-photon strong-coupling regime for a single quantum dot embedded in a microcavity
,”
Phys. Rev. Lett.
95
(
6
),
067401
(
2005
).
43.
T.
Yoshie
,
O. B.
Shchekin
,
H.
Chen
,
D. G.
Deppe
, and
A.
Scherer
, “
Quantum dot photonic crystal lasers
,”
Electron. Lett.
38
(
17
),
967
968
(
2002
).
44.
T.
Yoshie
,
A.
Scherer
,
J.
Hendrickson
,
G.
Khitrova
,
H. M.
Gibbs
,
G.
Rupper
,
C.
Ell
,
O. B.
Shchekin
, and
D. G.
Deppe
, “
Vacuum Rabi splitting with a single quantum dot in a photonic crystal nanocavity
,”
Nature
432
(
7014
),
200
203
(
2004
).
45.
G.
Björk
,
H.
Heitmann
, and
Y.
Yamamoto
, “
Spontaneous-emission coupling factor and mode characteristics of planar dielectric microcavity lasers
,”
Phys. Rev. A
47
(
5
),
4451
4463
(
1993
).
46.
M.
Uemoto
and
H.
Ajiki
, “
Large and well-defined Rabi splitting in a semiconductor nanogap cavity
,”
Opt. Express
22
(
19
),
22470
22478
(
2014
).
47.
N.
Le Thomas
,
U.
Woggon
,
O.
Schöps
,
M. V.
Artemyev
,
M.
Kazes
, and
U.
Banin
, “
Cavity QED with semiconductor nanocrystals
,”
Nano Lett.
6
(
3
),
557
561
(
2006
).
48.
D. E.
Gómez
,
K. C.
Vernon
,
P.
Mulvaney
, and
T. J.
Davis
, “
Surface plasmon mediated strong exciton–photon coupling in semiconductor nanocrystals
,”
Nano Lett.
10
(
1
),
274
278
(
2010
).
49.
D. E.
Gómez
,
S. S.
Lo
,
T. J.
Davis
, and
G. V. J.
Hartland
, “
Picosecond kinetics of strongly coupled excitons and surface plasmon polaritons
,”
J. Phys. Chem. B
117
(
16
),
4340
4346
(
2012
).
50.
C.
Gruber
,
A.
Trügler
,
A.
Hohenau
,
U.
Hohenester
, and
J. R.
Krenn
, “
Spectral modifications and polarization dependent coupling in tailored assemblies of quantum dots and plasmonic nanowires
,”
Nano Lett.
13
(
9
),
4257
4262
(
2013
).
51.
K.
Słowik
,
R.
Filter
,
J.
Straubel
,
F.
Lederer
, and
C.
Rockstuhl
, “
Strong coupling of optical nanoantennas and atomic systems
,”
Phys. Rev. B
88
(
19
),
195414
(
2013
).
52.
K.
Santhosh
,
O.
Bitton
,
L.
Chuntonov
, and
G.
Haran
, “
Vacuum Rabi splitting in a plasmonic cavity at the single quantum emitter limit
,”
Nat. Commun.
7
,
11823
(
2016
).
53.
N.
Zhou
,
M.
Yuan
,
Y.
Gao
,
D.
Li
, and
D.
Yang
, “
Silver nanoshell plasmonically controlled emission of semiconductor quantum dots in the strong coupling regime
,”
ACS Nano
10
(
4
),
4154
4163
(
2016
).
54.
P.
Zhang
,
G.
Song
, and
L.
Yu
, “
Optical trapping of single quantum dots for cavity quantum electrodynamics
,”
Photonics Res.
6
(
3
),
182
185
(
2018
).
55.
H.
Wang
,
H.-Y.
Wang
,
A.
Toma
,
T.-a.
Yano
,
Q.-D.
Chen
,
H.-L.
Xu
,
H.-B.
Sun
, and
R.
Proietti Zaccaria
, “
Dynamics of strong coupling between CdSe quantum dots and surface plasmon polaritons in subwavelength hole array
,”
J. Phys. Chem. Lett.
7
(
22
),
4648
4654
(
2016
).
56.
H.
Groß
,
J. M.
Hamm
,
T.
Tufarelli
,
O.
Hess
, and
B.
Hecht
, “
Near-field strong coupling of single quantum dots
,”
Sci. Adv.
4
(
3
),
eaar4906
(
2018
).
57.
T. B.
Hoang
,
G. M.
Akselrod
, and
M. H.
Mikkelsen
, “
Ultrafast room-temperature single photon emission from quantum dots coupled to plasmonic nanocavities
,”
Nano Lett.
16
(
1
),
270
275
(
2016
).
58.
J.
Vučković
,
D.
Fattal
,
C.
Santori
,
G. S.
Solomon
, and
Y.
Yamamoto
, “
Enhanced single-photon emission from a quantum dot in a micropost microcavity
,”
Appl. Phys. Lett.
82
(
21
),
3596
3598
(
2003
).
59.
Y.
Wang
,
T.
Yang
,
M. T.
Tuominen
, and
M.
Achermann
, “
Radiative rate enhancements in ensembles of hybrid metal-semiconductor nanostructures
,”
Phys. Rev. Lett.
102
(
16
),
163001
(
2009
).
60.
S. M.
Sadeghi
,
W. J.
Wing
,
R. R.
Gutha
,
R. W.
Goul
, and
J. Z.
Wu
, “
Functional metal-oxide plasmonic metastructures: Ultrabright semiconductor quantum dots with polarized spontaneous emission and suppressed Auger recombination
,”
Phys. Rev. Appl.
11
(
2
),
024045
(
2019
).
61.
A. V.
Akimov
,
A.
Mukherjee
,
C. L.
Yu
,
D. E.
Chang
,
A. S.
Zibrov
,
P. R.
Hemmer
,
H.
Park
, and
M. D.
Lukin
, “
Generation of single optical plasmons in metallic nanowires coupled to quantum dots
,”
Nature
450
,
402
(
2007
).
62.
A. G.
Curto
,
G.
Volpe
,
T. H.
Taminiau
,
M. P.
Kreuzer
,
R.
Quidant
, and
N. F.
van Hulst
, “
Unidirectional emission of a quantum dot coupled to a nanoantenna
,”
Science
329
(
5994
),
930
(
2010
).
63.
P.
Törmä
and
W. L.
Barnes
, “
Strong coupling between surface plasmon polaritons and emitters: A review
,”
Rep. Prog. Phys.
78
(
1
),
013901
(
2014
).
64.
A.
Manjavacas
,
F. J. G.
de Abajo
, and
P.
Nordlander
, “
Quantum plexcitonics: Strongly interacting plasmons and excitons
,”
Nano Lett.
11
(
6
),
2318
2323
(
2011
).
65.
A.
Trügler
and
U.
Hohenester
, “
Strong coupling between a metallic nanoparticle and a single molecule
,”
Phys. Rev. B
77
(
11
),
115403
(
2008
).
66.
H.
Varguet
,
B.
Rousseaux
,
D.
Dzsotjan
,
H. R.
Jauslin
,
S.
Guérin
, and
G.
Colas des Francs
, “
Dressed states of a quantum emitter strongly coupled to a metal nanoparticle
,”
Opt. Lett.
41
(
19
),
4480
4483
(
2016
).
67.
C.
Van Vlack
,
P. T.
Kristensen
, and
S.
Hughes
, “
Spontaneous emission spectra and quantum light-matter interactions from a strongly coupled quantum dot metal-nanoparticle system
,”
Phys. Rev. B
85
(
7
),
075303
(
2012
).
68.
O.
Bitton
,
N.
Gupta Satyendra
, and
G.
Haran
, “
Quantum dot plasmonics: From weak to strong coupling
,”
Nanophotonics
8
,
559
(
2019
).
69.
A.
Muller
,
E. B.
Flagg
,
P.
Bianucci
,
X. Y.
Wang
,
D. G.
Deppe
,
W.
Ma
,
J.
Zhang
,
G. J.
Salamo
,
M.
Xiao
, and
C. K.
Shih
, “
Resonance fluorescence from a coherently driven semiconductor quantum dot in a cavity
,”
Phys. Rev. Lett.
99
(
18
),
187402
(
2007
).
70.
M.
Nomura
,
N.
Kumagai
,
S.
Iwamoto
,
Y.
Ota
, and
Y.
Arakawa
, “
Laser oscillation in a strongly coupled single-quantum-dot–nanocavity system
,”
Nat. Phys.
6
(
4
),
279
(
2010
).
71.
D.
Zhao
,
Y.
Gu
,
J.
Wu
,
J.
Zhang
,
T.
Zhang
,
B. D.
Gerardot
, and
Q. J. P. R. B.
Gong
, “
Quantum-dot gain without inversion: Effects of dark plasmon-exciton hybridization
,”
Phys. Rev. B
89
(
24
),
245433
(
2014
).
72.
K.
Hennessy
,
A.
Badolato
,
M.
Winger
,
D.
Gerace
,
M.
Atatüre
,
S.
Gulde
,
S.
Fält
,
E. L.
Hu
, and
A.
Imamoğlu
, “
Quantum nature of a strongly coupled single quantum dot–cavity system
,”
Nature
445
,
896
(
2007
).
73.
D. E.
Gomez
,
K. C.
Vernon
,
P.
Mulvaney
, and
T. J.
Davis
, “
Coherent superposition of exciton states in quantum dots induced by surface plasmons
,”
Appl. Phys. Lett.
96
(
7
),
073108
(
2010
).
74.
N. C.
Giebink
,
G. P.
Wiederrecht
, and
M. R.
Wasielewski
, “
Strong exciton-photon coupling with colloidal quantum dots in a high-Q bilayer microcavity
,”
Appl. Phys. Lett.
98
(
8
),
081103
(
2011
).
75.
K.
Karrai
,
R. J.
Warburton
,
C.
Schulhauser
,
A.
Högele
,
B.
Urbaszek
,
E. J.
McGhee
,
A. O.
Govorov
,
J. M.
Garcia
,
B. D.
Gerardot
, and
P. M. J. N.
Petroff
, “
Hybridization of electronic states in quantum dots through photon emission
,”
Nature
427
(
6970
),
135
(
2004
).
76.
M.
Winger
,
T.
Volz
,
G.
Tarel
,
S.
Portolan
,
A.
Badolato
,
K. J.
Hennessy
,
E. L.
Hu
,
A.
Beveratos
,
J.
Finley
,
V.
Savona
, and
A.
Imamoğlu
, “
Explanation of photon correlations in the far-off-resonance optical emission from a quantum-dot–cavity system
,”
Phys. Rev. Lett.
103
(
20
),
207403
(
2009
).
77.
F. P.
Laussy
,
E.
del Valle
, and
C.
Tejedor
, “
Strong coupling of quantum dots in microcavities
,”
Phys. Rev. Lett.
101
(
8
),
083601
(
2008
).
78.
I.
Diniz
,
S.
Portolan
,
R.
Ferreira
,
J. M.
Gérard
,
P.
Bertet
, and
A.
Auffèves
, “
Strongly coupling a cavity to inhomogeneous ensembles of emitters: Potential for long-lived solid-state quantum memories
,”
Phys. Rev. A
84
(
6
),
063810
(
2011
).
79.
R.
Esteban
,
J.
Aizpurua
, and
G. W.
Bryant
, “
Strong coupling of single emitters interacting with phononic infrared antennae
,”
New J. Phys.
16
(
1
),
013052
(
2014
).
80.
F.
Herrera
and
F. C.
Spano
, “
Cavity-controlled chemistry in molecular ensembles
,”
Phys. Rev. Lett.
116
(
23
),
238301
(
2016
).
81.
X.
Zhong
,
T.
Chervy
,
L.
Zhang
,
A.
Thomas
,
J.
George
,
C.
Genet
,
J. A.
Hutchison
, and
T. W.
Ebbesen
, “
Energy transfer between spatially separated entangled molecules
,”
Angew. Chem., Int. Ed.
56
(
31
),
9034
9038
(
2017
).
82.
A.
Thomas
,
J.
George
,
A.
Shalabney
,
M.
Dryzhakov
,
S. J.
Varma
,
J.
Moran
,
T.
Chervy
,
X.
Zhong
,
E.
Devaux
, and
C.
Genet
, “
Ground-state chemical reactivity under vibrational coupling to the vacuum electromagnetic field
,”
Angew. Chem., Int. Ed.
55
(
38
),
11462
11466
(
2016
).
83.
M. S.
Kodaimati
,
K. P.
McClelland
,
C.
He
,
S.
Lian
,
Y.
Jiang
,
Z.
Zhang
, and
E. A.
Weiss
, “
Challenges in colloidal photocatalysis and some strategies for addressing them
,”
Inorg. Chem.
57
,
3659
(
2018
).
84.
A.
Ben-Moshe
,
A.
Teitelboim
,
D.
Oron
, and
G.
Markovich
, “
Probing the interaction of quantum dots with chiral capping molecules using circular dichroism spectroscopy
,”
Nano Lett.
16
(
12
),
7467
7473
(
2016
).
85.
D.
Martin-Cano
,
H. R.
Haakh
, and
N.
Rotenberg
, “
Chiral emission into nanophotonic resonators
,”
ACS Photonics
6
(
4
),
961
966
(
2019
).
86.
R. J.
Coles
,
D. M.
Price
,
B.
Royall
,
E.
Clarke
,
M. S.
Skolnick
,
A. M.
Fox
, and
M. N.
Makhonin
, “
Path-dependent initialization of a single quantum dot exciton spin in a nanophotonic waveguide
,”
Phys. Rev. B
95
(
12
),
121401
(
2017
).
87.
M. T.
Frederick
and
E. A.
Weiss
, “
Relaxation of exciton confinement in CdSe quantum dots by modification with a conjugated dithiocarbamate ligand
,”
ACS Nano
4
(
6
),
3195
3200
(
2010
).
88.
M. T.
Frederick
,
V. A.
Amin
,
L. C.
Cass
, and
E. A.
Weiss
, “
A molecule to detect and perturb the confinement of charge carriers in quantum dots
,”
Nano Lett.
11
(
12
),
5455
5460
(
2011
).
89.
M. T.
Frederick
,
V. A.
Amin
,
N. K.
Swenson
,
A. Y.
Ho
, and
E. A.
Weiss
, “
Control of exciton confinement in quantum dot–organic complexes through energetic alignment of interfacial orbitals
,”
Nano Lett.
13
(
1
),
287
292
(
2013
).
90.
D.
Englund
,
A.
Majumdar
,
A.
Faraon
,
M.
Toishi
,
N.
Stoltz
,
P.
Petroff
, and
J.
Vučković
, “
Resonant excitation of a quantum dot strongly coupled to a photonic crystal nanocavity
,”
Phys. Rev. Lett.
104
(
7
),
073904
(
2010
).
91.
J.-W.
Jeon
,
J.
Zhou
,
J. A.
Geldmeier
,
J. F.
Ponder
,
M. A.
Mahmoud
,
M.
El-Sayed
,
J. R.
Reynolds
, and
V. V.
Tsukruk
, “
Dual-responsive reversible plasmonic behavior of core–shell nanostructures with pH-sensitive and electroactive polymer shells
,”
Chem. Mater.
28
,
7551
(
2016
).
92.
K.-D.
Park
,
M. A.
May
,
H.
Leng
,
J.
Wang
,
J. A.
Kropp
,
T.
Gougousi
,
M.
Pelton
, and
M. B.
Raschke
, “
Tip-enhanced strong coupling spectroscopy, imaging, and control of a single quantum emitter
,”
Sci. Adv.
5
(
7
),
eaav5931
(
2019
).