Strong and long-range radiative interaction between resonant transitions
Abstract
Enhancing the radiative interaction between quantum electronic transitions is of general interest. There are two important properties of radiative interaction: the range and the strength. There has been a trade-off between the range and the strength observed in the literature. Such apparent trade-off arises from the dispersion relation of photonic environments. A general recipe is developed to overcome such trade-off and to simultaneously enhance the range and the strength of radiative interactions.
I Introduction
The radiative interaction between electronic quantum transitions changes both the energy and lifetime of transitions. It leads to phenomena such as Frster energy transfer [1, 2], super-radiance [3, 4, 5], and collective Lamb shift [6]. Because radiative coupling is short range, these phenomena can only be observed when the distance between transitions is not much longer than the wavelength of the light that mediates the interaction. Researchers have been interested in enhancing the strength and extending the range of radiative interactions, with applications in energy sciences and quantum technologies. While lots of progress has been made in the past decades, we observed that there has a been a trade-off in many methods studied in the literature [7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22]. The trade-off occurs between the strength and the range: interactions with a long range are often weak in strength while strong interactions are very short range.
We first discuss how we define and quantify the range and strength of radiative interaction between quantum transitions. The radiative interaction includes two parts: via propagating modes and via evanescent modes. Here we focus on the propagating modes because they are generally responsible for long-range interaction. We consider two quantum transitions in vacuum. Figure 1(b) plots the real photon part of radiative coupling coefficient as a function of the distance between two quantum transitions in vacuum (see Appendix C for detailed derivation on the radiative interaction), which describes the rate of energy exchange. The strength is defined by the maximum value of the coupling, which is commonly realized when the distance is close to zero. The range is defined by the distance at which the coupling decays to its half maximum value. For example, the interaction strength and range in vacuum are the spontaneous emission rate and , respectively.

Radiative interaction via real photon exchange is mediated through optical modes in the continuum. A general strategy to enhance radiative interaction strength and range is to place transitions in engineered photonic environments, where the optical modes can be drastically different from those in vacuum. This strategy was used in many interesting works that enhance the range [7, 8, 9, 10, 11, 4, 13] or the strength [20, 21, 28, 29]. For example, zero-index materials can extend the range of coupling [7, 8, 9, 10, 11] to , where the enhancement factor can be a few hundreds. However, the strength decrease by a factor of . The greater the range is, the weaker the strength becomes. Another interesting example exploits local plasmonic resonance [26, 30, 31]. The strength can be enhanced. However, the range becomes shorter [26] due to the fast decay of plasmonic modes. There are other interesting works that use near fields [32, 15, 33, 34, 25] or far fields around the Weyl point [22] in photonic crystals. When extending the cases to three-dimensional space, they all exhibit similar trade-off between range and strength. The only exception is one-dimensional space, such as in a waveguide or along the one particular direction of photonic structures where the interaction range can be extended without compromising the strength [35, 13, 4, 5].
Figure 1(c) surveys recent works in which a trade-off between the interaction range and the strength can be observed. Although the specific mechanism for each case could vary, a common physics emerges in many of these cases. By identifying what contributes to the apparent trade-off, we show how to simultaneously enhance the strength and the range. The red hexagon in Fig. 1(c) for example has a - and -time enhancement of the strength and the range, respectively. Such enhancement could be very useful for quantum systems that use radiative interaction as the main mechanism for energy or information transduction.
II Long-range interaction
We choose Dirac point because its isosurface can change rapidly while maintaining a relatively constant group velocity, which helps to isolate the effect of the contribution from isosurface. The Hamiltonian near a Dirac point is given by , where is the Pauli matrices and is the isotropic group velocity. This Hamiltonian results in a linear dispersion relationship as , where is the location of the Dirac point in momentum space.
We study the interaction between two identical transitions placed inside the photonic crystal with the Dirac dispersion [36, 37] as shown in Fig. 1(a). We show three cases with their transition frequencies labeled in Fig. 2(c). From the point iii to point i in Fig. 2(c), the frequency gradually approaches the Dirac point. The radiative interaction as a function of the distance is shown in Fig. 2(d). The calculation is available in Appendix D. When the transition frequency is far away from the Dirac point, i.e. frequency-iii in Fig. 2(c), and the blue curve in Fig. 2(d), the strength and the range are not significantly different from those in vacuum. As we move the transition frequency close to the Dirac point, we see enhanced range at frequency-ii and frequency-i. In the case very close to the Dirac point (green curve in Fig. 2(d), the interaction decays slowly as the distance between the transitions increases. Conversely, the strength is weakened as the transition frequency approaches the Dirac point. Note that the green curve (point i) is displayed with a multiplication factor in Fig. 2(d). We summarize the strength and range of three cases in Fig. 2(b), and the trade-off is clearly observed: while the strength increases, the range decreases.
The trade-off observed here is quite general. We now discuss its origin. We consider a general continuum with a dispersion relationship , where is the wavevector. We focus on optical modes around the transition frequency because they are the most important for mediating the radiative interaction. In the momentum space, these modes form an isosurface defined by . For example, the isosurface in vacuum is a sphere with a radius of . In engineered photonic environments, the shape and size of isosurface can be quite different. We recently showed that the isosurface plays a significant role in radiative interactions [22, 38].


The radiative interaction can be written in the following form [23, 22]
(1) |
where the polarization factor and is the group velocity. The integration is performed over the isosurface . is the electric field of mode, . are the dipole moments of the two transitions, is the distance between the two transitions, and is the dielectric constant and is the refractive index.
Now we are ready to explain the range of interaction. The term is a fast-oscillating term. The integration of this fast-oscillating term cancels when , the distance between the two transitions, is large. This is why the interaction vanishes for large . In addition to , the photonic environment also plays an important role through the shape of the isosurface . To see this impact, we fix a distance between two transitions and consider photonic environments with different . If we have a large isosurface, the integration of cancels more compared to a small isosurface. As a result, a large isosurface leads to a shorter-range interaction than a small isosurface. This is what we observed in Fig. 2(b) and 2(c). The isosurface, which is an iso-contour in this two-dimensional space, is smaller around Dirac point, leading to a long interaction range. Away from the Dirac point, the isosurface is larger, resulting in a short interaction range.
Now we understand one aspect of the interaction-the range. We can examine the other side of the trade-off. The size of isosurface has a quite different impact on interaction strength. Larger isosurface generally leads to a greater interaction strength. The maximum interaction strength is obtained at when becomes unity. A large isosurface has a large integration area, and thus a greater integration value. This relationship can also be understood using the optical density of states (DOS), which is proportional to the size of isosurface. The larger the DOS, the greater the spontaneous emission rate, which is directly related to the interaction strength.
Now the trade-off in Fig. 2(d) can be easily understood. Close to the Dirac point, the isosurfaces become smaller and smaller (Fig. 2(b)). Correspondingly, the range of interaction increases. However, smaller isosurfaces reduce the density of states, which decreases the spontaneous emission rate and the interaction strength.
The understanding of interaction above is general for a photonic environment with an explicit dispersion relationship. The trade-off caused by isosurfaces is built into the mathematic integration Eq. (1) that determines the interaction. To mitigate this trade-off, we have to look into other physical mechanisms that contribute to the interaction: the first is the local field effect in the term , and the other is the group velocity in the term . By carefully balancing the isosurface, local field, and group velocity in judiciously designed photonic environments, we can overcome the apparent trade-off to realize simultaneous enhancement of interaction strength and range.
Our strategy starts by considering a photonic environment with a small isosurface because this is necessary to obtain a long range. However, it comes with a weak interaction strength. We can then use strong local fields and small group velocity to offset the negative impact from the small isosurface. The result is simultaneous enhancement of strength and range by many orders of magnitude. This strategy can be applied to many different types of engineered photonic media. Here we choose one of the simplest implementations in a Fabry-Pérot cavity [39, 40] to illustrate this method.
The cavity structure consisting two mirrors is shown in Fig. 3(a). The mirrors are extremely large and restrict the available modes and reduces the spherical isosurface in vacuum to iso-contours such as those rings and points shown in Fig. 3(c)-(f). Here we are particularly interested in the frequency region around the fundamental mode , where is the cavity length and is the refractive index of material inside the cavity. The isosurface of the fundamental mode reduces to two points (Fig. 3(f)). In comparison, high-order resonant modes consist of ring-shape isosurfaces (Fig. 3(d,e)). The diminishing isosurface at the fundamental mode is ideal for extending the interaction range. Unlike Dirac points and Weyl points, here the group velocity is zero in the - plane, which helps to off-set the negative impact of small isosurface on interaction strength.
We perform direct calculation of the resonant dipole-dipole interaction inside a Fabry-Pérot cavity (see more details in Appendix A). Fig. 3(b) shows the strength and range comparing the fundamental Fabry-Pérot modes to two other modes in the cavity. We can see the interaction range is longest for the fundamental mode because its smallest isosurface. Here we managed to maintain the strength of interaction because of the low group velocity.

III Enhancement of interaction strength
The fundamental mode in a Fabry-Pérot cavity provides a long interaction range while maintaining an interaction strength similar to that in vacuum. Next, we explore a second mechanism to greatly enhance the strength beyond that of vacuum using the local field effect [41]. This effect occurs at the interface between two materials of different refractive indices when a small volume of the lower index material is embedded in the higher index material. Polarization charge accumulated at the interfaces creates enhanced local electric field [41, 42, 43], which can be used to enhance the field in Eq. (1). To illustrate this effect, we consider an ellipsoidal defect of low index in a material of high index . The dimension of the ellipsoid is much smaller than wavelength (typically smaller than ). Fig. 4(a) shows the field distribution under plane wave illumination with electric field polarized in direction. The intensity of the field inside is enhanced by , where is the geometrical factor that depends on the major radius and minor radius of the ellipsoid (see details in Appendix B). Figure 4(b) shows the enhancement of the field intensity defects with different index contrast and geometrical ratio. With an index contrast around , times enhancement can be realized for elongated defect shapes. Since it is small with a deep sub wavelength size, the defect does not significantly affect the modes in the continuum which are mostly determined by larger and wavelength-scale structures. The local field effect explores a completely different physics of enhancement than the group velocity and isosurface. It does not interfere with other enhancement mechanisms.
Next, we combine the effects of local field enhancement and the fundamental mode of a Fabry-Pérot cavity to demonstrate a case of long and strong interaction. The cavity is filled with a high-index material . Inside the cavity, we have two small air holes as the low-index defects. We consider one resonant dipole inside each defect as shown in Fig. 4(c). The two dipoles are aligned to the short axis of the ellipsoidal defect. The radiative interaction between two quantum transitions is given by , where is the electric dyadic Green’s function of the photonic environment. Then dyadic Green’s function can be obtained numerically by a full-wave simulation shown in Appendix D. The numerical results are marked as the red diamonds in Fig. 4 (d), which agree with our theoretical estimation in Appendix B. For the cavity with a quality factor of , the same as the cavity in Fig. 3, the interaction strength can reach up to about . Meanwhile, the interaction range is times the range in vacuum. We also examine the results from non-Markovian calculation, which shows fairly consistent conclusions as the Markovian approximation used here. The details can be found in Appendix E. The retardation effect is included in the full electrodynamic treatment [44].
IV Discussion
To conclude, we discuss the origin of the apparent trade-off between the interaction strength and interaction range. We analyze three main factors that contribute to radiative interaction: the isosurface of photonic environment, the local field, and the group velocity. The isosurface leads to the trade-off between the strength and the range of interaction. The local field and group velocity can be modified relatively independently. With this recognition, we developed a recipe for long and strong range interaction. We start with a small isosurface to guarantee a slow decay with a long range. Then we engineer the photonic environment to exploit the local field and the group velocity. It leads to long and strong interaction between two resonant transitions. It would be interesting to study a network of quantum transitions where they interact with each other strongly over extended distance, which could lead to quite different results from networks that only consider nearest neighbor interactions.
Acknowledgments
This work was supported by the National Science Foundation (NSF) through the University of Wisconsin Materials Research Science and Engineering Center DMR-1720415. L. Y. was also supported by the National Key Research and Development Program of China 2019YFA0308100.
Appendix A Theory of strong-and-long range interaction
We consider the resonant dipole-dipole interaction between electric dipolar quantum transitions (QTs). The Hamiltonian of two identical QTs in an arbitrary photonic environment is given by
(2) |
They are explicitly written as [45] ()
(3) |
where and are the transition frequency and raising (lowering) operator of two-level QTs. and are the frequency and creation (annihilation) operator of photon.
(4) |
is the coupling between QTs and photonic mode , where the mode index contains the information of photonic wavevector and polarization . is the transition dipole moment of QTs. is the photon volume and is the refractive index of the photonic medium.
The transition probability from initial to final states is given by the Fermi’s Golden rule , where the transition matrix element can describe the resonant dipole-dipole interaction between two QTs. For the weak light-matter interaction, it can be written as the second-order form:
(5) |
Here, and denote initial and final states, where‘’ and ‘’ in the Dirac bracket notions represent excited and ground states, respectively, and the number ‘’ or ‘’ is the photon number in the photonic environment. The intermediate state has two options: with energy and with energy . The energy of the initial state is . Since two identical QTs are considered, we have . Then, Eq. (3) can be written as
(6) |
Analytical solutions of radiative interactions via propagating modes in a cavity — We consider a FabryPérot with two PEC plates. The volume is given by
(7) |
where is the thickness of the cavity and . Then, the summation over mode index can be written as an integral:
(8) |
Also, utilizing the relationship
(9) |
the interaction is written as
(10) |
where is the Cauchy principal value and . The polarization factor is written as
(11) |
where the dipole moment and (degenerate) electric polarizations of mode can written as
(12) |
where . The radiative interaction is given by
(13) |
where , and . is the circular isosurface with a same frequency . The group velocity in the plane is
(14) |
Then, the radiative interaction is written as
(15) |
where
(16) |
where
(17) |
Then, we consider the fundamental-mode cavity as shown in Fig. 3(f). The isosurface reduces to a pair of points. The thickness of the cavity is . Also, assume and . Then we have
(18) |
where is the spontaneous decay rate in vacuum.

The interaction via evanescent modes — The interaction includes real and imaginary parts in Eq. (10). We already have the imaginary part in Eq. (10). In the opposite, the real part of the interaction is responsible for the evanescent modes. From Eq. (10), we have
(19) |
Then, the interaction via evanescent modes and radiative modes satisfies the Kramers-Kronig relationship. The evanescent interaction scaling with the inter-QT distance is same as the radiative case, but their phases are opposite. For the fundamental mode of a perfect cavity, the interaction via evanescent modes is diverging while the radiative interaction is a constant as shown in Eq. (18).
Appendix B Local field enhancement
Our proposal to enhance interaction strength is based on a low-refractive-index defect, which only enhances the local field without any other negtive influence on the interaction range.
We consider an air defect with a radius along axis and two radiuses along axises in Fig. 4 (a) and its induced dipole moment along axis is given by
(20) |
Then, the polarizability is written as
(21) |
where the geometrical factor is with . Since we consider the size of the defect is much smaller than the wavelength (), the electric field is smooth inside the defect. Thus, we have the equation as
(22) |
where is the volume of the ellipsoid defect and is component of electric field inside the defect. Thus, the enhancement factor along axis is given by
(23) |
where and are the dielectric constants of defect and background, respectively. is the geometrical factor of the defect normal to - plane. For a spherical geometry, i.e. , the geometrical factor is and then the enhancement factor is . If , the enhancement factor will be greatly enhanced. For example, in Fig. 5 (c) with the diameter ratio .
For a two-dimensional case, the ellipsoid is reduced to an elliptical defect. Then, the geometrical factor becomes with .
Then, two dipoles are placed inside the air defects with the dipole orientation parallel to the axis. The interaction then can be written as
(24) |
where is the radiative interaction without low-index defects. We show the radiative interaction as a function separation distance for different geometrical defects in Fig. 5 (b-f). Also, in Fig 5 (h) and (g), we plot the interaction strength as a function of dielectric ratio and geometrical ratio, respectively. As we can see, the enhancement of interaction strength does not depend on any other range-related parameters.
We also do full-wave simulation for the local field enhancement by solving the Maxwell’s equations. As shown in Fig. 5, two air defects are placed in a high-index material. In the simulation, we fix the background index as and the long radius of the ellipsoid defect is . The short one is tuned as , , and . The left point dipole is set as the source and then we can get the electric field at the other dipole location. The electric field calculated from the full-wave simulation matches the theoretical estimations, as shown in Fig. 5 (h).
Appendix C Summary of interaction in various photonic environments
In this section, we will introduce the detailed information about the interaction range and interaction strength via propagating modes in Fig. 1 (c).
Uniform media and Index-near-zero metamaterials — In isotropic uniform media, the dispersion relation is given by with is the refractive index of the medium. The interaction via propagating modes is written as
(25) |
where and . The interaction strength is and the interaction range is , where is the spontaneous decay rate in vacuum and is the interaction range in vacuum. We plot the uniform medium with refractive index from to and the index-near-zero material [7, 13] with effective refractive index from to in Fig. 1, respectively. The dipole orientations of quantum transitions are same and they are vertical to .
Bandedge of Photonic crystals — The dispersion relation of the period dielectric photonic crystal is written as [24]
(26) |
with the condition of , where is the lattice constant and is the radius of dielectric region in a unit cell. Then, we have the size of isosurface and the group velocity for a fix frequency from the dispersion relation. Then, we substitute these quantities into Eq. (9). The data in Fig. 1 (c) is plotted with the refractive index from to and the frequency near the band edge.
Hyperbolic meta-materials — The dispersion relation of the hyperbolic material [21] is with . The interaction between quantum transitions can be calculated from the dyadic Green’s function. Because the interaction via propagating modes is diverging at , the interaction strength is infinite and the interaction range is zero. We assume the interaction strength is and the interaction range is . The data in Fig. 1 is based on two sets of parameters and with imaginary part of dielectric constant from to .
Plasmonic plates — The electric field of air-silver interface [26] is , where is the magnitude of incident dipole radiating field, is the coupling efficiency between the quantum transition and the surface plasmon polariton (SPP) modes, is the SPP wave vector and . is the permittivity of silver. The interaction via propagating modes is given by
(27) |
In Fig. 1 (c), the parameters are , , , and ; , , , and ; , , , and ; , , , and , respectively.
Weyl photonic crystals — In the Weyl photonic crystal, the interaction strength and the interaction range via propagating modes are from the data in Ref. [22]. The dielectric constant of the double-gyroid structure is .
Appendix D Numerical method of Green’s function
The cavity can be considered as a multi-layer structure. Then, the interaction between two QTs in a cavity is given by
(28) |
where is the dyadic Green’s function. Its matrix form is written as
(29) |
where the element is the -component electric field at with a point source with dipole orientation at and .
Numerical method in multi-layered structures (FabryPérot cavities) — For a multilayered structure, the dyadic Green’s function is given by [46]
(30) |
with . is the Bessel function of the first kind of order .
(31) |
where
(32) |
The transmission-line Green’s functions are given by
(33) |
where
(34) |
with . The coefficients are
(35) |
The generalized reflection coefficients are given by
(36) |
The Fresnel reflection coefficients are given by
(37) |
Full-wave simulation for dyadic Green’s function — For complex structures, the full-wave simulation is needed to calculate the dyadic Green’s function. The matrix element of the dyadic Green’s function is defined by
(38) |
where is the source -polarized dipole moment and is the component of electric field at due to the dipole radiation.
Interaction in Dirac photonic slabs — We use a triangular lattice with two basis and . The dielectric constant of the slab is and hole radius is , where is the lattice constant. Numerically, we use the MPB software package [47] to calculate eigen modes of the Dirac photonic slab in Fig. 2. We set the resolution in the unit cell as . Then, the frequency of the Dirac points is , which falls between th and th bands. Then we calculate the radiative interaction via propagating modes by the definition of Eq. (10).
Appendix E Non-Markovian dynamics between two quantum transitions

In this section, we introduce the dynamics between two quantum transitions in the non-Markovian limit by solving the time-dependent Schrodinger equation
(39) |
where the Hamiltonian has been given by Eq. (2) and the wavefunction is
(40) |
The propability amplitudes of quantum transitions and are given by
(41) |
where the memory kernel is written as
(42) |
Here, is the cutoff function due to the limit of Bohr frequency [48] , and with . We consider a pair of quantum transitions with a separation distance of . Fig. 6 shows the populations of quantum transitions and as a function of time for our proposed structure of a cavity doped with low-index defects, cavity, and vacuum, respectively. For the cavity with field-enhanced defect, the strong-and-long interaction results in fast oscillation between two quantum transitions.
References
- Clegg [1995] R. M. Clegg, Fluorescence resonance energy transfer, Current opinion in biotechnology 6, 103 (1995).
- Garcia-Vidal and Feist [2017] F. J. Garcia-Vidal and J. Feist, Long-distance operator for energy transfer, Science 357, 1357 (2017).
- Scully and Svidzinsky [2009] M. O. Scully and A. A. Svidzinsky, The super of superradiance, Science 325, 1510 (2009).
- Solano et al. [2017] P. Solano, P. Barberis-Blostein, F. K. Fatemi, L. A. Orozco, and S. L. Rolston, Super-radiance reveals infinite-range dipole interactions through a nanofiber, Nature communications 8, 1857 (2017).
- Sinha et al. [2020] K. Sinha, P. Meystre, E. A. Goldschmidt, F. K. Fatemi, S. L. Rolston, and P. Solano, Non-markovian collective emission from macroscopically separated emitters, Phys. Rev. Lett. 124, 043603 (2020).
- Meir et al. [2014] Z. Meir, O. Schwartz, E. Shahmoon, D. Oron, and R. Ozeri, Cooperative lamb shift in a mesoscopic atomic array, Phys. Rev. Lett. 113, 193002 (2014).
- Fleury and Alu [2013] R. Fleury and A. Alu, Enhanced superradiance in epsilon-near-zero plasmonic channels, Physical Review B 87, 201101 (2013).
- Mahmoud et al. [2017] A. Mahmoud, I. Liberal, and N. Engheta, Dipole-dipole interactions mediated by epsilon-and-mu-near-zero waveguide supercoupling, Optical Materials Express 7, 415 (2017).
- Liberal and Engheta [2018] I. Liberal and N. Engheta, Multiqubit subradiant states in n-port waveguide devices: -and--near-zero hubs and nonreciprocal circulators, Physical Review A 97, 022309 (2018).
- Gundogdu et al. [2015] F. T. Gundogdu, A. E. Serebryannikov, A. Ozgur Cakmak, and E. Ozbay, Asymmetric transmission in prisms using structures and materials with isotropic-type dispersion, Optical Express 23, 24120 (2015).
- Serebryannikov et al. [2019] A. E. Serebryannikov, H. Hajian, M. Krawczyk, G. A. Vandenbosch, and E. Ozbay, Embedded arrays of annular apertures with multiband near-zero-index behavior and demultiplexing capability at near-infrared, Optical Materials Express 9, 3169 (2019).
- Sato et al. [2012] Y. Sato, Y. Tanaka, J. Upham, Y. Takahashi, T. Asano, and S. Noda, Strong coupling between distant photonic nanocavities and its dynamic control, Nature Photonics 6, 56 (2012).
- Liberal and Engheta [2017] I. Liberal and N. Engheta, Zero-index structures as an alternative platform for quantum optics, Proceedings of the National Academy of Sciences 114, 822 (2017).
- Polesana et al. [2007] P. Polesana, A. Couairon, D. Faccio, A. Parola, M. A. Porras, A. Dubietis, A. Piskarskas, and P. Di Trapani, Observation of conical waves in focusing, dispersive, and dissipative kerr media, Phys. Rev. Lett. 99, 223902 (2007).
- Hughes [2007] S. Hughes, Coupled-cavity qed using planar photonic crystals, Phys. Rev. Lett. 98, 083603 (2007).
- Yao and Hughes [2009] P. Yao and S. Hughes, Macroscopic entanglement and violation of bell¡¯s inequalities between two spatially separated quantum dots in a planar photonic crystal system, Optics express 17, 11505 (2009).
- Minkov and Savona [2013] M. Minkov and V. Savona, Radiative coupling of quantum dots in photonic crystal structures, Phys. Rev. B 87, 125306 (2013).
- Hung et al. [2013] C.-L. Hung, S. M. Meenehan, D. E. Chang, O. Painter, and H. J. Kimble, Trapped atoms in one-dimensional photonic crystals, New Journal of Physics 15, 083026 (2013).
- Vasco et al. [2014] J. P. Vasco, P. S. S. Guimarães, and D. Gerace, Long-distance radiative coupling between quantum dots in photonic crystal dimers, Phys. Rev. B 90, 155436 (2014).
- Biehs et al. [2016] S.-A. Biehs, V. M. Menon, and G. Agarwal, Long-range dipole-dipole interaction and anomalous förster energy transfer across a hyperbolic metamaterial, Physical Review B 93, 245439 (2016).
- Cortes and Jacob [2017] C. L. Cortes and Z. Jacob, Super-coulombic atom–atom interactions in hyperbolic media, Nature communications 8, 14144 (2017).
- Ying et al. [2019] L. Ying, M. Zhou, M. Mattei, B. Liu, P. Campagnola, R. H. Goldsmith, and Z. Yu, Extended range of dipole-dipole interactions in periodically structured photonic media, Phys. Rev. Lett. 123, 173901 (2019).
- Lehmberg [1970] R. H. Lehmberg, Radiation from an -atom system. i. general formalism, Phys. Rev. A 2, 883 (1970).
- John and Wang [1991] S. John and J. Wang, Quantum optics of localized light in a photonic band gap, Phys. Rev. B 43, 12772 (1991).
- Notararigo et al. [2018] V. Notararigo, R. Passante, and L. Rizzuto, Resonance interaction energy between two entangled atoms in a photonic bandgap environment, Scientific Reports 8, 5193 (2018).
- Zhou et al. [2011] F. Zhou, Y. Liu, and Z.-Y. Li, Surface-plasmon-polariton-assisted dipolešcdipole interaction near metal surfaces, Optics Letters 36, 1969 (2011).
- García-Elcano et al. [2020] I. n. García-Elcano, A. González-Tudela, and J. Bravo-Abad, Tunable and robust long-range coherent interactions between quantum emitters mediated by weyl bound states, Phys. Rev. Lett. 125, 163602 (2020).
- Hood et al. [1998] C. J. Hood, M. S. Chapman, T. W. Lynn, and H. J. Kimble, Real-time cavity qed with single atoms, Phys. Rev. Lett. 80, 4157 (1998).
- 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, 1 (2014).
- Gonzalez-Tudela et al. [2011] A. Gonzalez-Tudela, D. Martin-Cano, E. Moreno, L. Martin-Moreno, C. Tejedor, and F. J. Garcia-Vidal, Entanglement of two qubits mediated by one-dimensional plasmonic waveguides, Physical review letters 106, 020501 (2011).
- Zhang et al. [2012] S. Zhang, Z. Ye, Y. Wang, Y. Park, G. Bartal, M. Mrejen, X. Yin, and X. Zhang, Anti-hermitian plasmon coupling of an array of gold thin-film antennas for controlling light at the nanoscale, Phys. Rev. Lett. 109, 193902 (2012).
- André and Lukin [2002] A. André and M. D. Lukin, Manipulating light pulses via dynamically controlled photonic band gap, Phys. Rev. Lett. 89, 143602 (2002).
- Douglas et al. [2017] J. S. Douglas, H. Habibian, C. L. Hung, A. V. Gorshkov, H. J. Kimble, and D. E. Chang, Quantum many-body models with cold atoms coupled to photonic crystals, Nature Photonics 8, 1388 (2017).
- Gonz¨¢lez-Tudela et al. [2015] A. Gonz¨¢lez-Tudela, C.-L. Hung, D. E. Chang, J. I. Cirac, and H. J. Kimble, Subwavelength vacuum lattices and atom¨catom interactions in two-dimensional photonic crystals, Nature Photonics 9, 320 (2015).
- Zheng and Baranger [2013] H. Zheng and H. U. Baranger, Persistent quantum beats and long-distance entanglement from waveguide-mediated interactions, Phys. Rev. Lett. 110, 113601 (2013).
- Plihal and Maradudin [1991] M. Plihal and A. A. Maradudin, Photonic band structure of two-dimensional systems: The triangular lattice, Phys. Rev. B 44, 8565 (1991).
- Sepkhanov et al. [2007] R. A. Sepkhanov, Y. B. Bazaliy, and C. W. J. Beenakker, Extremal transmission at the dirac point of a photonic band structure, Phys. Rev. A 75, 063813 (2007).
- Zhou et al. [2015] M. Zhou, L. Ying, L. Lu, L. Shi, J. Zi, and Z. Yu, Electromagnetic scattering laws in weyl systems, Nature Communications 9, 326 (2015).
- Kobayashi et al. [1995] T. Kobayashi, Q. Zheng, and T. Sekiguchi, Resonant dipole-dipole interaction in a cavity, Phys. Rev. A 52, 2835 (1995).
- Vaidya et al. [2018] V. D. Vaidya, Y. Guo, R. M. Kroeze, K. E. Ballantine, A. J. Kollár, J. Keeling, and B. L. Lev, Tunable-range, photon-mediated atomic interactions in multimode cavity qed, Phys. Rev. X 8, 011002 (2018).
- Landau and Lifshitz [1960] L. Landau and E. Lifshitz, Electrodynamics of Continuous Media (Pergamon Press, London, 1960).
- Anderson et al. [2006] P. A. Anderson, B. S. Schmidt, and M. Lipson, High confinement in silicon slot waveguides with sharp bends, Optical Express 14, 9197 (2006).
- Yu et al. [2010] Z. Yu, A. Raman, and S. Fan, Fundamental limit of nanophotonic light trapping in solar cells, Proceedings of the National Academy of Sciences of the United States of America 107, 17491 (2010).
- Milonni and Knight [1974] P. W. Milonni and P. L. Knight, Retardation in the resonant interaction of two identical atoms, Phys. Rev. A 10, 1096 (1974).
- Bay et al. [1997] S. Bay, P. Lambropoulos, and K. Mølmer, Atom-atom interaction in strongly modified reservoirs, Physical Review A 55, 1485 (1997).
- Michalski and Mosig [1997] K. A. Michalski and J. R. Mosig, Multilayered media green’s functions in integral equation formulations, IEEE Transactions on Antennas and Propagation 45, 508 (1997).
- Johnson and Joannopoulos [2001] S. G. Johnson and J. D. Joannopoulos, Block-iterative frequency-domain methods for maxwell’s equations in a planewave basis, Optics express 8, 173 (2001).
- Kofman and Kurizki [2000] A. G. Kofman and G. Kurizki, Acceleration of quantum decay processes by frequent observations, Nature 405, 546 (2000).