Strong Coupling and Entanglement of Quantum Emitters Embedded in

Sep 29, 2017 - Establishing strong coupling between spatially separated and thus selectively addressable quantum emitters is a key ingredient to compl...
0 downloads 15 Views 7MB Size
Subscriber access provided by UNIV OF ESSEX

Article

Strong coupling and entanglement of quantum emitters embedded in a nanoantenna enhanced plasmonic cavity Matthias Hensen, Tal Heilpern, Stephen K. Gray, and Walter Pfeiffer ACS Photonics, Just Accepted Manuscript • DOI: 10.1021/acsphotonics.7b00717 • Publication Date (Web): 29 Sep 2017 Downloaded from http://pubs.acs.org on September 29, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Photonics is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

Strong coupling and entanglement of quantum emitters embedded in a nanoantenna enhanced plasmonic cavity Matthias Hensen,*,† Tal Heilpern,‡ Stephen K. Gray,‡ and Walter Pfeiffer§ †

Institut für Physikalische und Theoretische Chemie, Universität Würzburg, Am Hubland, 97074 Würzburg, Germany



Center for Nanoscale Materials, Argonne National Laboratory, 9700 Cass Ave., Lemont, USA §

Fakultät für Physik, Universität Bielefeld, Universitätsstr. 25, 33615 Bielefeld, Germany

Keywords: nanoscale quantum optics, plasmonic cavity, strong coupling limit, entanglement, quantum plasmonics, quantum dynamical Purcell factor

Abstract: Establishing strong coupling between spatially separated and thus selectively addressable quantum emitters is a key ingredient to complex quantum optical schemes in future technologies. Insofar as many plasmonic nanostructures are concerned, however, the energy transfer and mutual interaction strength between distant quantum emitters can fail to provide strong coupling. Here, based on mode hybridization, the longevity and waveguide character of an elliptical plasmon cavity is combined with intense and highly localized field modes of suitably designed nanoantennas. Based on FDTD simulations a quantum emitter–plasmon coupling

ACS Paragon Plus Environment

h

1

ACS Photonics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 29

strength hg = 16.7 meV is reached while simultaneously keeping a small plasmon resonance line width hγs = 33 meV. This facilitates strong coupling and quantum dynamical simulations reveal an oscillatory exchange of excited state population and a notable degree of entanglement between the quantum emitters spatially separated by 1.8 µm, i.e. about twice the operating wavelength.

Main Text: Efficient light-matter interaction in nanophotonic structures and the related quantum emitter (QE) coupling phenomena provide one possible route for downsizing quantum optics based devices to scales compatible with usual microelectronics. Quantum correlations such as entanglement1,2 or squeezing3 persist in plasmonic nanostructures and quantum phenomena both for matter and light play an increasingly important role4,5. Therefore, strong coupling phenomena between a plasmonic mode and a QE ensemble6–9 or, especially, a single QE10–14 have received intense attention in nanophotonics15. For a long time strong coupling between plasmons and a single QE was only subject to theoretical studies, but recently Chikkaraddy et al. have experimentally demonstrated strong coupling of a single dye molecule with a plasmonic mode16. This observation marks a major step towards future applications of plasmonic nanostructures in quantum optical based devices since this demonstrates that QE excitation can be transferred on very short time scales, i.e. on a few 10-14 s, to a plasmonic resonator. Achieving this strong QE-plasmon coupling relies on the strong field enhancement and reduced electromagnetic mode volume in a gap plasmon formed between a gold nanoparticle and a gold film16. The next essential steps are now to employ this strong coupling to a plasmonic resonator to efficiently transfer energy, strongly couple, and establish quantum entanglement between two or more spatially well separated QEs. Coupling multiple QEs and selective addressing (excitation,

ACS Paragon Plus Environment

h

2

Page 3 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

or read out) of individual QEs in an ensemble of many interacting QEs is an essential ingredient for the realization of coherent control schemes needed for example for implementing of quantum computation algorithms. Also in this respect plasmonics provides helpful functionality. Suitably designed plasmonic nanostructures can serve to channel energy launched by one QE to another QE that is spatially sufficiently separated to allow selective excitation, control and read-out. In theoretical investigations plasmonic wedge waveguides17,18 and metal nanoparticle chains19 were investigated. In all these schemes the resulting coupling between QEs remains in the weak coupling limit, i.e. the excitation energy leaves one QE, is transmitted to the other QE, and so efficiently dissipated either by radiative or non-radiative processes that no periodic energy exchange between both emitters is observed. Achieving efficient energy transfer between two separated QEs or even the strong coupling limit in quantum plasmonics is strongly affected by the lossy character of plasmonic resonators. In conventional cavity quantum electrodynamics (cQED) resonators with extremely high quality (Q) factors in the range of 1012 are achievable20 and consequently strong coupling physics is reached even for large mode volumes. In plasmonics, resonances having a large mode volume show indeed rather long lifetimes21 compared to strongly localized modes of nanoantennas. Unfortunately such strongly localized nanoantenna modes, i.e. the intense and short-lived gap plasmons, are required to achieve strong coupling between a single QE and a plasmon16. In other words, it seems that properties like longevity and intense optical modes mutually exclude each other in the field of plasmonics. We propose that hybridization of plasmonic modes22 could solve this dilemma: When two or more systems hybridize to a new entity the properties of the constituents are imprinted on the new eigenmodes. Therefore, one needs to find a combination of different plasmonic structures

ACS Paragon Plus Environment

h

3

ACS Photonics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 29

which, on the one hand, provide opposed features like strongly localized modes and rather high Q-factors, and, on the other hand, can be efficiently coupled. In this context a recently realized hybridized plasmonic device could serve well. Aeschlimann et al. demonstrated strong coupling of optical nanoantennas embedded in an elliptical plasmonic cavity23. The antennas are placed in the two foci of the elliptical cavity and periodic energy exchange between the antennas separated by about 1.6 µm was observed. Note that this large distance allows selective excitation of specific antennas from the far-field. The corresponding hybridized plasmonic modes thus form an interesting platform for plasmon assisted long range QE coupling provided means are identified to strongly couple the QEs to the embedded nanoantennas. Here we make use of this experimentally demonstrated plasmonic coupling scheme23 in a theoretical study and predict strong coupling of two QEs that are spatially separated by more than 1.6 µm, as well as a short-lived meaningful entanglement, by using realistic quantum emitter properties. The manuscript is organized as follows: After a brief presentation of the plasmonic structures and the FDTD based electric field simulations of plasmonic cavity and antenna structures the quantum mechanical coupling phenomena at the boundary of strong coupling are discussed. Finally, based on these coupling constants, strong coupling between widely separated quantum emitters is demonstrated and the dynamics of quantum entanglement is discussed. Discussion and Results We start the discussion with an introduction of the single building blocks of the hybridized plasmonic system. All structural parameters were chosen in such a way that the various building blocks exhibit resonant behavior at νs = 375 THz (wavelength 800 nm, photon energy 1.55 eV). The corresponding electric field distributions of the resonant modes, which are shown in Figure

ACS Paragon Plus Environment

h

4

Page 5 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

1a-c, were retrieved from finite-difference time-domain (FDTD) simulations by placing a point electric dipole source close to the respective mode maximum but 4 nm separated from the nearest metal surface (see Methods section for details). The elliptic SPP cavity24 and its characteristic mode pattern, which is dominated by field components pointing perpendicular to the cavity floor, are presented in Figure 1a. In general, the problem of a vibrating elliptical membrane is solved by Mathieu functions25 and the prevailing mode is determined by the size of the major and minor axis size a and b, respectively. Here, a is set to 2072 nm and b is set to 1450 nm so that injected SPPs which are reflected at the cavity walls form a mode that shows two pronounced field anti-nodes near the geometrical focal spots. The wall height of 800 nm prevents leakage loss since the SPP penetration depth into the vacuum half-space is about 600 nm for Au substrates. Consequently, the SPP mode exhibits a rather high Q-factor of 137.

Figure 1. Elliptical SPP cavity and QE - antenna coupling schemes. (a) Cross section of mode field amplitude distribution for the ‘bare’ elliptical SPP cavity for excitation with a point dipole

ACS Paragon Plus Environment

h

5

ACS Photonics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 29

source located 4 nm above the right focal point. The inset shows the lateral mode pattern in the cavity. (b,c) Mode field amplitude cross sections for the WGM antenna (b) and the gaptenna (c) used to more efficiently couple QE to the elliptical SPP cavity mode. Again point dipole sources were used for excitation, now placed 4 nm above the center mesa of the WGM antenna and in the center of the 8 nm gap of the gaptenna structure. All mode cross sections are shown on the same logarithmic scale. The inset in part (b) shows a 3D representation of the WGM antenna for clarity. The longevity of the cavity mode will be later exploited by the nanoantennas when they are placed in the focal field anti-nodes since all constituents then form a hybridized system23. The nanoantenna simply consists of a circular groove (Figure 1b) and it can be resonantly excited when channel plasmon polaritons26 propagating inside those grooves form a standing wave27. Here, the groove depth dgr was set to 76 nm and the radius was set to 70 nm so that this so-called whispering gallery mode (WGM) resonator exhibits the m = 0 mode which resembles the field distribution of a dipole oscillating perpendicular to the substrate28. This field distribution overlaps well with the spatial cavity mode near the focal spots which is beneficial for energy transfer between antennas at opposite sites. Also, the interaction with attached quantum emitters is expected to be enhanced since far-field radiation loss is suppressed due to the orientation of the modal dipole, and the Q-factor amounts to 24. To further boost light-matter interaction in our structure we utilize the gap-plasmon approach of Chikkaraddy et al.16 and position a 100 nm nanoparticle on top of the WGM antenna so that the point dipole source is located in the center of an 8 nm wide gap between antenna mesa and the nanoparticle surface (Figure 1c). Note that dgr is reduced to 46 nm to compensate for the nanoparticle induced red-shift of the WGM antenna. This structure is labeled gaptenna in the following.

ACS Paragon Plus Environment

h

6

Page 7 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

After having discussed the response of cavity and antenna structures independently we now present results about the combined system consisting of two gaptennas positioned at the focal anti-nodes of the elliptic SPP cavity (Figure 2a). Note, that the structure is excited by a dipole source which is located at the same position as in the case of the pure gaptenna system, i.e. the source is located 4 nm from, both WGM inner mesa and nanoparticle. The obtained field strength is increased by one order of magnitude in the cavity center of the combined system compared to the case of an empty cavity (Figure 1a) and therefore the mode exhibits a delocalized character, as it is expected from a hybridized system.

Figure 2. Strong coupling of gaptennas embedded in an elliptical SPP cavity. (a) Cross section of mode field amplitude distribution for elliptical SPP cavity with gaptennas embedded at the

ACS Paragon Plus Environment

h

7

ACS Photonics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 29

two foci of the cavity mode. The inset shows a close-up of the gap of the right, in which the placement of the point dipole source is indicated by a white arrow. (b) Frequency-dependent or spectral Purcell functions f(ω) for gaptenna and ‘bare’ cavity as solid and dotted line, respectively. (c) f(ω) for the hybridized modes (dotted line) in the combined structure shown in part (a) together with the best fit of a sum of Lorentzian line-shape functions (red solid line) and their individual contributions (solid black lines). The low frequency resonance (grey shaded peak) is used in the following to mediate efficient coupling between QE placed in the gaps of both gaptennas. The resonance width is hγs = 30 meV using γs/2π = 7.5 THz from (c). To discuss the spectral properties of the combined system we first show in Figure 2b the spectral Purcell function f(ω) of the ‘bare’ cavity and the gaptenna as retrieved from a transmission monitor that measures the power radiated by the dipole source (see Methods section for details). According to a Q-factor of 137, f(ω) for the ‘bare’ cavity exhibits a sharp peak while the function of the gaptenna shows a broad peak corresponding to a ten times faster decay time, directly reflected in the different spectral Purcell functions. In the combined system the three modes, i.e. the two antenna modes and the cavity mode, hybridize resulting in the spectral Purcell function f(ω) shown in Figure 2c (dashed line): Three clearly separated peaks are visible and the maximum splitting, defined by the outer modes, amounts to 37 THz (153 meV), i.e. 10% of the operating frequency. Note that the clearly resolved mode splitting shows that the gaptenna modes are strongly coupled via the cavity mode. Interestingly, the coupling strength of the plasmonic structures is increased by more than a factor of 1.5 when gaptennas are used instead of WGM antennas where a splitting of 23 THz (95 meV) was reported23. This increase is attributed to the presence of the nanoparticles which also interact with the SPP mode. The impact of the cavity mode on this hybridized system is made visible by

ACS Paragon Plus Environment

h

8

Page 9 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

fitting the spectral Purcell function with a sum of Lorentzian line shape functions (solid black and red lines in Figure 2c): The line width of the outermost peaks corresponds to a decay time that is only about three times larger than the decay time of the ‘bare’ cavity despite the presence of gaptennas. Still, the gaptennas contribute to the new eigenmodes with their large Purcell enhancement leading to an even higher peak in f(ω) for the red-shifted mode, located at 356 THz (842 nm), compared to the gaptenna case alone (Figure 2b). The hybridization scheme thus indeed combines favorable properties of both modes and enhances coupling between the two gaptennas. Below we will utilize this particular mode to study the plasmon-mediated formation of quantum entanglement between two widely separated QEs since it offers, besides the cavityinduced reduction of line width, the highest Purcell enhancement among the three hybridized modes. For a quantitative assessment of the achieved QE-plasmon and QE-QE coupling in the different schemes considered here and for performing quantum dynamics simulations we now determine the corresponding coupling constants g. According to Fermi’s golden rule, i.e. in the weak coupling limit, the strength of light-matter interaction depends on the photon density of states (PDOS) in which the QE is embedded29,30. Plasmonic resonances modify the local PDOS and thus affect g. In the following this effect is accounted for by the spectral density |g(ω)|2 which is proportional to the PDOS. |g(ω)|2 is expressed using the Green’s function formalism12 yielding 2

g (ω ) =

1

ω2

hπ ε0 c

2

µTq Im G s (ω, rq , rq ) µ q ,

(1)

where the dyadic Green’s function Gs(ω,rq,rq) reflects the local field effects at the position rq of the QE induced by environmental effects, i.e. the presence of a plasmonic resonator. In addition, the dipole moment µq of the actually chosen QE must be known to determine the coupling

ACS Paragon Plus Environment

h

9

ACS Photonics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 29

strength. Here, we assume that the QE exhibits a decay width of hγq = 0.66 µeV (corresponding to a 1 ns lifetime or γq = 1 GHz) at its transition frequency ωq that is chosen in resonance with the plasmon resonance ωs. From γq the dipole moment µq was derived using the relation30

γq =

ωq3 µ q

2

3π ε 0 h c 3

.

(2)

The spectral density |g(ω)|2 according to eq 1 is retrieved from FDTD simulations and the assumed µq. For the various QE-plasmon coupling schemes the Green’s functions are calculated using pre-built scripts31. In the present case Gs and thus also |g(ω)|2 are dominated by distinct plasmon resonances and hence |g(ω)|2 is approximated by a Lorentzian given by 12 2

g (ω ) =

g2 2π

γs

( ω − ωs )

2

+

γ s2

,

(3)

4

with the coupling constant g between QE and the surface plasmon resonance and the plasmon resonance width γs. By fitting |g(ω)|2 using eq 3 the corresponding QE-plasmon coupling constants g, as listed in Table 1, are obtained. To check for consistency a second method to determine g is applied. FDTD simulation provide directly the spectral Purcell function f(ω) (see Methods for more details). Fitting f(ω) with Lorentzians yields the Purcell factor F, i.e. the resonance peak height in f(ω), that is commonly used to characterize the enhanced decay rate of a QE because of the presence of local PDOS effects30 for the plasmon resonances in f(ω) (last column in Table 1). In the weak coupling limit, i.e. if the coupling constant g is much smaller than other decay constants, F can be expressed as32

F=

4 g2

γsγq

,

(4)

ACS Paragon Plus Environment

h

10

Page 11 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

providing a simple relation between F and g. Hence, g can directly be inferred from F using eq 4. Note (see discussion below) that g so-obtained also depends on other system decay constants. While not as accurate as the spectral density approach outlined above, we find that g values obtained with eq 4 differ by less than 10% even in the moderate-strong coupling limit. We should note, however, that neither the spectral density approach nor the weak coupling limit formula properly account for the back-reaction effects, i.e. energy transfer from the plasmon back to the emitter which is at the heart of strong coupling. This is because the underlying classical electrodynamics calculations to yield the spectral density and Purcell factor do not allow such back reaction to occur. Thus one should always exercise some caution with such coupling constant estimates.

Table 1. Quantum emitter coupling parameter g, plasmon resonance width γs, and Purcell factor F for various configurations

hg / meV

hg / meV

eq 3

eq 4

‘Bare’ cavity (A)

0.3

WGM (B)

hγs / meV

F (FDTD)

0.3

11

47.6

2.5

2.5

64

597

Gaptenna (C)

16.1

15.8

135

11.3⋅103

Hybrid - 8 nm gap (D)

9.0

8.6

31

14.3⋅103

Hybrid - 4 nm gap (E)

16.7

15.6

33

45⋅103

Gap plasmon16 (F) *)

NA

90

147

3.5⋅106

The capital letters A to F serve in the following as labels for the different QE coupling schemes. With exception of the gap plasmon case hγq = 0.66 µeV, i.e. a 1 ns lifetime of the QE in vacuum, is assumed. * For the gap plasmon eq 4 is used to determine F based on the measured Rabi splitting for a single molecule that directly yields g, hγs extracted from the upper part in Fig. 3a in ref. 16, and hγq = 0.07 µeV as reported in ref. 16.

ACS Paragon Plus Environment

h

11

ACS Photonics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 29

As expected, the gaptenna features the highest coupling strength of 16 meV for the QEplasmon coupling schemes considered here. It is more than 50 times higher compared to a ‘bare’ cavity and even six times higher than for the WGM antenna. Interestingly it is approaching a value that gets close to thermal energies at room temperature indicating that the design concept can become relevant for room temperature applications. For the hybridized system combining two gaptennas with the elliptical cavity the coupling strength is reduced by about a factor of two compared to the gaptenna case. This reduced coupling can be understood qualitatively: In case of the hybridized system the corresponding plasmonic mode is delocalized between both gaptennas and consequently the mode volume in the gap between the nanoparticle and the WGM mesa increases by a factor of about two since the fields are highly localized in the gaps. This reduction in the coupling constant because of the delocalization between two gaptennas can be compensated by reducing the gap size by a factor of two. With a gap width of 4 nm we get close to the limits of the used FDTD method. Nevertheless it could make sense to consider even higher possible coupling constants g as they have been recently demonstrated experimentally for a single dye molecule embedded in a 1 nm wide gap plasmon mode16. However, for such dimensions simple FDTD calculations are no longer trustworthy and we have therefore kept the gap width to values larger than 4 nm. In the weak coupling limit F is a well-established quantity to quantify the coupling strength. For increasing coupling this simple concept breaks down and the nonlinearity introduced by the back-action of the system on the QE has to be considered. In the following we investigate the transition between the weak and strong coupling regime for QE-plasmon interaction. Based on quantum dynamics calculations (see Methods) with various values of the emitter and plasmon decay constants and QE-plasmon coupling constants discussed above we obtain further insight to

ACS Paragon Plus Environment

h

12

Page 13 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

the weak and strong coupling regimes, as well as to conveniently compare the various systems under study here. The approach we use is to consider a single QE - plasmon system (i.e., include only k = 1 in eqs 5, 7, and 6) and numerically solve for the time evolution of the system. With the same initial condition of only QE1 (now the only QE) initially excited, we obtain the probability P1(t) of QE1 being in the excited state as a function of time, and infer an effective decay constant,

γeff, by fitting P1(t) to an exponential. In the weak coupling limit this probability function will be exponential in time and γeff can be obtained from a fit of the ln P1(t) to a line. For larger values of the coupling parameter P1(t) will exhibit damped oscillations but such a linear fit to ln P1(t) will still yield the underlying decay constant. These damped oscillations reflect the degree of coupling g in each system – in the limit of strong coupling they have frequency 2g or period π/g. We normalize this effective QE decay rate in the presence of the plasmonic system by the isolated QE decay rate to obtain what we refer to as the dynamical Purcell factor: FQD = γeff / γq. Figure 3a shows plots of P1(t) on a logarithmic scale for systems A-F that we have introduced (Table 1). Note that to completely define the quantum system we also specify a value of the QE dephasing parameter, γd. This parameter, just as long as it is small compared to the coupling parameter g and the plasmon resonance decay parameter γs, does not significantly impact the dynamics and we have made the arbitrary choice here that γd = 0.1 g. Unless otherwise stated, we are always assuming this for the QE dephasing. Note also that QE and plasmonic transition frequencies are assumed to be resonant and in this limit the results do not depend on ω0. From Figure 3a we see that systems A-C do indeed exhibit exponential decay. However, systems D-F all exhibit some degree of oscillatory behavior superimposed on an underlying exponential decay indicative of larger QE-plasmon coupling.

ACS Paragon Plus Environment

h

13

ACS Photonics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 29

Actually, the systems A-F can be represented as isolated points in a much wider range of possibilities. It turns out that a universal plot can be made if FQD is plotted as a function of the dimensionless decay constants γs/g and γq/g, as in Figure 3b. Strong coupling manifests itself by leading to the bent, Λ-shaped contours. For the depicted ranges, the strong-coupling limit is the region to the left of the “apex” of the Λ-contours, i.e. the region for log10( γs/g) < 0.5. The deviations of the solid lines from the dashed ones, more noticeable in this left side of Figure 3b also suggest that the use of simple formulae like eq 4 and possibly also the approach of Hümmer et al.12 may not be appropriate, although a detailed analysis of better approaches to obtaining g must be deferred to future work.

Figure 3. (a) Comparison of the QE decay dynamics for different QE-plasmon coupling schemes. Quantum simulations for an initially excited single QE placed at the positions used for

ACS Paragon Plus Environment

h

14

Page 15 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

point dipole emitter excitation as shown in Figure 1b,c for WGM antenna (B, red solid line) and gaptenna (C, black solid line) and in Figure 2a for the hybridized system with 8 nm gap width (blue solid line). In addition, the QE decay for the hybridized system with reduced gap width (E, magenta solid line) and the gap plasmon coupling case reported in ref. 16 (F, green solid line) are shown. The dashed lines for case E and F represent linear fits for determining FQD in the intermediate and strong coupling case. (b) The dynamical Purcell factor FQD, as a function of the plasmon decay constant γs and QE emission rate γq normalized to the QE-plasmon coupling constant g. The QE dephasing rate has been fixed at γd/g = 0.1. The points labeled A-F refer to the QE coupling schemes listed in Table 1. The solid black contour lines correspond to values FQD = 102, 103, …,106. The dashed straight lines correspond to the weak-coupling limit Purcell factor determined using eq 4. In the following we consider the quantum dynamics in a bipartite qubit system coupled via the plasmonic modes introduced above. We use the coupling constants, g, and decay constants inferred for our systems in quantum dynamics calculations (see Methods) to assess how effectively one QE, about 1.6 µm separated from the other QE can nonetheless be coupled to it. The time evolution of two QEs in the hybridized cavity-gaptenna system is presented in Figure 4a-c, where in order to clearly see the added effects of the WGM antenna and the gaptenna, the evolution of the two emitters in the ‘bare’ elliptic SPP cavity are plotted in Figure 4d-f. In all cases the system is initiated with one QE excited and with both the other QE and the plasmonic system in their ground states. The time evolution is computed from eqs 5-7 (see Methods), where now k = 1,2 in eq 7, using the coupling and decay parameters taken from Table 1. Note that again the QE dephasing γd is chosen as 0.1 g resulting in hγd ≈ 1 meV and 0.03 meV for gaptenna and bare cavity, respectively. Whereas the value for the bare cavity might only be reached at low

ACS Paragon Plus Environment

h

15

ACS Photonics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 29

temperature a linewidth in the range of 1 meV is a reasonable estimate for room temperature 33,34 In both cases the initially excited QE decay and some energy is transferred to the other QE. Whereas the dynamics in the ‘bare’ cavity occur roughly on a 30 ps time scale (Figure 4d) this process occurs much faster. In the ‘bare’ cavity case the dynamics are dominated by an almost exponential decay of the QE1 excitation and very little transfer to QE2.

In contrast, the

population dynamics in the hybrid system are much more complicated: Within about 100 fs almost 30 % of the QE1 excitation is transferred to QE2 and after a small overshooting both excitations decay on a ps time scale. As discussed below the small overshooting is a first indication of the onset of strong coupling effects between both QEs.

Figure 4. Time evolution of two QE embedded in the hybridized cavity-gaptenna system, (a)-(c), and the elliptic SPP cavity, (d)-(f). The hybridized system is a result of all components of the system involved, i.e., the cavity, the WGM antenna and the gaptenna of dgap = 8 nm, while on the right the dynamics of the QEs in the ‘bare’ cavity are plotted. The insets in (a) and (d) depict the corresponding coupling schemes. (a) and (d) show the time evolution of the populations of both QE, and (b) and (e) show the evolution of the populations of the symmetric and anti-symmetric states for the two configurations. (c) and (f) show the evolution of the concurrence, which is a measure of the entanglement between the two QEs.

ACS Paragon Plus Environment

h

16

Page 17 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

Figures 4b,e show the quantum dynamics of both configurations projected on the symmetric and antisymmetric basis vectors that serve to better convey the underlying physics of this behavior18,19. In the hybridized system the symmetric state exhibits a fast almost complete initial decay whereas the anti-symmetric state decays with a much longer time constant18. In the latter case the plasmon related dissipation cancels at least partially and thus the decay rate for this state is smaller. This behavior is analog to the radiative damping of symmetrically and antisymmetrically coupled oscillating dipole emitters. In this case the anti-symmetric mode, often denoted as dark mode, decays less efficiently since it does not couple to transverse modes of the electromagnetic field. This behavior is closely related to the dynamics of the bipartite entanglement in the present situation. Here we employ the concurrence C as measure of entanglement between both QEs35. C starts from an unentangled initial state and increases with the rapid decay of the symmetric state to about 0.4 in the case of the hybrid system. Afterwards the entanglement decays with the same time constant as the anti-symmetric state. That is, while the initial state is a separable product of the QE states, it can be decomposed as an equal amplitude superposition of completely entangled symmetric and anti-symmetric states. The more rapid decay of the symmetric state leads to a state with more (slowly decaying) entangled anti-symmetric character and thus greater concurrence. However, the dynamics is relatively fast, with the maximum in concurrence occurring after about 100 fs (Figure 4c). In comparison to the hybridized system dynamics, the cavity case (Figures 4d-f) displays longer decays and a lower concurrence maximum. We once again would like to stress that the currently discussed onset of strong coupling was reached with a rather large gap size of 8 nm. Although the emitter properties of quantum systems are expected to be quenched close to metallic surfaces36, Baumberg and co-workers found

ACS Paragon Plus Environment

h

17

ACS Photonics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 29

extraordinary high Purcell factors16 (see Table 1) for a gap size of about 1 nm. Therefore we finally demonstrate that our proposed cavity design explicitly allows strong coupling of widely separated QEs by reducing the gap size from 8 nm to 4 nm.

Figure 5. Onset of strong long-range QE coupling in hybridized cavity with 4 nm gaptenna gap width. (a) Spectral Purcell function f(ω) for the hybridized modes in the elliptical cavity with embedded gaptennas (dotted line) together with the best fit of a sum of Lorentzian line-shape functions (red solid line) and their individual contributions (solid black lines). The low frequency resonance (grey shaded peak) is used in the following to mediate efficient coupling between QE placed in the gaps of both gaptennas. Here the QE dephasing parameter hγs = 1.7 meV. (b) Population and concurrence dynamics for the hybrid system with reduced gap width (E). The arrangement of panels (b-d) is the same as in Figure 4a-c.

ACS Paragon Plus Environment

h

18

Page 19 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

Without loss of generality, and to show the flexibility of our system, we adhere to the identical WGM antenna geometry for this scenario. Because of the reduced gap size the antenna resonance shift to slightly lower frequencies. To account for this resonance shift the cavity’s major axis is increased by 222 nm moving the cavity mode again into resonance with the gaptenna mode. The spectral resonance of both systems is then centered at λ0 = 860 nm (photon energy 1.44 eV or ν0 = 348.6 THz). The spectral Purcell function f(ω), retrieved from FDTD simulations in the weak coupling limit, is shown in Figure 5a: Again, three clearly separated eigenmodes are observed, spanning a frequency range of 37 THz, i.e. more than 10% of the center frequency. By applying the above mentioned fitting procedure (red solid line in Figure 5a) we retrieve a Purcell factor of 44,900 for the red-shifted eigenmode at 331.6 THz. Hence, a reduction of gap size from 8 nm to 4 nm leads to a 3.5 times higher Purcell enhancement, and the coupling constant increases to hg = 16.7 meV while the plasmon resonance line-width remains rather constant at hγs = 33.1 meV. Maintaining plasmon dissipation properties while increasing the coupling strength of the QE-plasmon system facilitates to finally pass the threshold towards strong coupling: Figure 5b shows the population dynamics of two QEs having the same properties as before (γq = 1 GHz at

ν0 = 348.6 THz) but positioned now in narrower gaps. The QEs undergo a partial oscillation in their respective excited state populations, however, dissipation and decoherence still prevents additional oscillations. In contrast to previous works in which plasmon-mediated quantum entanglement was realized via cooperative decay effects18,19, i.e. subradiance, we clearly see modulations in the state populations (Figures 5b,c) and also in the concurrence and (Figure 5d). While a concurrence of unity is the optimal value which indicates a maximally entangled state, the transient values achieved over a short period (< 0. 25 ps) in the hybridized cavity-gaptenna

ACS Paragon Plus Environment

h

19

ACS Photonics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 29

system are similar to those reported for other systems18,19. However, analogous calculations (not shown) for parameters corresponding to the QE-plasmon coupling conditions reported by Chikkaraddy et al. for single dye molecules in a gap plasmon16 yields slightly larger transient values of C, but they also do not exceed C = 0.5. To conclude, we studied a hierarchy of coupling schemes, ranging from a micrometer-sized plasmonic cavity, over small mode volume nanostructures featuring whispering gallery modes and gap plasmons, to an optimal hybrid system that combines the plasmonic properties of the aforementioned structures. The Purcell factors and spectral density associated with quantum emitters in proximity to these systems were calculated with the finite-difference time-domain method, and analyzed to obtain reliable emitter-plasmon coupling constants. A universal plot was introduced that allowed us to place all these systems in perspective: Our hybridized system approach was consistent with a moderate to strong coupling limit although the related plasmon mode is delocalized over a distance of more than 1.8 µm. Consequently, by employing associated coupling constants and decay parameters in quantum dynamics calculations, we demonstrated that two quantum emitters, separated by this large distance, are strongly coupled via the delocalized plasmon mode and that a meaningful degree of entanglement is establishes with about 100 fs. To combine the best properties of two worlds by hybridization, i.e. to imprint the longevity of cavity modes on the intense light-matter interaction of strongly localized plasmon resonances, constitutes a relevant approach for implementing nanoscale quantum information and sensing applications in the near future.

Methods

ACS Paragon Plus Environment

h

20

Page 21 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

FDTD simulations. Maxwell’s equations describing electrodynamic phenomena in different plasmonic environments are numerically solved by using a commercial software package (FDTD Solutions 8.11.337, Lumerical Solutions Inc.). The designed plasmonic structures consist only of gold due to its inert character and we used the data of Johnson and Christy37 to model gold’s dielectric function. All nanoscale structural features which give rise to intense electric near-fields are positioned at a distance of at least 900 nm (> λ0) from the simulation volume boundaries. We employed absorbing boundary conditions using a minimum number of 32 perfectly matched layers at each boundary in order to prevent reflections of propagating fields. The adaptive mesh accuracy of the simulation volume was set to 22 sampling points per effective wavelength λ inside the corresponding material and we omitted a conformal mesh refinement. Instead, we encased the WGM antenna, particle, and the gap in between with additional meshes providing a resolution of 2 nm in each direction, except the axial gap direction which was resolved by 1 nm. In the case of the system featuring the 4 nm gap size the structures are resolved by 1 nm and 0.5 nm, respectively. Note that these monitors are active in all simulations, irrespective of the presence of antennas or nanoparticles, in order to guarantee similar mesh conditions. The excitation source is a pulsed dipole (4.3 fs and λ0 = 800 nm) which exhibits a minimum distance of four mesh cells to interfaces in all simulations. Purcell factors were calculated by encasing the dipole source with a built-in transmission monitor: 2 nm x 2 nm x 1 nm in case of a 4 nm gap and 2 nm x 2 nm x 1 nm in the case of 8 nm gap size. The monitor integrates the Poynting vector over the monitor surface and normalizes the result to the source power of the dipole in a homogeneous environment so that the output directly gives the Purcell enhancement in the weak coupling limit. Note that in all cases the simulation time was chosen to guarantee a decay of the electric field intensity by at least three orders of magnitude.

ACS Paragon Plus Environment

h

21

ACS Photonics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 29

In the case of the hybridized and pure cavity system we employed a symmetric boundary condition along the major axis of the ellipse in order to reduce computational effort. For the different antenna systems we employed symmetric boundary conditions along both lateral directions.

Quantum dynamics calculations. Following earlier treatments of quantum emitter/plasmonic system coupling 17,19, one can imagine a system of two, two-state quantum emitter (QEs) coupled to one another indirectly via mutual interaction with a plasmonic system. The relevant system states are denoted as q1 , q2 , s , where qi = 0 or 1 denotes QE i being in its ground or excited state, and s = 0, 1, 2, …, smax denotes the possible (bosonic or harmonic oscillator) states associated with a single plasmon mode. A density operator formalism is used with



ρ (t ) =

ρ q ,q ,s ,q′ ,q′ ,s′ q1 , q2 , s q1′, q2′ , s ′ , 1

q1 , q2 , s , q1′ , q2′ , s ′

where ρ q ,q 1

2 , s , q1′ , q2′ , s ′

2

1

(5)

2

are complex-valued density matrix elements. The density operator satisfies the

equation of motion

∂ i ρ = − [H , ρ ] + L (ρ ) , ∂t h

(6)

where H is the Hamiltonian and L the Lindblad superoperator

38

that describes QE spontaneous

emission/dephasing and plasmon decay/dephasing. For our purposes we do not include any external fields and express the Hamiltonian simply as that of the QE and plasmon, H = h ω s b † b + h ωq

∑σ

σ k− + h g

+ k

k =1,2

where

b+

and

b

are

∑ (σ

k =1,2

the

+ k

b + σ k− b† ) ,

harmonic

oscillator

(7) raising

and

lowering

operators,

σ k− = 0k 1k , σ k+ = 1k 0k , and we will take the plasmon and QE transition frequencies to be equal, ωs = ωd = ω0 . Each QE is assumed to be coupled to the plasmon mode with the same

ACS Paragon Plus Environment

h

22

Page 23 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

dipolar coupling constant g. See Ref.

32

for the explicit form of L, and we note that it is

parameterized by the QE’s spontaneous emission rate γq, the environmentally-induced dephasing rate γd, and the plasmon decay constant γs. Insertion of eq 5 into eq 6 with eq 7 leads to a set of linear differential equations for the

16 ( smax + 1) density matrix elements ρ q ,q 2

1

2 , s , q1′ , q2′ , s ′

that can be solved numerically to yield, via

appropriate traces of the density matrix, populations and other quantities of interest as a function of time. We focus on an initial condition corresponding to the QE 1 being in its excited state with the rest of the system cold, and for such calculations it suffices to include just two plasmon states (s = 0 and 1) and simply solve the density matrix equations numerically with a standard ordinary differential equations numerical solver. In addition to the excited state populations of the two QE, we calculate the populations in the symmetric and antisymmetric entangled states,

S =1

2 ( 10 + 01

)

and A = 1

2 ( 10 − 01 ) , where q1 q2 = q1 q2

denotes a product

of the two QE states. We also compute the bipartite concurrence, C(t), which is a measure of the degree of entanglement between the two QE 35 and thus the potential relevance of the systems to quantum information and sensing applications. The concurrence is between 0 and 1, with 0 indicating a separable (unentangled) state. While any C > 0 indicates the state is an entangled state, i.e. a state that cannot be written as a separable state. Quantum information applications generally use “maximally entangled” states with C = 1.

Corresponding Author *E-mail: [email protected]

Author Contributions The manuscript was written through contributions of all authors.

ACS Paragon Plus Environment

h

23

ACS Photonics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 29

ACKNOWLEDGMENT The research leading to these results has received funding from the German Research Foundation (DFG) within the priority program SPP 1391 “Ultrafast Nanooptics” under Grant Agreement PF 317/5. This work was performed, in part, at the Center for Nanoscale Materials, a U.S. Department of Energy Office of Science User Facility under Contract No. DE-AC0206CH11357. ABBREVIATIONS cQED, cavity QED; FDTD, finite difference time domain; QE, quantum emitter. REFERENCES (1)

Altewischer, E.; van Exter, M. P.; Woerdman, J. P. Plasmon-Assisted Transmission of

Entangled Photons. Nature 2002, 418 (6895), 304–306. (2)

Berthel, M.; Huant, S.; Drezet, A. Spatio-Temporal Second-Order Quantum Correlations

of Surface Plasmon Polaritons. Opt. Lett. 2016, 41, 37-40. (3)

Huck, A.; Smolka, S.; Lodahl, P.; Sørensen, A.; Boltasseva, A.; Janousek, J.; Andersen,

U. Demonstration of Quadrature-Squeezed Surface Plasmons in a Gold Waveguide. Phys. Rev. Lett. 2009, 102, 246802. (4)

Tame, M. S.; McEnery, K. R.; Özdemir, Ş. K.; Lee, J.; Maier, S. A.; Kim, M. S. Quantum

Plasmonics. Nat. Phys. 2013, 9, 329–340. (5)

Fitzgerald, J. M.; Narang, P.; Craster, R. V.; Maier, S. A.; Giannini, V. Quantum

Plasmonics. Proc. IEEE 2016, 104, 2307–2322.

ACS Paragon Plus Environment

h

24

Page 25 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

(6)

Vasa, P.; Wang, W.; Pomraenke, R.; Lammers, M.; Maiuri, M.; Manzoni, C.; Cerullo, G.;

Lienau, C. Real-Time Observation of Ultrafast Rabi Oscillations between Excitons and Plasmons in Metal Nanostructures with J-Aggregates. Nat. Photon. 2013, 7, 128–132. (7)

Delga, A.; Feist, J.; Bravo-Abad, J.; García-Vidal, F. J. Quantum Emitters Near a Metal

Nanoparticle: Strong Coupling and Quenching. Phys. Rev. Lett. 2014, 112, 253601. (8)

Delga, A.; Feist, J.; Bravo-Abad, J.; García-Vidal, F. J. Theory of Strong Coupling

between Quantum Emitters and Localized Surface Plasmons. J. Opt. 2014, 16, 114018. (9)

Zhou, N.; Yuan, M.; Gao, Y.; Li, D.; Yang, D. Silver Nanoshell Plasmonically

Controlled Emission of Semiconductor Quantum Dots in the Strong Coupling Regime. ACS Nano 2016, 10, 4154–4163. (10) Trügler, A.; Hohenester, U. Strong Coupling between a Metallic Nanoparticle and a Single Molecule. Phys. Rev. B 2008, 77, 115403. (11) Barthes, J.; Bouhelier, A.; Dereux, A.; Francs, G. C. des. Coupling of a Dipolar Emitter into One-Dimensional Surface Plasmon. Sci. Rep. 2013, 3. (12) Hümmer, T.; García-Vidal, F. J.; Martín-Moreno, L.; Zueco, D. Weak and Strong Coupling Regimes in Plasmonic QED. Phys. Rev. B 2013, 87, 115419. (13) Hakami, J.; Wang, L.; Zubairy, M. S. Spectral Properties of a Strongly Coupled Quantum-Dot - Metal-Nanoparticle System. Phys. Rev. A 2014, 89, 053835. (14) Varguet, H.; Rousseaux, B.; Dzsotjan, D.; Jauslin, H. R.; Guerin, S.; Colas des Francs, G. Dressed States of a Quantum Emitter Strongly Coupled to a Metal Nanoparticle. Opt. Lett. 2016, 41, 4480.

ACS Paragon Plus Environment

h

25

ACS Photonics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 29

(15) Törmä, P.; Barnes, W. L. Strong Coupling between Surface Plasmon Polaritons and Emitters: A Review. Rep. Prog. Phys. 2015, 78, 013901. (16) Chikkaraddy, R.; de Nijs, B.; Benz, F.; Barrow, S. J.; Scherman, O. A.; Rosta, E.; Demetriadou, A.; Fox, P.; Hess, O.; Baumberg, J. J. Single-Molecule Strong Coupling at Room Temperature in Plasmonic Nanocavities. Nature 2016, 535, 127–130. (17) González-Tudela, A.; Martín-Cano, D.; Moreno, E.; Martín-Moreno, L.; Tejedor, C.; García-Vidal, F. J. Entanglement of Two Qubits Mediated by One-Dimensional Plasmonic Waveguides. Phys. Rev. Lett. 2011, 106, 020501. (18) Martín-Cano, D.; González-Tudela, A.; Martín-Moreno, L.; García-Vidal, F. J.; Tejedor, C.; Moreno, E. Dissipation-Driven Generation of Two-Qubit Entanglement Mediated by Plasmonic Waveguides. Phys. Rev. B 2011, 84, 235306. (19) Otten, M.; Shah, R. A.; Scherer, N. F.; Min, M.; Pelton, M.; Gray, S. K. Entanglement of Two, Three, or Four Plasmonically Coupled Quantum Dots. Phys. Rev. B. 2015, 92, 125432. (20) Gleyzes, S.; Kuhr, S.; Guerlin, C.; Bernu, J.; Deléglise, S.; Busk Hoff, U.; Brune, M.; Raimond, J.-M.; Haroche, S. Quantum Jumps of Light Recording the Birth and Death of a Photon in a Cavity. Nature 2007, 446, 297–300. (21) Zhu, X.; Zhang, J.; Xu, J.; Li, H.; Wu, X.; Liao, Z.; Zhao, Q.; Yu, D. Dispersion Control in Plasmonic Open Nanocavities. ACS Nano 2011, 5, 6546–6552. (22) Prodan, E.; Radloff, C.; Halas, N. J.; Nordlander, P. A Hybridization Model for the Plasmon Response of Complex Nanostructures. Science 2003, 302, 419–422.

ACS Paragon Plus Environment

h

26

Page 27 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

(23) Aeschlimann, M.; Brixner, T.; Cinchetti, M.; Frisch, B.; Hecht, B.; Huber, B.; Hensen, M.; Kramer, C.; Krauss, E.; Loeber, T. E.; Pfeiffer, W.; Piecuch, M.; Thielen, P. Cavity-Assisted Ultrafast Long-Range Periodic Energy Transfer between Plasmonic Nanoantennas. Light Sci. Appl. 2017, accepted for publication. (24) Schoen, D. T.; Coenen, T.; García de Abajo, F. J.; Brongersma, M. L.; Polman, A. The Planar Parabolic Optical Antenna. Nano Lett. 2013, 13, 188–193. (25) Gutiérrez-Vega, J. C.; Rodrı́guez-Dagnino, R. M.; Meneses-Nava, M. A.; Chávez-Cerda, S. Mathieu Functions, a Visual Approach. Am. J. Phys. 2003, 71, 233–242. (26) Moreno, E.; García-Vidal, F. J.; Rodrigo, S. G.; Martín-Moreno, L.; Bozhevolnyi, S. I. Channel Plasmon-Polaritons: Modal Shape, Dispersion, and Losses. Opt. Lett. 2006, 31, 3447. (27) Vesseur, E. J. R.; García de Abajo, F. J.; Polman, A. Modal Decomposition of SurfacePlasmon Whispering Gallery Resonators. Nano Lett. 2009, 9, 3147–3150. (28) Vesseur, E. J. R.; Polman, A. Plasmonic Whispering Gallery Cavities As Optical Nanoantennas. Nano Lett. 2011, 11, 5524–5530. (29) Fox, M. Quantum Optics: An Introduction; Oxford master series in physics; Oxford University Press: Oxford; New York, 2006. (30) Novotny, L.; Hecht, B. Principles of Nano-Optics; Cambridge University Press: Cambridge; Camebridge, 2006. (31) Green’s

Function

and

LDOS

|

Lumerical

Knowledge

Base

https://kb.lumerical.com/en/nanophotonic_applications_greens_function_and_ldos.html (accessed Jul 1, 2017).

ACS Paragon Plus Environment

h

27

ACS Photonics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 29

(32) Shah, R. A.; Scherer, N. F.; Pelton, M.; Gray, S. K. Ultrafast Reversal of a Fano Resonance in a Plasmon-Exciton System. Phys. Rev. B 2013, 88, 075411. (33) Bayer, M.; Forchel, A. Temperature Dependence of the Exciton Homogeneous Linewidth in In 0.60 Ga 0.40 As / GaAs Self-Assembled Quantum Dots. Phys. Rev. B 2002, 65, 041308. (34) Richter, M.; Gegg, M.; Theuerholz, T. S.; Knorr, A. Numerically Exact Solution of the Many Emitter–Cavity Laser Problem: Application to the Fully Quantized Spaser Emission. Phys. Rev. B 2015, 91, 035306. (35) Wootters, W. K. Entanglement of Formation of an Arbitrary State of Two Qubits. Phys. Rev. Lett. 1998, 80, 2245–2248. (36) Anger, P.; Bharadwaj, P.; Novotny, L. Enhancement and Quenching of Single-Molecule Fluorescence. Phys. Rev. Lett. 2006, 96, 113002. (37) Johnson, P. B.; Christy, R. W. Optical Constants of the Noble Metals. Phys. Rev. B 1972, 6, 4370–4379. (38) Blum, K. Density Matrix Theory and Applications, 3rd ed.; Springer series on atomic, optical and plasma physics; Springer: Heidelberg; New York, 2012.

ACS Paragon Plus Environment

h

28

Page 29 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

For Table of Contents Use Only

Strong coupling and entanglement of quantum emitters embedded in a nanoantenna enhanced plasmonic cavity Matthias Hensen, Tal Heilpern, Stephen K. Gray, and Walter Pfeiffer

A hybridized plasmonic structure (left) is proposed that combines strong electric fields and long-lived modes. It therefore allows for strong coupling of quantum emitters (QE) which are separated by 1.8 µm. Through this coupling, a selectively excited emitter can exchange energy with the distant emitter on an ultrafast time scale (right) and a significant degree of entanglement is obtained.

ACS Paragon Plus Environment

h

29