Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Monolithically integrated single quantum dots coupled to bowtie nanoantennas

Open Access Open Access

Abstract

Deterministically integrating semiconductor quantum emitters with plasmonic nano-devices paves the way towards chip-scale integrable, true nanoscale quantum photonic technologies. For this purpose, stable and bright semiconductor emitters are needed, which moreover allow for CMOS-compatibility and optical activity in the telecommunication band. Here, we demonstrate strongly enhanced light-matter coupling of single near-surface (< 10 nm) InAs quantum dots monolithically integrated into electromagnetic hot-spots of sub-wavelength sized metal nanoantennas. The antenna strongly enhances the emission intensity of single quantum dots by up to ~ 16×, an effect accompanied by an up to 3.4× Purcell-enhanced spontaneous emission rate. Moreover, the emission is strongly polarised along the antenna axis with degrees of linear polarisation up to ~ 85 %. The results unambiguously demonstrate a pronounced coupling of individual quantum dots to state-of-the-art nanoantennas. Our work provides new perspectives for the realisation of quantum plasmonic sensors, step-changing photovoltaic devices, bright and ultrafast quantum light sources and efficient nano-lasers.

© 2016 Optical Society of America

1. Introduction

The field of nanoplasmonics [1] has already demonstrated outstanding potential to tailor and enhance electromagnetic fields on sub-wavelength lengthscales [2,3]. As such it represents the most promising route to interface state-of-the-art electronics with true nano-photonic devices on the same chip [4, 5]. To this end, the study, optimisation and integration of nano-scale plasmonic components, such as antennas [6] and waveguides [5], on high-quality semiconductor substrates [7–9] is essential in order to prove their applicability in telecommunication [10] and quantum [11] applications. Monolithically integrated, self-assembled quantum dots [12] exhibit outstanding electrical and optical properties, they do not suffer from bleaching or blinking and have near-unity internal quantum efficiencies. These properties stem from the efficient decoupling from environmental perturbations in the solid-state matrix material and distinguish self-assembled quantum dots from alternative quantum emitters, such as nitrogen vacancy centres [13], colloidal nano-crystals [14,15], single molecules [16] or fluorescent dyes [17]. Lithographically defined plasmonic dimer antennas, such as bowties [6], are most prominent amongst the zoo of metallic nanoparticles since they simultaneously provide strong light confinement in sub-wavelength sized hot-spots, large-range spectral tunability and facilitate electrical access [18] and full control of the emission polarisation [19]. As a result, the quantum dot coupled nanoantennas offer new perspectives to probe light-matter-couplings and cavity quantum electrodynamics (cQED) effects beyond the point-dipole approximation [20].

Here, we coupled individual quantum dots to plasmonic nanoantennas to form a promising cQED-system schematically illustrated in Fig. 1(a), which consists of near-surface (d ~ 10 nm) InAs/AlGaAs quantum dots [21] and lithographically defined triangular bowtie nanoantennas [6] arranged in a square array (left scanning electron microscopy image in Fig. 1(b)) with a lattice constant a = 1.5 µm. A typical nanoantenna with triangle size s = 87 ± 3 nm and feedgap size g = 26 ± 3 nm is shown in the right scanning electron microscopy image in Fig. 1(b). The self-assembled InAs quantum dots are grown by solid-source molecular beam epitaxy and embedded into high-band gap AlGaAs (see Appendix) in order to suppress carrier tunnelling from the quantum dots to trap states at the nearby sample surface. In the top panel of Fig. 1(c), we present simulations for a bowtie nanoantenna (nominal s0 = 90 nm and g0 = 15 nm) of the scattering cross-section σscatt (blue curve) and the normalised decay rate Γ/Γ0 (red curve), where Γ and Γ0 denote the decay rate of a point-dipole with and without the proximal bowtie nanoantenna, respectively. We observe a Δ ~ 30 meV red-shift of the maximum Γ/Γ0 with respect to the maximum σscatt, which is attributed to the difference between the near-and the far-field coupling [22]. In the bottom panel of Fig. 1(c), we show the experimentally determined differential reflectivity spectrum (raw data in grey, 50-point-smoothed average in blue) of the according nanoantenna that reveals the surface plasmon resonance and find a good agreement with the simulated σscatt [23]. Moreover, the red and green curves show typical photoluminescence spectra of the quantum dot ensemble and the GaAs substrate (including carbon impurities [24]) recorded spatially displaced from the nanoantennas. A good spectral overlap clearly exists between the quantum dot emission and the near-field plasmonic mode.

 figure: Fig. 1

Fig. 1 Sample layout, structural and optical characterisation: (a) Schematic illustration of the monolithically integrated quantum dot-nanoantenna system. (b) Scanning electron microscopy image of the bowtie nanoantenna array (left) and an individual bowtie nanoantenna (right) with s = 87 nm and g = 26 nm. Scale bar, 20 nm. (c) Top: Simulated scattering cross-section σscatt and normalised decay rate Γ/Γ0 of a bowtie nanoantenna with nominal s0 = 90 nm and g0 = 15 nm in blue and red, respectively. Bottom: Measured quantum dot spectrum (L: laser, C: carbon impurities) and differential reflectivity of a bowtie nanoantenna with s = 87 nm and g = 26 nm in red/green and grey, respectively. Blue curve shows a 50-point-smoothed average.

Download Full Size | PDF

2. Identification of coupled quantum dots

We identify quantum dots that are spatially coupled to the bowtie nanoantenna using spatially resolved photoluminescence spectroscopy. Hereby, we simultaneously record the quantum dot emission and the partly reflected excitation laser that is co-polarised along the main antenna axis (i.e. y-axis in Fig. 1(b)). The latter enables us to determine the spatial position of the bowtie nanoantennas with an accuracy of 250 nm and, thus, to correlate it with the quantum dot emission. In Fig. 2(a), we present experimental data of the laser reflectivity IR, the quantum dot emission in co-polarised Ico and cross-polarised Icross detection geometry and the calculated differential signal ΔI = IcoIcross from left to right, respectively. The ΔI-map exhibits four spatially distinct positions (highlighted by arrows) with clearly enhanced responses which are spatially well correlated to the bowtie nanoantenna positions extracted via the IR-map (highlighted by the grid). Spectrally resolved photoluminescence recorded from three uncoupled (i.e. reference) quantum dots located outside the bowtie nanoantenna array and the same four coupled quantum dot-nanoantenna systems investigated in panel (a) are shown for co- (red) and cross- (blue) polarised detection in Figa. 2(b) and 2(c), respectively. For the quantum dots coupled to antennas we observe a maximum enhancement ηmax of the photoluminescence intensity of ~ 16×, when compared to uncoupled quantum dots in a co-polarised detection geometry. The average enhancement η˜ of all seven coupled quantum dots measured is 8.7×, which is in good agreement with other reports on similar systems [25]. We attribute this enhancement to a pronounced coupling of the emitter to the plasmonic mode, resulting in a combination of excitation enhancement [26], Purcell-enhanced emission [17] and spatial redistribution [27]. We note that quantum dots with enhanced photoluminescence intensities are not necessarily be positioned perfectly below the feedgap of the antenna since no active positioning [25] was used in our work. Therefore, the number for the maximum enhancement above can only be seen as a lower bound.

 figure: Fig. 2

Fig. 2 Enhanced luminescence from coupled quantum dot-nanoantenna systems: (a) Spatially resolved laser reflectivity, co-, cross-polarised and differential quantum dot emission from left to right, respectively. (b), (c) Photoluminescence spectra of uncoupled and coupled quantum dots for co- and cross-polarised emission in blue and red, respectively.

Download Full Size | PDF

3. Strongly polarised quantum dot luminescence

In order to explore the coupling mechanisms between the quantum dots and the bowtie nanoantennas, we present in Figs. 3(a) and 3(b) polarisation-resolved photoluminescence studies for a coupled quantum dot-nanoantenna system (top) and an uncoupled quantum dot (bottom) in a false-colour and polar-plot representation, respectively. We observe strongly polarised emission from the coupled quantum dot-nanoantenna systems along the main antenna axis, yielding degrees of linear polarisation ρ = (Imax−Imin)/(Imax + Imin) of up to ρcoupled = 85%. In strong contrast, for the reference quantum dots we observe predominantly unpolarised emission with ρref = 9 ± 3% as shown in Fig. 3(c). The results presented show that the degree of polarisation and the according polarisation angle of self-assembled InAs quantum dots are determined by the coupling to the highly polarised dipolar fields of the bowtie nanoantennas [7].

 figure: Fig. 3

Fig. 3 Strong linearly polarised emission from coupled quantum dot-nanoantenna systems: (a) Polarisation-resolved photoluminescence for a coupled quantum dot-nanoantenna system (top) and an uncoupled quantum dot (bottom). (b) Corresponding peak intensities of a coupled quantum dot-nanoantenna system (top) and an uncouped quantum dot (bottom) in a polar-plot representation. (c) Degree of linear polarisation ρ for coupled quantum dot-nanoantenna systems (red) and uncoupled quantum dots (blue).

Download Full Size | PDF

4. Purcell-enhanced quantum dot luminescence

We continue to study the contribution of the Purcell-enhancement to the overall intensity increase of coupled quantum dot-nanoantenna systems by applying time-resolved photoluminescence spectroscopy and compare our results to uncoupled quantum dots. In Fig. 4(a), we present the time-resolved transients of a coupled quantum dot-nanoantenna system (red) and a reference quantum dot (blue) in a semi-logarithmic representation. We observe a significantly reduced spontaneous emission lifetime for the coupled quantum dot-nanoantenna system of τc = 0.14 ± 0.1 ns (time-resolution limited), as compared to the uncoupled dot (τref = 0.44 ± 0.1 ns). In general, we find average lifetimes of coupled and uncoupled quantum dots of τ˜c=0.21±0.1 ns and τ˜ref=0.48±0.1 ns, respectively, as summarised in Fig. 4(b), which we attribute to varying spatial displacements between the quantum dot positions and the antenna hot-spots [25]. The clear decrease in spontaneous emission lifetime accompanied with the pronounced increase of emission intensity unambiguously demonstrates a modification of the radiative emission dynamics via the Purcell-effect [28], giving rise to Purcell-factors FP > 3.4. Moreover, we observe a clear correlation between the measured spontaneous emission lifetimes and the corresponding linear degrees of polarisation as shown in Fig. 4(c), indicating that pronounced modifications of the spontaneous emission lifetimes are accompanied by highly linear polarised emission.

 figure: Fig. 4

Fig. 4 Purcell-enhanced emission from coupled quantum dot-nanoantenna systems: (a) Time-resolved photoluminescence transients of a coupled quantum dot-nanoantenna system (red), an uncoupled quantum dot (blue) and the instrument response function (black). (b) Histogram of the spontaneous emission lifetimes for coupled quantum dot-nanoantenna systems (red) and uncoupled quantum dots (blue). (c) Correlation between the spontaneous emission lifetime and the degree of linear polarisation.

Download Full Size | PDF

5. Conclusion

In summary, we demonstrated pronounced light-matter-coupling of bright (~16× enhanced) and fast (< 200 ps), monolithically integrated, semiconductor InAs quantum dots positioned 10 nm below small feedgap (25 nm) lithographically defined Au bowtie nanoantennas, resulting in Purcell-factors FP > 3.4. The combination of such exquisite quantum emitters with lithographically engineered metallic nanoantennas [6,26] enables intrinsic electrical connectivity [18] and, thus, expands the currently available portfolio of experimental techniques towards photocurrent spectroscopy and Stark tuning [29]. Moreover, the studied hybrid system facilitates the extension to so-called cross-antennas [19], enabling full control of emission polarisation and, thus, represents a novel platform for advanced resonant fluorescence spectroscopy [30]. Further potential for optimising the light-matter-interaction is envisioned by deterministic quantum dot-antenna positioning [25], by engineering the capping of the semiconductor-heterostructure [31] and the in-situ epitaxial growth of monocrystalline metals [32] that most likely also leads to a surface passivation of the sample. Moreover, the presented InAs quantum dot-nanoantenna system is a natural candidate to study in detail light-matter-interactions in mesoscopic quantum systems [20], since the confined electromagnetic field is on the order or even smaller as compared to the spatial extensions of the electronic wavefunction. This would be expected to result in strongly modified optical selection rules and, thus, enables further enhancement of spontaneous emission, an essential requirement for applications in diverse scientific fields such as quantum information processing, optical quantum computing or light harvesting devices.

Appendix

Sample fabrication

The semiconductor heterostructure was grown by molecular beam epitaxy on a (001) GaAs substrate. After deposition of 500 nm GaAs, we grew 50 nm of Al0.3Ga0.7 As. Self-assembled InAs quantum dots were formed at a temperature of 485 °C during a growth time of 35 s. After a growth interruption of 70 s the quantum dots were capped by 10 nm of Al0.3Ga0.7 As to make them optically active. The quantum dot density is estimated to ~ 100 µm−2, based on a study of different quantum dot samples grown under nominally identical growth conditions, which allowed atomic force microscopy measurements of uncapped surface quantum dots [33].

The plasmonic antennas investigated were defined on semi-insulating (001) GaAs wafers. After cleavage, the samples were flushed with acetone and isopropanol (IPA). In order to get a better adhesion of the e-beam resist, the samples were put on a hot plate (170 °C) for 300 s. An e-beam resist (Polymethylmethacrylat 950 K, AR-P 679.02, ALLRESIST) was coated at 4000 rpm for 40 s at an acceleration of 2000 rpm/s and baked out at 170 °C for 300 s, producing a resist thickness of 70 ± 5 nm. The samples were illuminated in a Raith E-line system using an acceleration voltage of 30 kV and an aperture of 10 µm. A dose test was performed for every fabrication run, as this crucial parameter depends on the varying e-beam current, with typical values ranging between 1200 µC/cm2 and 1500 µC/cm2. All samples were developed in Methylisobutylketon diluted with IPA (1 : 3) for 45 s. To stop the development, the sample was rinsed with pure IPA. For the metallisation an e-beam evaporator was used to deposit 35 nm of gold at a low rate of 0.1 nm/s. The lift-off was performed in warm acetone, leaving behind high-quality nanostructures with feature sizes on the order of 10 nm.

The sample contains several arrays of nominally identical bowtie nanoantennas with varying electron beam dose, feedgap size g and triangle size s and Au marker act as a reference for the spatially resolved photoluminescence scans as presented in Fig. 2.

Simulations

We simulated the scattering cross-section of the bowtie nanoantenna using a commercially available finite difference time domain solver (Lumerical Solutions, Inc., FDTD solutions, version: 8.11.387). We used a three-dimensional simulation cell that is terminated by perfectly matched layers. The bowtie was modelled using the extruded N-sided equilateral polygon with rounded corners that is provided on the Lumerical homepage (https://kb.lumerical.com/en/ref_sim_obj_creating_rounded_corners.html). At the centre of the simulation cell, i.e. around the bowtie feedgap region, we used a mesh size of 2 nm, whereas in the outer regions the value was set to 4 nm. The computation of the scattering cross-section is based on the Mie scattering tutorial that can be found on the Lumerical homepage (https://kb.lumerical.com/en/sp_fluorescence_enhancement.html). Consequently, we excited the structures using a total field scattered field (TFSF) source and a finite-difference time-domain (FDTD) scattered field monitor to compute the scattering cross-section. The simulation principle of the hybrid system is fully based on the “Fluorescence Enhancement” tutorial that is also published on the Lumerical homepage (https://kb.lumerical.com/en/sp_fluorescence_enhancement.html).

Optical measurements

For differential reflection spectroscopy, a white light continuum laser (Fianium WhiteLaseTM micro) is focused onto individual bowtie nanoantennas using an 100× microscope objective (Mitutoyo) with numerical aperture NA = 0.5. Excitation power density and linear polarisation are adjusted with a continuous power attenuator wheel and a broadband linear polariser (Thorlabs, LBVIS100 − MP2). The reflected signal from the bowtie nanoantennas is collected via the same objective, directed via a multimode optical fibre to a 0.5 m imaging spectrometer (Princeton Instruments Acton SP2500i, grating: 300 l/mm) and analysed with either a Si-CCD camera (Princeton Instruments Spec-10) or InGaAs diode array (Princeton Instruments, OMA V). The differential reflection spectrum is obtained from reflectivity spectra recorded at the position of the bowtie nanoantenna Ron and at a position on the bare substrate as a reference R0, by calculating (RonR0)/R0 [23].

Photoluminescence spectroscopy was carried out in a dip-stick cryostat at liquid Helium temperature T = 4.2 K. We used a continuous wave titanium-sapphire laser (Spectra-Physics Model 3900) with an excitation energy at 1.455 eV to excite the quantum dots via the two-dimensional wetting layer. In addition, a defocused HeNe laser emitting above-bandgap at an emission energy of 1.962 eV was used to stabilise the fluctuating charge environment around the quantum dot. Hereby, we used an excitation power density low enough to guarantee that no considerable photoluminescence signal from the quantum dots is generated, but high enough to reduce the quantum dot linewidth. The excitation power density was continuously monitored and stabilised to less than 0.1 % fluctuations and a uniform input polarisation was achieved by a fixed linear polariser (Thorlabs, LBVIS100 − MP2). The excitation was focused onto the sample by a home-built low-temperature, achromatic microscope objective with NA = 0.7. In the excitation path, an additional half wave plate (Thorlabs, AHWP10M-980) was introduced that allowed for tuning of the excitation polarisation. The emitted photoluminescence signal from individual quantum dots was collected via the same objective and analysed subsequently with a linear polariser (Thorlabs, LBV IS100 − MP2), before guided via a polarisation-maintaining optical fibre to a spectrometer (Jobin Yvon Triax 550, grating 600 l/mm) equipped with a Si-CCD camera (Jobin Yvon Symphony CCD). For time-resolved measurements, the spectrometer was used as a monochromator to spectrally filter the emission of a single quantum dot before it was detected by a single photon avalanche detector (EG&G Photon Counting Module SPCM-200) and analysed with time tagging electronics (Picoquant TimeHarp 220).

Funding

Deutsche Forschungsgemeinschaft (DFG) (SFB 631, the German Excellence Initiative via NIM); Technische Universität München (TUM) (Institute for Advanced Study, TUM International Graduate School of Science and Engineering (IGSSE)); RFBR (project no. 16-37-60075 mol_a_dk); Fellowship of Russian President (SP-3014.2016.3);

Acknowledgments

A.A.L. acknowledges financial support of RFBR via the project no. 16-37-60075 mol_a_dk and the fellowship of Russian President SP-3014.2016.3.

References and links

1. S. Maier, Plasmonics: Fundamentals and Applications (Springer US, 2007).

2. S. Lal, S. Link, and N. J. Halas, “Nano-optics from sensing to waveguiding,” Nat. Photonics 1, 641–648 (2007). [CrossRef]  

3. J. A. Schuller, E. S. Barnard, W. Cai, Y. C. Jun, J. S. White, and M. L. Brongersma, “Plasmonics for extreme light concentration and manipulation,” Nat. Mater. 9, 193–204 (2010). [CrossRef]   [PubMed]  

4. E. Ozbay, “Plasmonics: merging photonics and electronics at nanoscale dimensions,” Science 311, 189–193 (2006). [CrossRef]   [PubMed]  

5. R. Zia, J. A. Schuller, A. Chandran, and M. L. Brongersma, “Plasmonics: the next chip-scale technology The development of chip-scale electronics and photonics has led to,” Mater. Today 9, 20–27 (2006). [CrossRef]  

6. D. P. Fromm, A. Sundaramurthy, P. J. Schuck, G. Kino, and W. E. Moerner, “Gap-Dependent Optical Coupling of Single “Bowtie” nanoantennas resonant in the visible,” Nano Lett. 4, 957–961 (2004). [CrossRef]  

7. K. Schraml, M. Spiegl, M. Kammerlocher, G. Bracher, J. Bartl, T. Campbell, J. J. Finley, and M. Kaniber, “Optical properties and interparticle coupling of plasmonic bowtie nanoantennas on a semiconducting substrate,” Phys. Rev. B 90, 035435 (2014). [CrossRef]  

8. G. Bracher, K. Schraml, M. Blauth, J. Wierzbowski, N. Coca-López, M. Bichler, K. Müller, J. J. Finley, and M. Kaniber, “Imaging surface plasmon polaritons using proximal self-assembled InGaAs quantum dots,” J. Appl. Phys. 116, 033101 (2014). [CrossRef]  

9. G. Bracher, K. Schraml, M. Ossiander, S. Frédérick, J. J. Finley, and M. Kaniber, “Optical study of lithographically defined, subwavelength plasmonic wires and their coupling to embedded quantum emitters,” Nanotechnology 25, 075203 (2014). [CrossRef]   [PubMed]  

10. E. Goldmann, M. Paul, F. F. Krause, K. Müller, J. Kettler, T. Mehrtens, A. Rosenauer, M. Jetter, P. Michler, and F. Jahnke, “Structural and emission properties of InGaAs/GaAs quantum dots emitting at 1.3 µ m,” Appl. Phys. Lett. 105, 152102 (2014). [CrossRef]  

11. J. L. O’Brien, A. Furusawa, and J. Vučković, “Photonic quantum technologies”, Nat. Photonics 3, 687–695 (2010). [CrossRef]  

12. D. Bimberg, M. Grundmann, and N. Ledentsov, Quantum Dot Heterostructures (Wiley, 1999).

13. A. W. Schell, G. Kewes, T. Hanke, A. Leitenstorfer, R. Bratschitsch, O. Benson, and T. Aichele, “Single defect centers in diamond nanocrystals as quantum probes for plasmonic nanostructures,” Opt. Express 19, 7914–7920 (2011). [CrossRef]   [PubMed]  

14. T. B. Hoang, G. M. Akselrod, C. Argyropoulos, J. Huang, D. R. Smith, and M. H. Mikkelsen, “Ultrafast spontaneous emission source using plasmonic nanoantennas,” Nat. Commun. 6, 7788 (2015). [CrossRef]   [PubMed]  

15. T. B. Hoang, G. M. Akselrod, and M. H. Mikkelsen, “Ultrafast room-temperature single photon emission from quantum dots coupled to plasmonic nanocavities,” Nano Lett. 16, 270–275 (2016). [CrossRef]  

16. A. Kinkhabwala, Z. Yu, S. Fan, Y. Avlasevich, K. Müllen, and W. E. Moerner, “Large single-molecule fluorescence enhancements produced by a bowtie nanoantenna,” Nat. Photonics 3, 654–657 (2009). [CrossRef]  

17. G. M. Akselrod, C. Argyropoulos, T. B. Hoang, C. Ciracì, C. Fang, J. Huang, D. R. Smith, and M. H. Mikkelsen, “Probing the mechanisms of large Purcell enhancement in plasmonic nanoantennas,” Nat. Photonics 8, 835–840 (2014). [CrossRef]  

18. J. C. Prangsma, J. Kern, A. G. Knapp, S. Grossmann, M. Emmerling, M. Kamp, and B. Hecht, “Electrically connected resonant optical antennas,” Nano Lett. 12, 3915–3919 (2012). [CrossRef]   [PubMed]  

19. P. Biagioni, J. S. Huang, L. Duò, M. Finazzi, and B. Hecht, “Cross resonant optical antenna,” Phys. Rev. Lett. 102, 256801 (2009). [CrossRef]   [PubMed]  

20. M. L. Andersen, S. Stobbe, A. S. Sørensen, and P. Lodahl, “Strongly modified plasmon-matter interaction with mesoscopic quantum emitters,” Nat. Phys. 7, 215–218 (2011). [CrossRef]  

21. K. Adlkofer, E. F. Duijs, F. Findeis, M. Bichler, A. Zrenner, E. Sackmann, G. Abstreiter, and M. Tanaka, “Enhancement of photoluminescence from near-surface quantum dots by suppression of surface state density,” Phys. Chem. Chem. Phys. 4, 785–790 (2002). [CrossRef]  

22. J. Zuloaga and P. Nordlander, “On the energy shift between near-field and far-field peak intensities in localized plasmon systems,” Nano Lett. 11, 1280–1283 (2011). [CrossRef]   [PubMed]  

23. M. Kaniber, K. Schraml, A. Regler, J. Bartl, G. Glashagen, F. Flassig, J. Wierzbowski, and J. J. Finley, “Surface plasmon resonance spectroscopy of single bowtie nano-antennas using a differential reflectivity method,” Sci. Rep. 6, 23203 (2016). [CrossRef]   [PubMed]  

24. T.-P. Hsieh, J.-I. Chyi, H.-S. Chang, W.-Y. Chen, T. M. Hsu, and W.-H. Chang, “Single photon emission from an InGaAs quantum dot precisely positioned on a nanoplane,” Appl. Phys. Lett. 90, 073105 (2007). [CrossRef]  

25. M. Pfeiffer, K. Lindfors, H. Zhang, B. Fenk, F. Phillipp, P. Atkinson, A. Rastelli, O. G. Schmidt, H. Giessen, and M. Lippitz, “Eleven nanometer alignment precision of a plasmonic nanoantenna with a self-assembled GaAs quantum dot,” Nano Lett. 14, 197–201 (2014). [CrossRef]  

26. M. Pfeiffer, K. Lindfors, C. Wolpert, P. Atkinson, M. Benyoucef, A. Rastelli, O. G. Schmidt, H. Giessen, and M. Lippitz, “Enhancing the optical excitation efficiency of a single self-assembled quantum dot with a plasmonic nanoantenna,” Nano Lett. 10, 4555–4558 (2010). [CrossRef]   [PubMed]  

27. A. G. Curto, G. Volpe, T. H. Taminiau, M. P. Kreuzer, R. Quidant, and N. F. van Hulst, “Unidirectional emission of a quantum dot coupled to a nanoantenna,” Science 329, 930–933 (2010). [CrossRef]   [PubMed]  

28. E. M. Purcell, “Spontaneous emission probabilities at radio frequencies,” Phys. Rev. 69, 681 (1946).

29. P. W. Fry, I. E. Itskevich, D. J. Mowbray, M. S. Skolnick, J. J. Finley, J. A. Barker, E. P. O’Reilly, L. R. Wilson, I. A. Larkin, P. A. Maksym, M. Hopkinson, M. Al-Khafaji, J. P. David, A. G. Cullis, G. Hill, and J. C. Clark, “Inverted electron-hole alignment in InAs-GaAs self-assembled quantum dots,” Phys. Rev. Lett. 84, 733–736 (2000). [CrossRef]   [PubMed]  

30. E. B. Flagg, A. Muller, J. W. Robertson, S. Founta, D. G. Deppe, M. Xiao, W. Ma, G. J. Salamo, and C. K. Shih, “Resonantly driven coherent oscillations in a solid-state quantum emitter,” Nat. Phys. 5, 203–207 (2009). [CrossRef]  

31. H. Zhang, Y. Huo, K. Lindfors, Y. Chen, O. G. Schmidt, A. Rastelli, and M. Lippitz, “Narrow-line self-assembled GaAs quantum dots for plasmonics,” Appl. Phys. Lett. 106, 101110 (2015). [CrossRef]  

32. A. R. Smith, K.-J. Chao, Q. Niu, and C.-K. Shih, “Formation of atomically flat silver films on GaAs with a “Silver Mean” quasi periodicity,” Science 273, 226–228 (1996). [CrossRef]   [PubMed]  

33. A. A. Lyamkina, S. P. Moshchenko, D. V. Dmitriev, A. I. Toropov, and T. S. Shamirzaev, “Exciton-plasmon interaction in hybrid quantum dot/metal cluster structures fabricated by molecular-beam epitaxy,” JETP Letters 99, 219 (2014). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (4)

Fig. 1
Fig. 1 Sample layout, structural and optical characterisation: (a) Schematic illustration of the monolithically integrated quantum dot-nanoantenna system. (b) Scanning electron microscopy image of the bowtie nanoantenna array (left) and an individual bowtie nanoantenna (right) with s = 87 nm and g = 26 nm. Scale bar, 20 nm. (c) Top: Simulated scattering cross-section σscatt and normalised decay rate Γ/Γ0 of a bowtie nanoantenna with nominal s0 = 90 nm and g0 = 15 nm in blue and red, respectively. Bottom: Measured quantum dot spectrum (L: laser, C: carbon impurities) and differential reflectivity of a bowtie nanoantenna with s = 87 nm and g = 26 nm in red/green and grey, respectively. Blue curve shows a 50-point-smoothed average.
Fig. 2
Fig. 2 Enhanced luminescence from coupled quantum dot-nanoantenna systems: (a) Spatially resolved laser reflectivity, co-, cross-polarised and differential quantum dot emission from left to right, respectively. (b), (c) Photoluminescence spectra of uncoupled and coupled quantum dots for co- and cross-polarised emission in blue and red, respectively.
Fig. 3
Fig. 3 Strong linearly polarised emission from coupled quantum dot-nanoantenna systems: (a) Polarisation-resolved photoluminescence for a coupled quantum dot-nanoantenna system (top) and an uncoupled quantum dot (bottom). (b) Corresponding peak intensities of a coupled quantum dot-nanoantenna system (top) and an uncouped quantum dot (bottom) in a polar-plot representation. (c) Degree of linear polarisation ρ for coupled quantum dot-nanoantenna systems (red) and uncoupled quantum dots (blue).
Fig. 4
Fig. 4 Purcell-enhanced emission from coupled quantum dot-nanoantenna systems: (a) Time-resolved photoluminescence transients of a coupled quantum dot-nanoantenna system (red), an uncoupled quantum dot (blue) and the instrument response function (black). (b) Histogram of the spontaneous emission lifetimes for coupled quantum dot-nanoantenna systems (red) and uncoupled quantum dots (blue). (c) Correlation between the spontaneous emission lifetime and the degree of linear polarisation.
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.