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

Prospects and limitations of bottom-up fabricated hollow-core waveguides

Open Access Open Access

Abstract

We perform a numerical simulation study of hollow-core anti-resonant reflection optical waveguides (ARROWs) fabricated using lithography and material deposition in the context of their suitability as a platform for on-chip photonic quantum information processing. We explore the effects of the core size, the number of pairs of anti-resonant layers surrounding the hollow core, and the refractive index contrast between the anti-resonant layer materials on propagation losses in the waveguide. Additionally, we investigate the feasibility of integrating these waveguides with Bragg gratings and dielectric metasurfaces to form on-chip cavities that could act as nonlinear optical elements controllable with single photons when loaded with atomic ensembles.

© 2016 Optical Society of America

1. Introduction

Hollow-core waveguides filled with atomic ensembles offer a promising platform for realizing mediated interactions [1] between single photons needed to implement a variety of quantum information protocols. The waveguides can combine low losses for the propagating light with small optical mode areas, which in turn can give rise to strong light-matter interactions enabling extremely low-power optical nonlinearities in dilute gasses confined in the waveguides, especially when coherent light-matter interaction techniques are used. Several notable demonstrations of such strong light-matter interaction in hollow-core waveguides have been reported in recent years, enabling for example, all-optical switching with a few-hundred photons in a hollow-core photonic crystal fibre filled with laser-cooled atoms [2], cross-phase modulation with few photons [3], and single-photon broadband quantum memory [4] in a photonic-crystal fibre filled with room-temperature alkali atoms, as well as demonstration of quantum state control of warm alkali vapor in a hollow-core anti-resonant reflecting optical waveguides (ARROW) on a chip [5].

Used extensively for microfluidics applications [6], the lithographically defined hollow-core ARROWs have propagating modes of similar areas (~10λ2) as hollow-core photonic-crystal fibres (HCPCFs) [7]. However, compared to HCPCFs, hollow-core ARROWs offer some complicated trade-offs between loss performance and ease of fabrication and monolithic integration. Unlike hollow-core fibres, these on-chip waveguides do not need a drawing tower and can be made with a combination of photolithography and relatively basic material deposition tools. Since they are fabricated in a bottom-up fashion through deposition of alternating layers of two dielectric materials, one can additionally envision including steps into the fabrication process to control the photonic environment of the waveguide, such as adding a Bragg grating as shown in Fig. 1(d). At the same time, propagation losses reported for on-chip air-core ARROWs range between ~ 0.1 cm−1 (~ 0.4 dB/cm) for numerical simulations and ~ 1 cm−1 (~ 4 dB/cm) for fabricated structures, which does not compare favorably with the losses achieved in HCPCFs ranging from < 250 dB/km for off-the-shelf fibres down to ~ 1 dB/km [8] – a value comparable to high-performance conventional solid-core fibres [9].

 figure: Fig. 1

Fig. 1 Schematic drawing of the cross sections of an ARROW: (a) lengthwise; transverse cross section of (b) a conventional hollow-core self-aligned pedestal ARROW that has a high index layer beneath the core (n1 > n2) and (c) a hollow-core self-aligned pedestal ARROW that has a low index layer beneath the core; (d) conceptual vision of an integrated opto-electronic system combining an ARROW loaded with atomic or molecular vapor with a Bragg grating and micro-wires that could be used to, e.g., generate magnetic field.

Download Full Size | PDF

Here, we explore the limits of hollow-core ARROWs posed by their propagation loss and evaluate the feasibility of integrating mirrors and cavities with these waveguides, focusing on mirror structures that would leave the hollow core of the waveguide unobstructed and allow loading of atomic vapors.

2. Design principles and propagation loss

Antiresonant reflection optical waveguides (ARROWs) were first proposed and demonstrated by Duguay et al. [10] in 1986 for light propagation in low index media. ARROWs consist of a core surrounded by multiple dielectric layers whose thicknesses are chosen such that there is destructive interference of the multiple incident and reflected waves in each cladding layer, and constructive interference in the core for a given wavelength. The cladding acts as a highly reflective mirror that confines light to the core and results in low loss propagation. The initial demonstration guided light in a silicon dioxide core using a single pair of Si/SiO2 antiresonant layers below and total internal reflection from the air above for confinement. Hollow core ARROW waveguides were first demonstrated by Delonge and Fouckhardt [11] in 1995. Combining the constructive interference condition in the core with the destructive interference conditions in the cladding, one can obtain an analytic expression for the thickness of each cladding layer i of the ARROW with a core of refractive index nc and thickness dc [10] (Fig. 1(a)):

ti=λ4ni(2N+1)(1nc2ni2+λ24ni2dc2)12
where usually N=0 but can be any non-negative integer. We note that this antiresonance condition differs slightly from the Bragg condition for normal incidence (ti = λ/4ni) on a quarter-wavelength stack. One of the advantages of ARROWs is that the cladding layers don’t need to be periodic with respect to the refractive index or their thicknesses as long as they satisfy the antiresonance condition specified in equation (1), although we will only consider two different materials to be used for the cladding layers here.

The loss of the waveguide can be estimated as the average of the losses through the left and right cladding layers, αhorizontal, and the losses through the top and bottom cladding layers, αvertical, with

αvertical=2RtopRbottom2hctan(θc,v)αhorizontal=2RleftRright2wctan(θc,h)
obtained by calculating the reflectivities Rtop, Rbottom, Rleft, and Rright of the dielectric stacks surrounding the core given the incident glancing angle [12, 13]. Here, hc is the height of the core, wc is the width of the core, and θc,v(h) is the glancing angle along the vertical (horizontal) direction in the core defined by sin θc,v = λ/2nchc (sin θc,h = λ/2ncwc), so αwaveguide = (αvertical + αhorizontal)/2.

Several possible design variants for the hollow-core ARROW are described in the literature, out of which the self-aligned pedestal (SAP) ARROW (Fig. 1(b)) has been reported to have the lowest losses both in numerical predictions (0.36cm−1) and for fabricated devices (2.2cm−1) [14] thanks to having air as the terminating layer on both sides [15]. The higher experimental losses compared to the results of numerical simulations arise from the deviation of the cladding layer thicknesses from the ideal values and from the surface roughness which results in additional scattering losses.

The modal characteristics of the ARROW can be found using photonic modeling software such as the finite element solver FemSIM by RSoft. Fig. 2(a) shows the electric field distribution of the fundamental mode of the ARROW waveguide and, at at the same time, visualizes the dynamics of the relatively large propagation losses. As the waveguide relies on anti-resonant reflection instead of, e.g. bandgap or inhbited coupling, even the fundamental mode remains inherently leaky, radiating energy into surrounding space as it propagates. The losses of a structure based on the analytical description of equation (1) can be somewhat reduced by numerically optimizing the thickness of each cladding layer, e.g. by using the RSoft MOST optimization tool, for a specific wavelength and the loss spectrum for a waveguide with 5.8µm × 12µm hollow core surrounded by three pairs of anti-resonant layers designed for 852nm is shown in Fig. 2(b). The waveguide loss is calculated from the imaginary part of the effective propagation index of the fundamental mode as α=4πIm{neff}λ

 figure: Fig. 2

Fig. 2 Numerical simulation (a) of the transverse profile of the fundamental waveguide mode and (b) of the propagation loss spectrum of a hollow-core ARROW with a 5.8µm × 12µm core surrounded by 3 pairs of Si3N4(1.74)/SiO2 cladding layers arranged as shown in Fig. 1(b), designed and optimized for a wavelength of 852 nm. (c) Numerical simulation of propagation losses in waveguides optimized for 852 nm light with assorted hollow core sizes and three pairs of of Si3N4(1.74)/SiO2 cladding layers as shown in Fig. 1(b). (d) Effects of the number of cladding layer pairs and of the refractive index contrast between the two materials forming each pair on the simulated propagation losses for waveguides with 5.8µm × 12µm hollow cores. The square data points correspond to the cladding layers arranged as shown in Fig. 1(b), while the triangular data points represent cladding layers arranged as shown in Fig. 1(c). The value in brackets denotes the refractive index of the cladding layer material.

Download Full Size | PDF

We now proceed to explore the effects of what one could call ’major’ design parameters, namely the core size, the number of anti-resonant layer pairs, and the refractive index contrast between the layers in each pair, on the propagating losses of the fundamental mode in the waveguide. The results are presented in Fig. 2(c) and 2(d). The points in these plots represent waveguides with the thickness of each anti-resonant cladding layer optimized using the RSoft MOST optimization tool.

Fig. 2(c) presents the waveguide loss for a selection of waveguide core sizes, for which the area of the propagating mode is defined by

Amode=ϵ(x,y)|E(x,y)|2dxdymax{ϵ(x,y)|E(x,y)|2}
To achieve low-power optical nonlinearities, small mode field areas are desirable, as the tighter transverse confinement increased the field intensity of individual photons. However, as the individual dimensions of the hollow core decrease, the light will propagate with a larger glancing angle and increased loss, as can be seen from equation (2). Additionally, for the fundamental TE mode, the light has TM like polarization relative to the sidewalls. As a result, due to the same phenomenon that gives rise to the Brewster angle, the reflectivity of the TM like waves off the sidewalls will be lower, which increases the propagation loss. This is why the width of the core is typically chosen to be larger than its height but, overall, decreasing the size of the hollow core leads to a rapid increase of the propagation loss. These loss values in Fig. 2(c) and 2(d) were found by first choosing the thicknesses of the cladding layers according to the antiresonance condition given by equation (1), and locating the fundamental quasi-Gaussian mode of the waveguide. Using this mode as a seed, an optimization procedure is performed whereby the thicknesses of each layer are varied according to a genetic algorithm in the finite element modeling and the optimization tool by RSoft. The loss is slowly minimized and the optimization tool determines the corresponding cladding layer thicknesses to achieve it.

The dielectrics used for the cladding layers have included Si3N4(n = 1.7 − 2), SiO2 (n = 1.44), TiO2 (n = 2.4), and Ta2O5 (n = 2.1). A Si3N4/SiO2 pair has been commonly used [14–16], although a combination of TiO2 [11] or Ta2O5 [17] along with SiO2 has been explored as well. Si3N4 has been typically used as as the high index antiresonant layer, given its ease of deposition using plasma enhanced chemical vapor deposition (PECVD), relatively high refractive index (n = 1.7 − 2, depending on the deposition parameters), and negligible absorption in the desired optical range [18].

Fig. 2(d) presents the results of our numerical simulations exploring how the waveguide loss changes with the number of layer pairs surrounding the core for three kinds of cladding layer pairs. We chose pairs consisting of SiO2 (n = 1.44) and Si3N4(n = 2), Si3N4(n = 1.7), and TiO2(n = 2.4) as the high index material. The low index Si3N4 corresponds to the material deposited by the PECVD machine at University of Waterloo Quantum Nanofab, while n = 2.0 is the material’s refractive index commonly reported in the literature. We included TiO2 as its refractive index appears to be one of the highest available for non-absorptive materials at our target wavelength. Note that our numerical optimization included the conformality factor arising in the PECVD process, where vertical and horizontal films are deposited at different rates. In all of our simulations, the ratio between the thicknesses of the horizontal and vertical cladding layers is set to be 1.25 for Si3N4 and TiO2 and 1.58 for SiO2 (1.63 for the thick outer layer). Increasing the refractive index contrast between the cladding materials appears to have the strongest effect on reducing the propagation loss and we observe a loss reduction by a factor of approximately three for a given number of layer pairs when the index contrast changes from 0.3 to 0.96. On the other hand, we observe that adding more periods does not necessarily yield a steady decrease in the mode loss – going from 2 to 3 periods decreases the loss more than by going from 3 to 4 periods – and there seems to be a lower bound on the loss for a given material pair. It should be noted that the 4 period TiO2/SiO2 waveguide loss could not be lowered below that of the 3 period version during the multiple optimization runs. However, we believe that with some more tweaking of the optimization parameter space, this should be possible. In absolute values, the propagation losses presented here are somewhat higher that than those reported in Ref. [14] a 5.8µm × 12µm hollow core waveguide with three Si3N4(n = 2.0)/SiO2 cladding layer pairs, which is most likely the result of slightly higher (n = 2.05) refractive index of Si3N4 and using fewer mesh divisions in the FIMMWAVE simulation used in that work.

Lastly, the waveguides in in the literature [11, 14–19] surround the core with the high index layer as shown in Fig. 1(b) because it is intuitively expected for the higher Fresnel reflection coefficient at the interface to result in a lower overall waveguide loss. However, another possibility that has not yet been explored is to have the layer immediately beneath the core be the low index layer as shown in Fig. 1(c). The propagation losses of the optimized waveguides for this variant are compared with that of the typical configurations in Fig. 2(d) and we find that there is not a significant difference in the propagation loss between the two variants given the same cladding material pair, and number of anti-resonant layers used. In each case, the waveguides can be optimized such that both variants have comparable propagation losses. We were forced to explore this new design variant with a SiO2 rather than a silicon nitride layer used beneath the core, when we discovered that the SU8 sacrificial core had poor adhesion to silicon nitride films from our PECVD machine and was washed out during the development step. While the SU8 core had excellent adhesion to silicon substrate and silicon oxide films, our silicon nitride films seem to have a lower ratio of silicon to nitrogen that results in problems with resist adhesion and a low refractive index.

3. Bragg mirrors in hollow-core ARROWs

The interaction between light and an atomic ensemble confined to the hollow-core ARROW could be further increased by integrating the waveguide with a pair of mirrors to form a cavity. For solid-core waveguides, mirrors may be realized with Bragg gratings which are implemented by periodically modulating the refractive index in the waveguide core [20]. This presents a challenge for hollow-core waveguides as there is no material in the core that could be modified. Here, we explore implementing a grating by modulating the effective index of propagation. We propose to do this by etching a grating in the various anti-resonant layers. The waveguide will then have a slightly different propagation index along the section where one of cladding layers above or below the core has been etched. Given that the ARROW has a bottom up fabrication process, the grating can be implemented in any particular layer. This approach has the advantage of leaving the hollow core unobstructed so that atomic ensembles can be loaded into the waveguide.

As shown in Figure 3(a), we will consider cases where the grating is etched either into the layer immediately above the core, the layer beneath the core, or the very top layer. Figure 3(b) then presents how the effective index of propagation and the propagation loss change for the etched region for the three different locations of the grating, in this case for a Si3N4(n = 1.74)/SiO2 hollow core ARROW with three cladding layer pairs. We see that changing the thickness of a particular anti-resonant layer increases the propagation loss because the waveguide is no longer optimized. Layers closer to the core have a greater overlap with the fundamental leaky mode, hence etching them creates a larger perturbation relative to the original waveguide. This explains why etching the layer beneath the core causes a large change in the effective index of propagation (~ 5 × 10−4) whereas it is only 2.9 × 10−4 when etching the very top layer. Overall, we found that modulating the thickness of the layer beneath the core results in highest Δneff and the smallest increase in propagation loss. At the same time, increasing the depth of modulation past certain value for all three layers results in complete loss of guiding for the fundamental mode. We found that using higher refractive index contrast cladding materials yields a waveguide with lower loss where the optical mode is more confined to the core. There is then less overlap between the anti-resonant layers and the the fundamental mode, causing a smaller perturbation. For comparison, the same geometry ARROW as the one from Fig. 3(b) but with n = 2 for the Si3N4 will have maximum effective index change of approximately (~ 4 × 10−4). Using more periods of anti-resonant layers has the same effect because the lower loss waveguide has a smaller overlap between the cladding layers and the fundamental mode which is again more tightly confined.

 figure: Fig. 3

Fig. 3 (a) Etched gratings in the various anti-resonant layers: first layer above the core (left), first layer beneath the core (middle), and the top layer (right). (b) Change in the effective index of propagation Δneff for a hollow-core ARROW with 5.8µm × 12µm core surrounded by three pairs of Si3N4(1.74)/SiO2 cladding layers. The x-axis refers to the etch depth δt relative to the original thickness t of that particular layer. When etching the very top layer, at etch depths shown to induce uncharacteristically large effective propagation index contrasts and losses, the mode is no longer quasi-Gaussian within the core. (c) Estimated best mirror reflectivities of the Bragg gratings formed by etching the three different cladding layers obtained using MSE with plane wave incidence. The dash-dot horizontal lines mark the maximum reflectivity for grating in a particular layer. (d) Estimated reflectivity spectrum of a 10000 period Bragg mirror created by periodically etching the anti-resonant layer beneath the core. The etch depth is 70% of this layer, which creates an effective propagation index contrast of 5 × 10−4 with a 8% increase in loss.

Download Full Size | PDF

Given that the waveguide has non-negligible loss, the standard transfer matrix method [21] cannot be used to determine the reflectivity of these gratings. We can instead use the Method of Single Expression (MSE) proposed by Baghdasaryan et al. [22] to properly account for the absorption loss in each grating layer. Starting from the transmitting side, the field in each Bragg layer is calculated as we travel back to the illuminated side by numerically solving the coupled differential equations below.

dU(z)d(k0z)=Y(z)
dY(z)d(k0z)=P2(z)U3(z)Re{ϵ(z)}U(z)
dP(z)d(k0z)=Im{ϵ(z)}U2(z),
where the electric field is E(z) = U(z)eiS(z), Y(z) = dU(z)/d(k0z), and the power is P(z) = U2(z)/[dS(z)/d(k0z)]. Given a grating of length L, the initial conditions are U(L) = 1, Y(L) = 0, and P(L)=ϵt where ϵt is the relative permittivity of the transmitting side. The reflectivity is then given by
R=|U2(0)ϵiP(0)jU(0)Y(0)U2(0)ϵi+P(0)+jU(0)Y(0)|2
where ϵi is the relative permittivity of the incident side. Now our grating can be thought of as being composed of two alternating dielectric materials n1 (the effective index of propagation of the waveguide), and n2 (the effective index of propagation when one of the anti-resonant layers has been etched). This approximation is valid because the modes in the two regions are similarly quasi-Gaussian in the core. The corresponding Bragg-condition thicknesses are t1 = λ/4·Re{n1}, and t2 = λ/4 · Re{n2}. For the three period Si3N4(n = 1.74)/SiO2 ARROW from Fig. 3(b), the reflectivity of the mirror is shown in Fig. 3(c) for various etched layer cases where the etch depth is chosen to have the highest effective index contrast relative to the original waveguide. While this is an approximate result obtained by assuming plane wave propagation, we have found previously [23] for a different type of hollow-core waveguide that this estimate has an excellent agreement with a full three dimensional simulation. Implementing the grating in the anti-resonant layer beneath the core yields the highest reflectivity (~84%) mirror. The next best mirror with ~ 55% reflectivity can be implemented using the layer above the core, while using the top layer yields a poor mirror (~ 23%) due to its low effective index contrast. Note that for a typical lossless Bragg grating, the reflectivity increases with higher refractive contrast materials and larger number of periods and, regardless of the refractive index contrast, eventually approaches 1 as we add more periods. However, with our ARROW gratings the upper bound on the reflectivity is less than 1 due to the presence of absorption loss.

4. Integrated cavities

Depending on the application, several parameters can be used to characterize the performance of an optical cavity, including its mode volume Vmode, quality factor Q, and cooperativity factor C = g2/κγ. Cavities with high-cooperativity factor, C ≫ 1 or at least C > 1, are of particular interest as the presence of a single atom in such cavity will result in a significant change of the cavity transmission [24]. We now proceed to discuss the feasibility of creating a high-cooperativity cavity a formed by a pair Bragg gratings around a spacer region inside the hollow-core ARROW.

In the expression for the cooperativity factor, κ = ω/Q describes the field decay rate from the cavity, γ is the linewidth of the excited state of the atom placed in the cavity, and g is the atom-field coupling coefficient or single-photon Rabi frequency:

g=μω2ϵMVmode.
Here, µ is the atomic transition dipole moment, ω is the cavity resonant frequency, and ϵM is the permittivity where the field intensity is highest and the mode volume is given by
Vmode=Vϵ|E|2dVmax{ϵ|E|2}=AmodeLeff
with Amode being the waveguide mode area and Leff the effective length of the cavity.

At the same time, the cooperativity can be expressed independent of the atomic species present in the cavity, such that

g2κγ=316π2QVmode/λ3
and is determined solely by the properties of the cavity. To estimate the Q factor of a cavity formed by a pair of Bragg mirrors in a hollow core arrow, we consider the Q factor of a simple Fabry-Perot cavity of length L with perfectly thin mirrors with reflectivities R, which is filled with a medium with effective index n and attenuation α:
Q=ωnLc(lnRαL)

The cavity length, L, can then be approximated as Leff = Lspacer + 2zp, with the penetration depth [25] into the Bragg mirrors zpaneff. Here, aλ/2neff is the Bragg grating period and the approximate expression for zp assumes neff ≈ 1 and Δneff ≪ 1. Further assuming that the propagation loss in the etched sections of the waveguide is roughly the same as in the non-etched sections, αwg, we can estimate the reflectivity of the Bragg mirror as Re2zpαwg [23]. Combining these, we arrive at an expression estimating the upper limit for the cooperativity of our Bragg grating cavity for Lspacerzp:

g2κγ316πλAmodeΔneffαwg
Thus to achieve g2/κγ > 1 at λ = 852nm in a hollow-core waveguide with Amode = 18µm2 and Δneff = 5 × 10−4, a propagation loss αwg < 10−2 cm−1 would be needed. This value is more than an order of magnitude lower than even the simulated propagation loss that does not account for fabrication imperfections and might be out of reach.

An alternative approach to integrating a cavity into a hollow-core ARROW relies on attaching photonic-crystal membranes acting as dielectric metasurfaces at the two ends of a waveguide section as shown in Fig. 4(a). A dielectric metasurface [26] can be designed to act as a highly reflective mirror for light incident perpendicularly on it [27]. When comprised of a thin dielectric sheet with a pattern of holes, atomic ensembles can be still loaded into the waveguide through this porous membrane. The main parameters used to design such membrane are the refractive index of the dielectric, radius of the holes, spacing between the holes, and the thickness of the membrane.

 figure: Fig. 4

Fig. 4 (a) Schematic showing how the dielectric metasurface mirror would be integrated with the hollow-core ARROW waveguide. (b) Cooperativity of hollow-core ARROW (Amode = 18µm2) cavity formed with two dielectric metasurface mirrors for various waveguide lengths and mirror reflectivities.

Download Full Size | PDF

For a cavity formed by a pair of such membranes, Leff is then given just by the length of the waveguide section between the membranes. The cooperativity of such cavity can be then described as

g2κγ=38πλ2Amode1lnRαLeff,
where R is the reflectivity of the metasurface mirror. In the limit of |ln R| ≪ αLeff the cooperativity will be independent of the cavity length and determined purely by the reflectivity of the metasurface, while for |ln R| ≪ αLeff the cooperativity will decrease with increasing Leff. Fig. 4(b) then presents predicted cooperativities for an ARROW with Amode = 18µm2 for assorted cavity lengths and metasurface reflectivities. We see that for this type of ARROW-integrated cavity, high cooperativity might achievable for some of the waveguides structures from Fig. 2(d).

5. Conclusion

Compared to single quantum emitters, ensembles of quantum emitters can frequently offer significant advantages for implementing controllable single photon interactions. These advantages include collective enhancement of light-matter interaction, robustness to noise and decoherence, and the inherent ability to handle light pulses containing multiple photons. Combining atomic ensembles with on-chip waveguides could open new possibilities for control over the design of the atoms’ photonic environment and for design of scalable devices and architectures utilizing recently demonstrated quantum phenomena, such as vacuum-induced transparency [28] and superradiant lasing [29].

Unfortunately, we found that the propagation loss in hollow-core ARROWs is likely to pose severe limitation for their application in quantum information processing unless reasonably short sections are used. Bragg grating integration in particular results in poor-performance mirrors for wavelengths associated with alkali atoms, such as cesium, although it is possible that decent Bragg mirrors could be implemented, for example, at telecom wavelengths. On the positive side, it appears that combining ARROWs with photonic-crystal membranes acting as high-reflectivity dielectric metasurfaces could yield on-chip high-cooperativity cavities with lengths of ~100µm, which is comparable to the lengths of high-density atomic clouds produced by laser cooling In references [28] and [29].

Funding

This work was supported by Industry Canada and by Canada’s Natural Sciences and Research Council (NSERC) under the Discovery Grants Program. J. F. was in part supported by Ontario Graduate Fellowship.

References and links

1. D. E. Chang, V. Vuletić, and M. D. Lukin, “Quantum nonlinear optics — photon by photon,” Nat. Photonics 8, 685–694 (2014). [CrossRef]  

2. M. Bajcsy, S. Hofferberth, V. Balic, T. Peyronel, M. Hafezi, A. S. Zibrov, V. Vuletic, and M. D. Lukin, “Efficient all-optical switching using slow light within a hollow fiber,” Phys. Rev. Lett. 102, 203902 (2009). [CrossRef]   [PubMed]  

3. V. Venkataraman, K. Saha, and A. L. Gaeta, “Phase modulation at the few-photon level for weak-nonlinearity-based quantum computing,” Nat. Photonics 7, 138–141 (2013). [CrossRef]  

4. M. R. Sprague, P. S. Michelberger, T. F. M. Champion, D. G. England, J. Nunn, X.-M. Jin, W. S. Kolthammer, A. Abdolvand, P. S. J. Russell, and I. a. Walmsley, “Broadband single-photon-level memory in a hollow-core photonic crystal fibre,” Nat. Photonics 8, 287–291 (2014). [CrossRef]  

5. B. Wu, J. F. Hulbert, E. J. Lunt, K. Hurd, A. R. Hawkins, and H. Schmidt, “Slow light on a chip via atomic quantum state control,” Nat. Photonics 4, 776–779 (2010). [CrossRef]  

6. H. Schmidt and A. R. Hawkins, “The photonic integration of non-solid media using optofluidics,” Nat. Photonics 5, 598 (2011). [CrossRef]  

7. C. M. Haapamaki, J. Flannery, G. Bappi, R. A. Maruf, S. V. Bhaskara, O. Alshehri, T. Yoon, and M. Bajcsy, “Mesoscale cavities in hollow-core waveguides for quantum optics with atomic ensembles,” Nanophotonics 5, 392 (2016). [CrossRef]  

8. P. Roberts, F. Couny, H. Sabert, B. Mangan, D. Williams, L. Farr, M. Mason, A. Tomlinson, T. Birks, J. Knight, and P. S. J. Russell, “Ultimate low loss of hollow-core photonic crystal fibres,” Opt. Express 13, 236–244 (2005). [CrossRef]   [PubMed]  

9. K. Nagayama, M. Kakui, M. Matsui, T. Saitoh, and Y. Chigusa, “Ultra-low-loss (0.1484 dB/km) pure silica core fibre and extension of transmission distance,” Electron. Lett. 38, 1168–1169 (2002). [CrossRef]  

10. M. A. Duguay, Y. Kokubun, T. L. Koch, and L. Pfeiffer, “Antiresonant reflecting optical waveguides in SiO2 − Si multilayer structures,” Appl. Phys. Lett. 49, 13–15 (1986). [CrossRef]  

11. T. Delonge and H. Fouckhardt, “7th international symposium on high performance capillary electrophoresis integrated optical detection cell based on bragg reflecting waveguides,” J. Chromatogr. A 716, 135–139 (1995). [CrossRef]  

12. W. Huang, R. M. Shubair, A. Nathan, and Y. L. Chow, “The modal characteristics of arrow structures,” J. Lightw. Tech. 10, 1015 (1992). [CrossRef]  

13. B. S. Phillips, “Tailoring the spectral transmission of optofluidic waveguides,” Ph.D. thesis, Brigham Young University (2011).

14. E. J. Lunt, “Low-loss hollow waveguide platforms for optical sensing and manipulation,” Ph.D. thesis, Brigham Young University (2010).

15. D. Yin, J. P. Barber, A. R. Hawkins, and H. Schmidt, “Waveguide loss optimization in hollow-core arrow waveguides,” Opt. Express 13, 9331–9336 (2005). [CrossRef]   [PubMed]  

16. D. Yin, H. Schmidt, J. Barber, and A. Hawkins, “Integrated ARROW waveguides with hollow cores,” Opt. Express 12, 2710–2715 (2004). [CrossRef]   [PubMed]  

17. Y. Zhao, M. Jenkins, P. Measor, K. Leake, S. Liu, H. Schmidt, and A. R. Hawkins, “Hollow waveguides with low intrinsic photoluminescence fabricated with Ta2O5 and SiO2 films,” Appl. Phys. Lett. 98, 091104 (2011). [CrossRef]  

18. E. J. Lunt, B. Wu, J. M. Keeley, P. Measor, H. Schmidt, and A. R. Hawkins, “Hollow ARROW waveguides on self-aligned pedestals for improved geometry and transmission,” IEEE Photon. Technol. Lett. 22, 1147–1149 (2010). [CrossRef]  

19. R. Bernini, S. Campopiano, L. Zeni, and P. M. Sarro, “ARROW optical waveguides based sensors,” Sensors and Actuators B: Chemical 100, 143–146 (2004). [CrossRef]  New materials and Technologies in Sensor Applications, Proceedings of the European Materials Research Society 2003 - Symposium N.

20. A. Othonos, “Fiber Bragg gratings,” Rev. Sci. Instrum. 68, 4309–4341 (1997). [CrossRef]  

21. E. Hecht, Optics (Addison-Wesley, 1998), 4th ed.

22. H. Baghdasaryan, T. Knyazyan, T. Baghdasaryan, B. Witzigmann, and F. Roemer, “Absorption loss influence on optical characteristics of multilayer distributed bragg reflector: wavelength-scale analysis by the method of single expression,” Opto-Electronics Review 18, 438–445 (2010). [CrossRef]  

23. J. Flannery, G. Bappi, V. Bhaskara, O. Alshehri, and M. Bajcsy, “Implementing Bragg mirrors in a hollow-core photonic-crystal fiber,” (under review) (2016).

24. H. Tanji-Suzuki, I. D. Leroux, M. Schleier-Smith, M. Cetina, A. Grier, J. Simon, and V. Vuletić, “Interaction between Atomic Ensembles and Optical Resonators. Classical Description,” Adv. At. Mol. Opt. Phy. 60, 201–237 (2011). [CrossRef]  

25. J. D. Joannopoulos, S. G. Johnson, J. N. Winn, and R. D. Meade, “Photonic crystals: Molding the flow of light,” Princeton University Press (2008).

26. A. Arbabi, Y. Horie, M. Bagheri, and A. Faraon, “Dielectric metasurfaces for complete control of phase and polarization with subwavelength spatial resolution and high transmission,” Nat. Nanotechnol. 10, 937–943 (2015). [CrossRef]   [PubMed]  

27. V. Lousse, W. Suh, O. Kilic, S. Kim, O. Solgaard, and S. Fan, “Angular and polarization properties of a photonic crystal slab mirror,” Opt. Express 12, 1575–1582 (2004). [CrossRef]   [PubMed]  

28. H. Tanji-Suzuki, W. Chen, R. Landig, J. Simon, and V. Vuletić, “Vacuum-induced transparency,” Science 333, 1266 (2011). [CrossRef]   [PubMed]  

29. J. G. Bohnet, Z. Chen, J. M. Weiner, D. Meiser, M. J. Holland, and J. K. Thompson, “A steady-state superradiant laser with less than one intracavity photon,” Nature 484, 79 (2012). [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 Schematic drawing of the cross sections of an ARROW: (a) lengthwise; transverse cross section of (b) a conventional hollow-core self-aligned pedestal ARROW that has a high index layer beneath the core (n1 > n2) and (c) a hollow-core self-aligned pedestal ARROW that has a low index layer beneath the core; (d) conceptual vision of an integrated opto-electronic system combining an ARROW loaded with atomic or molecular vapor with a Bragg grating and micro-wires that could be used to, e.g., generate magnetic field.
Fig. 2
Fig. 2 Numerical simulation (a) of the transverse profile of the fundamental waveguide mode and (b) of the propagation loss spectrum of a hollow-core ARROW with a 5.8µm × 12µm core surrounded by 3 pairs of Si3N4(1.74)/SiO2 cladding layers arranged as shown in Fig. 1(b), designed and optimized for a wavelength of 852 nm. (c) Numerical simulation of propagation losses in waveguides optimized for 852 nm light with assorted hollow core sizes and three pairs of of Si3N4(1.74)/SiO2 cladding layers as shown in Fig. 1(b). (d) Effects of the number of cladding layer pairs and of the refractive index contrast between the two materials forming each pair on the simulated propagation losses for waveguides with 5.8µm × 12µm hollow cores. The square data points correspond to the cladding layers arranged as shown in Fig. 1(b), while the triangular data points represent cladding layers arranged as shown in Fig. 1(c). The value in brackets denotes the refractive index of the cladding layer material.
Fig. 3
Fig. 3 (a) Etched gratings in the various anti-resonant layers: first layer above the core (left), first layer beneath the core (middle), and the top layer (right). (b) Change in the effective index of propagation Δneff for a hollow-core ARROW with 5.8µm × 12µm core surrounded by three pairs of Si3N4(1.74)/SiO2 cladding layers. The x-axis refers to the etch depth δt relative to the original thickness t of that particular layer. When etching the very top layer, at etch depths shown to induce uncharacteristically large effective propagation index contrasts and losses, the mode is no longer quasi-Gaussian within the core. (c) Estimated best mirror reflectivities of the Bragg gratings formed by etching the three different cladding layers obtained using MSE with plane wave incidence. The dash-dot horizontal lines mark the maximum reflectivity for grating in a particular layer. (d) Estimated reflectivity spectrum of a 10000 period Bragg mirror created by periodically etching the anti-resonant layer beneath the core. The etch depth is 70% of this layer, which creates an effective propagation index contrast of 5 × 10−4 with a 8% increase in loss.
Fig. 4
Fig. 4 (a) Schematic showing how the dielectric metasurface mirror would be integrated with the hollow-core ARROW waveguide. (b) Cooperativity of hollow-core ARROW (Amode = 18µm2) cavity formed with two dielectric metasurface mirrors for various waveguide lengths and mirror reflectivities.

Equations (13)

Equations on this page are rendered with MathJax. Learn more.

t i = λ 4 n i ( 2 N + 1 ) ( 1 n c 2 n i 2 + λ 2 4 n i 2 d c 2 ) 1 2
α v e r t i c a l = 2 R t o p R b o t t o m 2 h c tan ( θ c , v ) α h o r i z o n t a l = 2 R l e f t R r i g h t 2 w c tan ( θ c , h )
A m o d e = ϵ ( x , y ) | E ( x , y ) | 2 d x d y max { ϵ ( x , y ) | E ( x , y ) | 2 }
d U ( z ) d ( k 0 z ) = Y ( z )
d Y ( z ) d ( k 0 z ) = P 2 ( z ) U 3 ( z ) R e { ϵ ( z ) } U ( z )
d P ( z ) d ( k 0 z ) = I m { ϵ ( z ) } U 2 ( z ) ,
R = | U 2 ( 0 ) ϵ i P ( 0 ) j U ( 0 ) Y ( 0 ) U 2 ( 0 ) ϵ i + P ( 0 ) + j U ( 0 ) Y ( 0 ) | 2
g = μ ω 2 ϵ M V m o d e .
V m o d e = V ϵ | E | 2 d V max { ϵ | E | 2 } = A m o d e L e f f
g 2 κ γ = 3 16 π 2 Q V m o d e / λ 3
Q = ω n L c ( ln R α L )
g 2 κ γ 3 16 π λ A m o d e Δ n e f f α w g
g 2 κ γ = 3 8 π λ 2 A m o d e 1 l n R α L e f f ,
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.