You are on page 1of 8

High-Q nanobeam cavities on a silicon

nitride platform enabled by slow light


Cite as: APL Photonics 5, 066101 (2020); https://doi.org/10.1063/5.0007279
Submitted: 10 March 2020 . Accepted: 04 May 2020 . Published Online: 02 June 2020

Jiahao Zhan, Zeinab Jafari, Sylvain Veilleux, Mario Dagenais, and Israel De Leon

COLLECTIONS

This paper was selected as Featured

ARTICLES YOU MAY BE INTERESTED IN

Optical multi-stability in a nonlinear high-order microring resonator filter


APL Photonics 5, 056106 (2020); https://doi.org/10.1063/5.0002941

High quality factor photonic crystal nanobeam cavities


Applied Physics Letters 94, 121106 (2009); https://doi.org/10.1063/1.3107263

Optical parametric gain in CMOS-compatible sub-100 m photonic crystal waveguides


APL Photonics 5, 066108 (2020); https://doi.org/10.1063/5.0003633

APL Photonics 5, 066101 (2020); https://doi.org/10.1063/5.0007279 5, 066101

© 2020 Author(s).
APL Photonics ARTICLE scitation.org/journal/app

High-Q nanobeam cavities on a silicon nitride


platform enabled by slow light
Cite as: APL Photon. 5, 066101 (2020); doi: 10.1063/5.0007279
Submitted: 10 March 2020 • Accepted: 4 May 2020 •
Published Online: 2 June 2020

Jiahao Zhan,1 Zeinab Jafari,2 Sylvain Veilleux,3 Mario Dagenais,1,a) and Israel De Leon2,b)

AFFILIATIONS
1
Department of Electrical and Computer Engineering, University of Maryland, College Park, Maryland 20742, USA
2
School of Engineering and Sciences, Tecnológico de Monterrey, Monterrey, Nuevo Leon 64849, Mexico
3
Department of Astronomy, University of Maryland, College Park, Maryland 20742, USA

a)
Author to whom correspondence should be addressed: dage@ece.umd.edu
b)
Electronic mail: ideleon@tec.mx

ABSTRACT
Silicon nitride integrated photonic devices benefit from a wide working spectral range covering the visible and near-infrared spectra, which
in turn enables important applications in bio-photonics, optical communications, and sensing. High-quality factor optical resonators are
essential photonic devices for such applications. However, implementing such resonators on a silicon nitride platform is quite challenging
due to the low refractive index contrast attainable with this material. Here, we demonstrate that silicon nitride photonic cavities compris-
ing a slow-light waveguide bounded by mirrors can in principle exhibit quality factors in the order of several millions despite a relatively
low refractive index contrast. We show that the energy stored in such a slow-light cavity exhibits a cubic dependence on the cavity length,
which can enable extremely large quality factors with modest-length cavities. We present the design and experimental characterization of
silicon nitride slow-light nanobeam-type cavities. Two sets of nanobeam cavities were fabricated to experimentally verify the cubic depen-
dence of the Q factor on the cavity length. The highest measured Q factor in our devices is 4.42 × 105 , which is limited by fabrication
imperfections.
© 2020 Author(s). All article content, except where otherwise noted, is licensed under a Creative Commons Attribution (CC BY) license
(http://creativecommons.org/licenses/by/4.0/). https://doi.org/10.1063/5.0007279., s

I. INTRODUCTION Q factor, most SiN nanobeam cavity designs to date have resorted
to using suspended (air-cladded) architectures, reporting measured
Optical cavities with a high quality factor (Q factor) are of Q factors in the range of 104 –105 .15–17 However, the increased Q
great importance in various applications, such as narrow-band filter- factors come at the cost of a more complex fabrication process
ing, ultra-low energy switching and modulation, sensing, and cavity and structural fragility. A recent effort to tackle this problem has
quantum electrodynamics.1–5 In particular, nanobeam cavities (one- reported an encapsulated SiN nanobeam cavity design with theoret-
dimensional photonic crystal cavities) have attracted much atten- ical and measured Q factors of ∼105 and 7000, respectively.12 While
tion because they can achieve large Q factors in addition to offer- this represents important progress, more work is still required to
ing design flexibility and relatively simple fabrication.6,7 Recently, achieve Q factors comparable to those of suspended SiN or silicon
silicon nitride (SiN) nanobeam cavities have been proposed, bring- nanobeam cavities while maintaining the device’s ease of fabrication
ing advantages over their silicon counterparts such as a broader and structural robustness.
operation range covering the visible and the near-infrared spectra In this work, we use the concept of slow light to show, both
and the absence of two-photon absorption in the telecommunica- theoretically and experimentally, significant enhancements in the
tion band.8–10 Nonetheless, the low refractive index of this material Q factor of SiN cavities. Slow light refers to optical modes with a
poses challenges to achieving high Q factors because of the lower small group velocity. There are two standard mechanisms to gen-
index-contrast attainable compared to silicon.11–14 To improve their erate slow light, both of which require spectral regions of high

APL Photon. 5, 066101 (2020); doi: 10.1063/5.0007279 5, 066101-1


© Author(s) 2020
APL Photonics ARTICLE scitation.org/journal/app

dispersion. One of them uses material dispersion such as that created found as
near an atomic resonance, while the other uses structural dispersion U UC κ 2 3
such as that created in photonic crystals.18 Slow-light waveguides of Qm = ωm = ωm [L + ( ) L ], (2)
P P mπ
both types have been used in the past to increase the photon lifetime
in various cavities,19–24 showing that the Q factor increases with the where UC = U0 /L = 12 ε0 n20 E02 A is a constant and P is the power dis-
group index, ng . However, despite its potential, this approach has sipation. Assuming that the loss of the waveguide is small and hence
not been thoroughly investigated, thus far overseeing aspects such considering a constant P, the Q factor scales with the cube of the cav-
as the effect of the cavity length on the cavity performance. Here, we ity length L. Equation (2) clearly shows that it is possible to enhance
show experimentally that nanobeam cavities incorporating a slow- the Q factor by modestly increasing the length of the cavity. Note
light photonic crystal waveguide exhibit a Q factor that increases that the L3 dependence of Q is not expected for slow-light cavities
with the cube of the cavity length (as opposed to the characteris- based on slow light resulting from material dispersion.19
tic linear dependence of standard cavities25 ), enabling the design of Although the analysis leading to Eq. (2) considers a small sinu-
high-Q SiN nanobeam cavities with modest footprints despite a low soidal index variation, a similar result is expected for more compli-
refractive index contrast. Using this approach, we report experimen- cated periodic patterns. To verify the validity of Eq. (2) for other
tal measurements of Q factors as large as 4.42 × 105 and theoreti- cavities, we perform 2.5D finite-difference time domain (FDTD)
cal estimations of several millions for glass-cladded SiN nanobeam simulations (Lumerical) on three slow-light waveguides with differ-
cavities. These values are the largest Q factors reported to date for ent group indices (ng = 24, ng = 14, and ng = 7) and examine the
non-suspended SiN nanobeam cavities, and are comparable to those effects of ng and L on the Q factor. The designed slow-light waveg-
reported in suspended nanobeam cavities.15–17 Furthermore, our uides are made of SiN strip waveguides periodically patterned with
numerical simulations indicate that the Purcell factor of these slow- elliptical holes spaced by a distance of Λ = 515 nm. SiN is modeled by
light cavities increases quadratically with the cavity length despite the Sellmeier equation given in Ref. 27 to take into account the mate-
the increased mode volume, which could be useful for applications rial dispersion. Since the material absorption of our low-pressure
involving molecule sensing and strong interaction with quantum chemical vapor deposition (LPCVD) SiN is vanishingly small, it is
emitters. neglected in the simulation. The elliptical shape of the holes results
The remainder of the paper is organized as follows. In Sec. II, in a wider stopband and a larger mirror reflectivity compared to
the principles of operation, cavity design, and simulation results are the circular one.28 The cross section of the waveguides is 1200 ×
discussed. In Secs. III and IV, the experimental results are presented 300 nm2 , which is chosen to support only the fundamental trans-
and discussed, respectively. In Sec. V, the major conclusions are verse modes (TE and TM) maintaining at the same time a good
summarized. mode confinement. The entire structure is encapsulated in SiO2 ,
having substrate and upper-cladding thicknesses of 10 μm and 4 μm,
respectively.
II. PRINCIPLE OF OPERATION AND CAVITY DESIGN The band structures of the three slow-light waveguide designs
One of the techniques used to increase the Q factor in optical are plotted in Fig. 1(a). Only the first two bands corresponding to
cavities is based on utilizing a slow-light medium inside the cavi- the dielectric mode (solid line) and the air mode (dashed line) are
ties.19–24 The enhanced photon lifetime in the slow-light medium, shown. For the dielectric mode, most of the mode power is confined
which is a consequence of an increased ng and a reduced group in the high index material. Thus, larger Q factors could be obtained
velocity, augments the Q factor. In a structural slow-light waveg- for this mode in comparison to the air mode for which most of the
uide, it is feasible to increase ng by changing the structural param- power is in the low index material. To define a cavity, we simply
eters of the waveguide and adjusting the slope of the band at the truncate the number of holes in the slow-light waveguides, N W , as
wavelength of interest. Nonetheless, as we discuss below, the Q fac- shown in the inset of Fig. 1(b). This acts as a uniform-reflectivity
tor of a slow-light cavity (slow-light waveguide bounded by mir- mirror because the effective indices of the waveguide mode do not
rors) also depends critically on the cavity length (L). To obtain an vary much over the wavelength range of interest. As can be inferred
understanding of the basic physical mechanism leading to the Q- from Fig. 1(a), the first band (solid line) of the three waveguides is
factor enhancement in such a cavity, we first consider a simple case far away from the light cone (gray region), meaning that the radi-
of a 1D waveguide with periodic index variation in the form of ation loss to the light cone is negligible for all the cavities. Thus,
n(x) = n0 + n1 cos(2πx/Λ), where n1 is small compared to n0 . The the contribution of the radiation loss and mirror reflection to the
stored energy in such a structure is given by26 Q factor is decoupled from that of ng . Figure 1(b) illustrates how
the Q factor of the fundamental TE mode in each of these three
κ2 /γL cosh γL sinh γL − (Ω/ν)2 cavities is enhanced as N W is increased. As expected from our previ-
U = U0 [ ], (1) ous analysis leading to Eq. (2), the simulation results fit the form
γ2 cosh2 γL + (Ω/ν)2 sinh2 γL Q = aNW 3
+ bNW very well (note that L scales linearly with N W
as L = ΛN W ; here we have replaced L with N W ). It is also worth
where U0 = 12 ε0 n20 E02 AL is the energy stored in an unstructured noting that the larger the ng is, the larger the Q factor is for a
waveguide of the same length, L is the length of the waveguide, constant L.
ν = c/n0 is the phase velocity of light in the background material, For further analysis, we use the slow-light waveguide with the
κ is the coupling strength, Ω √ = ω − ω0 is the detuning from the largest group index (ng = 24 obtained for the hole radii of 500 nm
Bragg frequency, and γ = κ2 − (Ω/ν)2 . After some brief alge- and 110 nm) in the remaining of this paper. To achieve high Q fac-
braic calculations, the quality factor at resonances where γL = imπ is tors, the slow-light waveguide (with N W holes) is bounded by Bragg

APL Photon. 5, 066101 (2020); doi: 10.1063/5.0007279 5, 066101-2


© Author(s) 2020
APL Photonics ARTICLE scitation.org/journal/app

FIG. 1. (a) Band structures of one unit cell in various slow-light waveguides with ng
= 24 obtained for the hole radii of 500 nm and 110 nm, ng = 14 obtained for the hole
radii of 200 nm and 110 nm, and ng = 7 obtained for the hole radii of 100 nm and
110 nm. The solid (dashed) lines refer to the dielectric (air) mode, the gray regions
FIG. 2. (a) Top view of the proposed nanobeam cavity made of a slow-light waveg-
show the light cone, and the red dots illustrate the operating frequencies of the
uide with NW holes between two sets of taper and mirror with NT and NM holes,
cavities made of these slow-light waveguides. (b) The Q factor of the fundamental
respectively. (b) Magnitude of the magnetic field |H| of the fundamental resonant
resonant mode (TE polarization) vs NW in the three cavities made of the slow-light
mode in cavities with NW of 40, 120, and 160, respectively. The scale bar rep-
waveguides with different ng . The inset shows the geometry of the cavities.
resents 5 μm. (c) The Q factor and effective mode area vs NW . Red dots are
the Q factor found by FDTD simulations and the blue curve shows the fitting to
3
Q = aNW + bNW .
mirrors on both sides (with N M holes each) to form a nanobeam
cavity, as depicted in Fig. 2(a). Two tapers (with N T holes each) are
used between the slow-light waveguide and the mirrors to adiabati- calculated by
cally match their mode profiles and minimize scattering losses.6 The ( ∣H∣2 dA)2
minor and major radii of the holes in the mirror sections are 60 nm A= ∫ , (3)
∫ ∣H∣ dA
4
and 200 nm, respectively. The hole parameters in the taper sections
are adjusted to create a quadratic tapering of the filling factor and an where H is the magnetic field vector.29 Figure 2(c) plots the value of
exponential decay of the fields along the tapers.6 The magnetic field A as a function of N W as obtained from FDTD simulations. Contrary
distributions of the fundamental TE resonant mode in the proposed to the cubic scaling of the Q factor, the effective mode area follows
cavity for three different N W values of 40, 120, and 160 are plotted a linear trend. Note that the same linear trend is expected for the
in Fig. 2(b). These field profiles are given for a constant number of mode volume, V. Hence, the Purcell factor of the cavity, Pf ∝ Q/V,
holes in the taper and mirror sections (N T = 80 and N M = 60). The exhibits a quadratic dependence on the cavity length.
Q factor vs N W for the same number of holes in the taper and mir-
ror sections is also depicted in Fig. 2(c). The Q factor varies, again,
with the cube of L and a high Q factor of ∼1.7 × 107 is achieved at III. EXPERIMENTAL RESULTS
a wavelength of ∼1627 nm for N W = 200. It should be noted that We fabricated a set of devices of different parameters (N W ,
even larger Q factors are predicted by adding more holes to the mir- N T , and N M ) using the structural specification detailed in Sec. II.
ror sections (For instance, a Q factor of ∼2.8 × 107 is obtained for Details of the fabrication procedure are given in Sec. VI. Since the
N M = 70). nanobeam cavities are designed for TE polarization, when charac-
Finally, we estimate how the effective area of the mode terizing their performances, excess TM polarization is not desired.
in the (x, z) plane, A, scales with the cavity length and is To eliminate unwanted TM light caused by our laser source and

APL Photon. 5, 066101 (2020); doi: 10.1063/5.0007279 5, 066101-3


© Author(s) 2020
APL Photonics ARTICLE scitation.org/journal/app

how the residual polarization can affect the transmission spectrum).


For the TM polarization filter, we chose periodical circular holes as a
grating, and by adjusting its period, we can align its stopband around
the edge of the TE bandgap of the nanobeam cavity, where resonant
peaks are located. Scanning-electron micrographs of a section of a
TM filter and a section of a nanobeam cavity are shown in Figs. 3(a)
and 3(b), respectively.
The measurement of fabricated devices was performed using
the experimental setup as schematically drawn in Fig. 3(c). We
use a tunable laser source operating in the 1450–1640 nm wave-
length range and a power meter. The fiber rotator in Fig. 3(c) con-
trols the input polarization. A butt-coupling technique is applied
to couple light into the chip from a polarization-maintaining fiber,
and the input waveguide width is optimized for the maximum
coupling efficiency.30 An index-matching liquid is also applied to
minimize reflection losses. The offset between the input and out-
put waveguides is to avoid stray light coupling into the output
fiber.
Figure 4 plots the measured transmission spectra of a
nanobeam cavity with N W = 200, N T = 80, and N M = 60 for both
FIG. 3. (a) SEM image of a segment of a TM polarization filter. (b) SEM image of polarizations, respectively. The blue and red curves correspond to
a segment of a slow-light waveguide. (c) Schematic of the experimental setup to the measurement results obtained from the coarse and fine sweep-
characterize the fabricated devices. ing of the tunable laser, respectively. The two shaded regions mark
the locations of the stopband of the designed TM polarization filter.
Clearly, TE resonant peaks are revealed with a large contrast of at
polarization optics, an on-chip TM filter was also implemented in least 50 dB, due to the proper alignment of the TM filter stopband
addition to the nanobeam cavities. This allowed us to obtain a large with those TE resonant peaks.
TE/TM ratio of at least 50 dB required to observe the spectra of the It is worth mentioning that although the nanobeam cavities
TE resonant modes under study (see the supplementary material for are originally designed for TE polarization, they do have some

FIG. 4. (a) Measured transmission spectra of a nanobeam cavity with NW = 200, NT = 80, and NM = 60 for TE polarization. The blue and red curves correspond to the results
obtained from a coarse sweeping of the tunable laser with a 2 pm step size and a fine sweeping with a 0.2 pm step size, respectively. (b) Zoom-in of the resonant peak
indicated by the black arrow in (a), with the measured linewidth. (c) Measured transmission spectra of the same cavity for TM polarization. (d) Zoom-in of the resonant peak
indicated by the black arrow in (c), with the measured linewidth.

APL Photon. 5, 066101 (2020); doi: 10.1063/5.0007279 5, 066101-4


© Author(s) 2020
APL Photonics ARTICLE scitation.org/journal/app

TABLE I. List of parameters and highest measured Q factors of fabricated nanobeam


cavities.

Geometrical parameters Highest measured Q


NW NT NM TE TM

120 60 50 2.79 × 105 8.16 × 104


160 60 80 3.79 × 105 1.46 × 105
200 60 80 4.09 × 105 1.56 × 105
200 80 60 4.42 × 105 1.88 × 105
200 80 70 3.19 × 105 2.04 × 105

TM resonances as well, as indicated by the peaks appearing in the


1540–1560 nm wavelength range. However, we note that the Q fac-
tors are lower for TM polarization because the effective index of
the TM-polarized mode of the cavity is smaller than that for the TE
polarization. FIG. 5. (a) Comparison of the fitted curve (blue) and the measured Q factors (red
For TE polarization, the resonant peak at 1604.7 nm has dots) for TE polarization, with NM = NT = 15. (b) Comparison of the fitted curve
a 3-dB linewidth of 3.63 pm, corresponding to a Q factor (blue) and the measured Q factors (red dots) for TM polarization, with NM = NT
of 4.42 × 105 , with a transmission of −48.7 dB. The highest = 25. (c) Zoom-in of the shaded area in (a). The fitted curve is based on these 7
measured Q factor for TM polarization of the same cavity is data points and has an SSE of 1 and an R-squared of 0.9972. The Q values have
1.88 × 105 (with a 3-dB linewidth of 8.18 pm at 1542.8 nm). been scaled down by a factor of 1000 for a better fitting. (d) Zoom-in of the shaded
area in (b). The fitted curve is based on these 10 data points and has an SSE of
The measured Q factors for all fabricated devices are listed in 0.469 and an R-squared of 0.9979. The Q values have been scaled down by a
Table I. factor of 1000 for a better fitting.
To experimentally verify the cubic relation between the Q fac-
tor and cavity length, we fabricated two sets of nanobeam cavities
with (I) N M = N T = 15 and (II) N M = N T = 25 and N W increas-
IV. DISCUSSION
ing from 40 to 220 for both sets. Then, we measured the highest
Q factor of the TE resonant modes in the first set, and the high- The highest measured Q factor in the proposed cavity is
est Q factor of the TM resonant modes in the second set [note that 4.42 × 105 , which was obtained for N W = 200, N T = 80, and
the derivation of Eq. (2) does not require a specific polarization]. N M = 60, and the total cavity length of 247.2 μm. A Q factor as high
The two sets have different N M and N T values in order to keep as 1.29 × 105 was also measured for a smaller number of holes in
the Q factors in a similar regime (<∼2.5 × 104 ) for both polariza- the mirror and taper sections (N W = 200, N T = 15, and N M = 15
tions. From Figs. 5(a) and 5(b), we observe a nonlinear dependence and the total length of 133.9 μm), as shown in Fig. 5(a). To the best
of Q on N W . our knowledge, these are the highest reported Q factors obtained
We find that for both polarizations, the experimental data fol- experimentally for non-suspended SiN nanobeam cavities.11–13 The
low a cubic dependence for a limited range of cavity lengths, as indi- slow-light waveguide inside the proposed cavities affects the Q fac-
cated by the shaded areas in Figs. 5(a) and 5(b). The blue curves in tor in two ways: (I) its group index defines the minimum achievable
the two figures are fits of the cubic expression Q = aNW 3
+ bNW to the Q factor for a zero-length cavity with a fixed number of holes in the
data points in the shaded areas. Figures 5(c) and 5(d) are zoom-ins mirror and taper sections, and (II) its increased length can enhance
of the shaded areas to show the associated data points and fits. Note the Q factor significantly, considering that the Q factor has a cubic
that in this regime of relatively low Q, the data fit (SSE of 1 and 0.469 dependence on L, as opposed to the linear dependence in conven-
and R-squared of 0.9972 and 0.9979) the cubic relation very well. tional cavities.25 It is also in contrast to the nanobeam cavities with
However, we note a clear deviation from the cubic fit for the data out a conventional strip waveguide in between the mirrors in which the
of the shaded areas, corresponding to cavities with N W > 160 and Q maximum attainable Q factor is obtained for a zero-length cavity.6,31
> ∼2.5 × 104 . We attribute this deviation to the effect of power dissi- Furthermore, our analysis indicates that the Purcell factor varies
pation by light scattering, which would inevitably be present in our quadratically with L, suggesting the possibility of enhancing this fig-
devices due to the material roughness and fabrication imperfections. ure by increasing the cavity length. The maximum useful waveguide
As mentioned earlier, power dissipation, P, plays an important role length is limited due to the pronounced scattering loss in slow-light
in defining the Q factor, as seen by the 1/P dependence in Eq. (2). waveguides32,33 which consequently reduces the maximum Q factor
Thus, since power dissipation in our devices is always an increasing in practice. The effect of the loss on the Q factor and the Purcell fac-
function of the cavity length, it is expected that the dependence of tor will be further studied and reported elsewhere. It is also worth
Q on the cavity length deviates from the cubic relationship for long mentioning that the Q factors reported in this work are comparable
cavities. to those previously measured for suspended SiN nanobeam cavities

APL Photon. 5, 066101 (2020); doi: 10.1063/5.0007279 5, 066101-5


© Author(s) 2020
APL Photonics ARTICLE scitation.org/journal/app

TABLE II. Comparison of different SiN cavities.

Structure SiN thickness (nm) L (μm) λ (nm) Q factor References

Sa 200 ∼5.7 ∼630 5.5 × 104 15


S 400 ∼28.2 ∼1500 3.0 × 105 16
S 350 ∼21.6 ∼980 4.0 × 105 17
NSb 220 ∼18.5 ∼630 7.6 × 103 11
NS 330 ∼13.7 ∼780 7.0 × 103 12
NS 220 ∼23.3 ∼730 6.5 × 103 13
NS 300 ∼247.2 ∼1605 4.4 × 105 This work
NS 300 ∼133.9 ∼1605 1.3 × 105 This work
a
Suspended cavities.
b
Non-suspended cavities.

with larger index contrasts.15–17 A comparison of several reported AUTHOR’S CONTRIBUTIONS


SiN nanobeam cavities is summarized in Table II.
J.Z. and Z.J. contributed equally to this work. I.D.L and M.D.
supervised all aspects of the project.
V. CONCLUSION
In this paper, we report the design and experimental character-
ization of high-Q non-suspended nanobeam cavities fabricated on a ACKNOWLEDGMENTS
300-nm-thick SiN platform. The structures consist of 1D slow-light
photonic crystal waveguides (ng ∼ 24) of various lengths bounded We thank the University of Maryland Fablab staff member
by Bragg mirrors. We show that the Q factor of these slow-light Mark Lecates for assistance with the fabrication procedure. J. Zhan,
cavities exhibits a cubic dependence on the cavity length, which M. Dagenais, and S. Veilleux acknowledge the support from the
enables them to achieve large Q factors with modest cavity lengths. National Aeronautics and Space Administration (NASA) through
We report measured Q factors as large as 4.42 × 105 and numeri- Grant No. 16-APRA16-0064. J. Zhan, Z. Jafari, M. Dagenais, and I.
cally predicted values of up to 2.8 × 107 , which are comparable to De Leon were supported by the UMD-TEC Seed Award, Grant No.
those of suspended SiN nanobeam cavities previously reported. In 2-957016 (0020240I10). I. De Leon and Z. Jafari acknowledge the
addition, a TM polarization filter is designed and included to enable financial support from CONACyT Grant No. CN-17-109 and from
the high-contrast detection (>50 dB) of the TE resonant modes, the Federico Baur Endowed Chair in Nanotechnology.
which would otherwise be buried under the transmission of excess
TM polarization. The proposed cavities are mechanically robust and
have a great potential to be integrated with new photonic materials DATA AVAILABILITY
such as 2D materials. Thus, these high-Q cavities are promising for The data that support the findings of this study are available
a wide variety of applications such as narrowband filtering, ultra- from the corresponding author upon reasonable request.
low energy switching and modulation, sensing, and cavity quantum
electrodynamics.
REFERENCES
VI. METHODS
1
P. B. Deotare, L. C. Kogos, I. Bulu, and M. Loncar, “Photonic crystal nanobeam
The devices are fabricated in the following procedure. First, cavities for tunable filter and router applications,” IEEE J. Sel. Top. Quantum
300-nm-thick stoichiometric SiN is deposited via low-pressure Electron. 19, 3600210 (2013).
chemical vapor deposition (LPCVD) on silicon wafers with 10-μm 2
Z. Jafari, A. Zarifkar, M. Miri, and L. Zhang, “All-optical modulation in a
thermal SiO2 on top. Device patterning is performed by e-beam graphene-covered slotted silicon nano-beam cavity,” J. Lightwave Technol. 36,
lithography with a ZEP-520A resist, followed by a developing pro- 4051–4059 (2018).
3
cess. Next, a thin chromium layer is deposited as the hard mask E. Li, Q. Gao, R. T. Chen, and A. X. Wang, “Ultracompact silicon-conductive
oxide nanocavity modulator with 0.02 lambda-cubic active volume,” Nano Lett.
and nanobeam cavity patterns are then transferred into SiN by a
18, 1075–1081 (2018).
CHF3 –SF6 based plasma etching process. The hard mask is removed 4
B. Wang, M. A. Dündar, R. Nötzel, F. Karouta, S. He, and R. W. van der Hei-
with an acid and 4-μm-thick SiO2 cladding is deposited via plasma- jden, “Photonic crystal slot nanobeam slow light waveguides for refractive index
enhanced chemical vapor deposition (PECVD). sensing,” Appl. Phys. Lett. 97, 151105 (2010).
5
M. W. McCutcheon and M. Loncar, “Design of a silicon nitride photonic crys-
SUPPLEMENTARY MATERIAL tal nanocavity with a quality factor of one million for coupling to a diamond
nanocrystal,” Opt. Express 16, 19136 (2008).
See the supplementary material for further details regarding the 6
Q. Quan and M. Loncar, “Deterministic design of wavelength scale, ultra-high Q
TM polarization filter. photonic crystal nanobeam cavities,” Opt. Express 19, 18529 (2011).

APL Photon. 5, 066101 (2020); doi: 10.1063/5.0007279 5, 066101-6


© Author(s) 2020
APL Photonics ARTICLE scitation.org/journal/app

7
P. B. Deotare, M. W. McCutcheon, I. W. Frank, M. Khan, and M. Lončar, “High lifetime by induced slow light and nonlinear dispersions,” Opt. Express 20, 27403
quality factor photonic crystal nanobeam cavities,” Appl. Phys. Lett. 94, 121106 (2012).
(2009). 21
K. McGarvey-Lechable and P. Bianucci, “Maximizing slow-light enhancement
8
D. J. Moss, R. Morandotti, A. L. Gaeta, and M. Lipson, “New CMOS-compatible in one-dimensional photonic crystal ring resonators,” Opt. Express 22, 26032
platforms based on silicon nitride and hydex for nonlinear optics,” Nat. Photonics (2014).
7, 597–607 (2013). 22
K. McGarvey-Lechable, T. Hamidfar, D. Patel, L. Xu, D. V. Plant, and
9
A. Rahim, E. Ryckeboer, A. Z. Subramanian, S. Clemmen, B. Kuyken, A. Dhakal, P. Bianucci, “Slow light in mass-produced, dispersion-engineered photonic crystal
A. Raza, A. Hermans, M. Muneeb, S. Dhoore, Y. Li, U. Dave, P. Bienstman, N. Le ring resonators,” Opt. Express 25, 3916 (2017).
Thomas, G. Roelkens, D. Van Thourhout, P. Helin, S. Severi, X. Rottenberg, and 23
D. Goldring, U. Levy, I. E. Dotan, A. Tsukernik, M. Oksman, I. Rubin, Y. David,
R. Baets, “Expanding the silicon photonics portfolio with silicon nitride photonic and D. Mendlovic, “Experimental measurement of quality factor enhancement
integrated circuits,” J. Lightwave Technol. 35, 639–649 (2017). using slow light modes in one dimensional photonic crystal,” Opt. Express 16,
10
M. A. Porcel, A. Hinojosa, H. Jans, A. Stassen, J. Goyvaerts, D. Geuzebroek, 5585 (2008).
M. Geiselmann, C. Dominguez, and I. Artundo, “Silicon nitride photonic integra- 24
V. Huet, A. Rasoloniaina, P. Guillemé, P. Rochard, P. Féron, M. Mortier, A. Lev-
tion for visible light applications,” Opt. Laser Technol. 112, 299–306 (2019). enson, K. Bencheikh, A. Yacomotti, and Y. Dumeige, “Millisecond photon lifetime
11
Y. Chen, A. Ryou, M. R. Friedfeld, T. Fryett, J. Whitehead, B. M. Cossairt, and in a slow-light microcavity,” Phys. Rev. Lett. 116, 133902 (2016).
A. Majumdar, “Deterministic positioning of colloidal quantum dots on silicon 25
Y.-W. Hu, Y. Zhang, P. Gatkine, J. Bland-Hawthorn, S. Veilleux, and M. Dage-
nitride nanobeam cavities,” Nano Lett. 18, 6404–6410 (2018). nais, “Characterization of low loss waveguides using Bragg gratings,” IEEE J. Sel.
12
T. K. Fryett, Y. Chen, J. Whitehead, Z. M. Peycke, X. Xu, and A. Majumdar, Top. Quantum Electron. 24, 1–8 (2018).
“Encapsulated silicon nitride nanobeam cavity for hybrid nanophotonics,” ACS 26
H. G. Winful, “The meaning of group delay in barrier tunnelling: A re-
Photonics 5, 2176–2181 (2018). examination of superluminal group velocities,” New J. Phys. 8, 101 (2006).
13 27
Y. Chen, J. Whitehead, A. Ryou, J. Zheng, P. Xu, T. Fryett, and A. Majumdar, K. Luke, Y. Okawachi, M. R. E. Lamont, A. L. Gaeta, and M. Lipson, “Broadband
“Large thermal tuning of a polymer-embedded silicon nitride nanobeam cavity,” mid-infrared frequency comb generation in a Si3 N4 microresonator,” Opt. Lett.
Opt. Lett. 44, 3058 (2019). 40, 4823–4826 (2015).
14 28
S. Sergent, M. Takiguchi, H. Taniyama, A. Shinya, E. Kuramochi, and Q. Quan, I. B. Burgess, S. K. Y. Tang, D. L. Floyd, and M. Loncar, “High-Q, low
M. Notomi, “Design of nanowire-induced nanocavities in grooved 1D and 2D SiN index-contrast polymeric photonic crystal nanobeam cavities,” Opt. Express 19,
photonic crystals for the ultra-violet and visible ranges,” Opt. Express 24, 26792 22191 (2011).
(2016). 29
See https://support.lumerical.com/hc/en-us/articles/360034395374-Calculating-
15
M. Khan, T. Babinec, M. W. McCutcheon, P. Deotare, and M. Lončar, “Fab- the-modal-volume-of-a-cavity-mode for detailed information on how the mode
rication and characterization of high-quality-factor silicon nitride nanobeam volume is calculated in the software.
cavities,” Opt. Lett. 36, 421 (2011). 30
T. Zhu, Y. Hu, P. Gatkine, S. Veilleux, J. Bland-Hawthorn, and M.
16
M. Eichenfield, R. Camacho, J. Chan, K. J. Vahala, and O. Painter, “A picogram- Dagenais, “Ultrabroadband high coupling efficiency fiber-to-waveguide coupler
and nanometre-scale photonic-crystal optomechanical cavity,” Nature 459, 550– using Si3 N4 /SiO2 waveguides on silicon,” IEEE Photonics J. 8, 1–12 (2016).
555 (2009). 31
Q. Quan, P. B. Deotare, and M. Loncar, “Photonic crystal nanobeam cav-
17
K. E. Grutter, M. Davanco, and K. Srinivasan, “Si3 N4 nanobeam optomechani- ity strongly coupled to the feeding waveguide,” Appl. Phys. Lett. 96, 203102
cal crystals,” IEEE J. Sel. Top. Quantum Electron. 21, 61–71 (2015). (2010).
18 32
R. W. Boyd, “Material slow light and structural slow light: Similarities and J. Li, L. O’Faolain, I. H. Rey, and T. F. Krauss, “Four-wave mixing in photonic
differences for nonlinear optics,” J. Opt. Soc. Am. B 28, A38 (2011). crystal waveguides: Slow light enhancement and limitations,” Opt. Express 19,
19
M. Soljačić, E. Lidorikis, L. V. Hau, and J. D. Joannopoulos, “Enhancement of 4458 (2011).
microcavity lifetimes using highly dispersive materials,” Phys. Rev. E 71, 026602 33
L. O’Faolain, S. A. Schulz, D. M. Beggs, T. P. White, M. Spasenović,
(2005). L. Kuipers, F. Morichetti, A. Melloni, S. Mazoyer, J. P. Hugonin, P. Lalanne, and
20
P. Grinberg, K. Bencheikh, M. Brunstein, A. M. Yacomotti, Y. Dumeige, T. F. Krauss, “Loss engineered slow light waveguides,” Opt. Express 18, 27627
I. Sagnes, F. Raineri, L. Bigot, and J. A. Levenson, “Enhancement of a nano cavity (2010).

APL Photon. 5, 066101 (2020); doi: 10.1063/5.0007279 5, 066101-7


© Author(s) 2020

You might also like