ArticlePDF Available

Towards Visible Frequency Comb Generation Using a Hollow WGM Resonator

Authors:

Abstract and Figures

Optical frequency combs are widely used for metrology, optical clocks, optical communications and sensors. In whispering gallery mode (WGM) microcavity resonators, geometrical features enable the four-wave mixing phase match condition to be satisfied. Hence, frequency comb generation is achievable. Geometrical dispersion of hollow structure WGM cavities can broaden the comb span to the visible range.
Content may be subject to copyright.
Towards Visible Frequency Comb Generation using a Hollow WGM Resonator
Sho Kasumie, Yong Yang, Jonathan M. Ward, and S´ıle Nic Chormaic
Light-Matter Interactions Unit, Okinawa Institute of Science and
Technology Graduate University, Onna, Okinawa 904-0495, Japan.
(Dated: February 27, 2018)
Optical frequency combs are widely used for metrology, optical clocks, optical communications
and sensors. In whispering gallery mode (WGM) microcavity resonators, geometrical features enable
the four-wave mixing phase match condition to be satisfied. Hence, frequency comb generation is
achievable. Geometrical dispersion of hollow structure WGM cavities can broaden the comb span
to the visible range.
Keywords: Frequency comb, Four-wave mixing, Microbubble optical resonator
I. Introduction
The ability to confine light in microscopic circular or-
bits by a process of continuous total internal reflection
has pushed optical resonators towards the extremes of
optical quality [1], finesse [2] and mode volume [3]. These
so-called whispering gallery resonators (WGRs) [4, 5] and
their morphological dependent resonances have found ap-
plications from ultralow threshold nonlinear optics [6, 7]
and lasing [8] to bio/chemical sensing [9, 10], telecommu-
nications [11, 12], quantum optics [13] and optomechanics
[14, 15]. Apart from technical applications, these devices
have proven to be a test bed for fundamental physics
such as chaos theory [16] and quantum mechanics [17].
Frequency comb generation is arguably one of the most
important applications of WGRs. Frequency combs gen-
erated by phase-locked pulsed lasers produce equidistant
optical lines (or teeth) that can span from the UV to the
IR [18, 19]. The teeth act as an optical ruler and have
greatly advanced the field of optical metrology [20, 21].
Currently, optical combs require large lasers that con-
sume a lot of power, so significant effort is being made to
develop robust chip scale devices - WGRs are at the fore-
front of this development. In a silica WGR, a four-wave
mixing (FWM) process is employed to produce the fre-
quency comb [22, 23] where the signal and idler photons
are generated by the pump beam both of which are reso-
nant with WGMs of the cavity. This is particularly chal-
lenging because the free spectral range of a WGR is not
constant. In 2007, a pioneering approach for WGR comb
generation was proposed and experimentally confirmed
[23]. It was shown that a strong pump laser can intro-
duce phase modulations which shift resonance modes so
that the total dispersion of the cavity is in the anomalous
regime, thus making the FWM process possible [23].
Today it is possible to have on-chip WGRs, that use
continuous low power pump lasers to produce stable, oc-
tave spanning IR frequency combs [24]. However, ex-
tending these devices into the visible and UV regions is
technically challenging due to limitations in engineering
the dispersion in the resonator. While the material dis-
persion is fixed, geometrical dispersion can be modified
so that the total dispersion meets the required condition.
In such a context, WGRs with a hollow structure and
bottle shape were found to have additional degrees of
freedom in their design, namely the wall thickness and
curvature, which improves their geometrical dispersion
manipulability [25–28]. Here, we will discuss how this
dispersion engineering has been, to date, implemented in
hollow WGRs to create IR and visible combs.
II. Dispersion Management in Microbubble
Resonators
To achieve efficient FWM needed for generating a fre-
quency comb, a phase matching condition is required.
For WGMs, the mode structure is affected by both the
material and the geometric dispersion. In total, the dis-
persion of a WGR can be described as the variation of the
free-spectral-range (∆ωF SR ) and is categorized as normal
(∆ωF SR <0) or anomalous (∆ωF SR >0). The disper-
sion compensated regime is when the FSR is independent
of wavelength (i.e. ωF S R < γ and γis the linewidth of
the cavity mode). In one-dimensional resonators, such as
a Fabry-P´erot cavity, the total dispersion is determined
by the material which usually has normal dispersion. By
introducing confinement, the total dispersion can be flat-
tened yet still remains in the normal dispersion regime
due to the overlap with air [29]. In WGRs, the con-
finement of the optical modes and the geometry of the
resonator can be modified in three-dimensions. Chang-
ing the geometry of the WGR, and therefore the physical
path of the WGMs, one can modify the effective refrac-
tive index of the modes (this is also true for different
mode orders) and this is the origin of the geometric dis-
persion. Therefore, by carefully controlling the size of
the WGR, the material dispersion can be controlled. For
a microsphere whose with a diameter larger than 150
µm, at a wavelength of 1.55 µm, the resonator is in the
anomalous dispersion regime. With introduction of the
Kerr effect, by pumping with an appropriate laser power,
the total dispersion can be compensated and degenerate
FWM can be achieved [29, 30]. For microtoriods, disper-
sion compensation is easier to control by varying the mi-
nor diameter of the toroid [23]. Changing the diameter of
the WGR only results in a slight change in the total dis-
persion. By altering the material [31] or choosing higher
2
order radial and azimuthal modes in the WGR, the dis-
persion can also be slightly managed. This method was
used for achieving mid-IR comb generation in fluoride
WGRs [32].
In order to expand comb generation, the dispersion
should be more freely controlled, i.e. both the zero dis-
persion wavelength (ZDW) and the flatness of the dis-
persion curve need to be managed. According to the
principle mentioned above, a synthetic dimension needs
to be introduced to the 3D WGR. To-date there are three
ways reported: (i) Change the lateral profile of the WGR
and excite 3D confined bottle-like modes. The earliest
experimental demonstration was by Savchenkov et al.,
who observed 790 nm comb generation by exciting the
lateral modes in a CaF2cylinder [33]; (ii) Use a wedged
silica microdisk [34]. Microdisks with multiple wedges
were fabricated, so that the shorter wavelength modes
and longer wavelength modes occupied different space
at the wedges, thus experiencing different effective re-
fractive indices; (iii) Make a hollow structure such as a
microbubble. Microbubbles can be fabricated using sev-
eral techniques [27, 35, 36]. For example, in our earlier
work [37], two counter-propagating CO2laser beams were
used. The main point is to heat a small section of silica
capillary while pressurizing the air inside the capillary.
When the glass becomes soft, the air pressure pushes
out the wall of the capillary to form a bubble shape. In
order to get ultrathin walls and the intended bubble di-
ameter/shape, the capillary is first tapered using a CO2
laser and mechanical pulling stages.
It was first proposed in 2013 [26] that microbubbles
could be used for comb generation. The authors pointed
out that the dispersion can be altered not only by chang-
ing the diameter of the microbubble, but also by varying
the wall thickness or filling the core of the microbubble
with different materials. In this way, the ZDW of the
WGR can be shifted towards the ZDW of silica. They
designed and fabricated a microbubble with diameter of
134 µm and wall thickness of 3-4 µm.
FIG. 1. Degenerate FWM in a dispersion compensated mi-
crobubble resonator. Reproduced with permission from[26].
FIG. 2. Hyperparametric oscillation and IR comb genera-
tion in a microbubble resonator. Reproduced with permission
from [38].
With a pumping power of 3 mW, degenerated FWM
occurred at 1.55 µm, see Fig. 1. In comparison, degen-
erated FWM could not be excited in a microsphere of
the same diameter, which proved their claim. This work
showed that the wall thickness of the microbubble pro-
vides a way to control the dispersion. Later, in 2015,
Farnesi et al. fabricated a silica microbubble with diam-
eter of 475 µm and wall thickness of 3-4 µm [38]. By
increasing the pumping power up to 80 mW, they suc-
ceeded to obtain Type I and Type II frequency combs at
a wavelength of 1.56 µm [38], see Fig. 2. Hyperparamet-
ric oscillation incorporating FWM, a Raman process and
other nonlinear effects provided a much broader band-
width comb than the previous work [26]. These two works
show the capability of silica microbubbles for IR comb
generation. For comb generation in the visible spectrum,
the wall thickness needs to be reduced further so that the
ZDW of the microbubble can be shifted to even shorter
wavelengths, as will be discussed in Section IV.
III. Comb Generation in Bottle-Like
Microresonators
While the modes in conventional WGRs are gener-
ally confined to the equator, bottle-like microresonators
(BLMRs) are unique as the optical modes expand in two
dimensions along the surface and have been termed whis-
pering gallery etalons or bottle resonators (Fig. 3). The
two dimensional distribution of the optical mode also cre-
ates an equally spaced FSR. Consider a BLMR with a
parabolic curvature profile along the z-axis of its cylindri-
cal coordinate. The diameter is R(z) = R0[1 (ζz)2/2],
where R0is the maximum radius at z= 0 and ζis the
curvature. The eigenfrequency of the bottle modes is
3
FIG. 3. Illustration of a bottle mode. Reproduced with per-
mission from [25].
[25]:
νj,m =c
2sj2
R2
0
+2ζj
R0
(m+1
2) (1)
with jand mrepresenting its longitudinal and axial mode
numbers, cand nrepresent the speed of light in vacuum
and the refractive index in the material, respectively. In a
practical system, with very small curvature, the variation
of the FSR is:
νG
F SR c[ζR0(m+ 1/2)]2
2πnR0j3(2)
The material dispersion is ∆ν0
F SR c2λ2/4π2n3R2
GV D and GV D =(λ/c)(2n/∂λ2) with λrepresenting
the wavelength.
FIG. 4. Total dispersion of a BLM in terms of the FSR varia-
tion for different working wavelengths. Reproduced with per-
mission from [28]
The dispersion curves for bottle modes and WGMs in
BLMRs of radii 51 µm and 75 µm are plotted in FIG. 4
which shows that the dispersion is always in the anoma-
lous regime around 1.55 µm regardless of the bottle’s
diameter [28]. It can also be seen that, for the bottle
modes, geometric dispersion does not shift the ZDW sig-
nificantly, while the dispersion for the WGM shifts the
ZDW depending on the radius of the BLMR. In a hollow
WGR, the geometric dispersion can be managed by de-
signing the diameter, curvature and wall thickness. The
implication for this is that for a conventional WGM at
1.55 µm it is not possible to create a comb in a resonator
with a diameter below 150 µm due to the normal dis-
persion at small diameters. However, for bottle modes,
anomalous dispersion can be maintained at smaller diam-
eters due to the additional geometric dispersion of these
modes.
To test this hypothesis in experiment, a BLMR with a
diameter of 102 µm and a stem diameter 90 µm was fab-
ricated. The wall thickness was estimated to be 3 4µm
and did not play any role in this case. Light was coupled
to the BLMR via a tapered fibre that was mounted so
that its position along the z-axis could be finely adjusted.
Of course, not only the third nonlinearity FWM is ex-
cited, Raman scattering can also generate a frequency
comb [38]. To exclude the Raman process, the laser
power was kept above the FWM threshold but below the
Raman lasing threshold. Frequency comb generation by
FWM was confirmed by judicious selection of the fibre
position, i.e. when a WGM was pumped no FWM was
generated; however, when the tapered fiber position was
moved to pump a bottle mode then FWM was observed
(Fig. 5). Thus, it was shown that the dispersion can be
engineered in bottle modes to lift the restrictions imposed
by conventional WGMs. As stated, the wall thickness did
FIG. 5. FWM in a hollow BLMR when the taper is positioned
25 µmm away from the center (dashed line). Reproduced with
permission from [28]
not play a role in this experiment; however, combining
the effect of the parabolic curvature with a thin wall re-
mains to be explored as an additional means of dispersion
tuning.
4
IV. Visible Comb Generation in Microbubble
Resonators
To-date, most visible comb generation in WGRs is
done indirectly. An IR comb is first generated through
FWM and then is mapped to the visible range through
second [39] and third harmonic generation [40]. This was
demonstrated in a Si3N4microring by exploiting the high
nonlinearity of the material. The low efficiency of the
higher order harmonic generation process results in lower
quality combs, compared to direct IR Kerr combs. Thus,
combs generated directly by visible frequency FWM is
still being studied. As discussed in the introduction, and
the previous section, direct Kerr frequency comb genera-
tion in the visible range from silica microspheres, toroids
or disks is almost impossible due to the material disper-
sion limit. Better dispersion engineering is required to
go beyond this limitation.
A route to visible combs in hollow microcavities was
recently detailed theoretically [41] for a spherical bub-
ble WGR. The inner diameter, wall thickness and outer
diameter are ρ0,tand ρ1=ρ0+t, respetively. For a
general case, the inner medium, the bubble wall and sur-
rounding medium can all have different refractive indices,
n1,n2and n3, respectively. Then the eigenfrequency is
determined by the characteristic equation:
np
3
χ0
l(z31)
χl(z31)=np
2
Nlψ0
l(z21) + χ0
l(z21)
Nlψl(z21) + χl(z21 )(3)
where zij =kniρj,kis the complex wavenumber and
lis the azimuthal mode number. For simplicity, kis
real k= 2π/λ and only the first order radial mode is
considered. p=±1 is the polarization coefficient where
p= 1 for TE modes and p=1 for TM modes and the
coefficient Nlis:
Nl=np
1ψ0
l(z10)χl(z20 )np
2ψl(z10)χ0
l(z20)
np
1ψl(z10)ψ0
l(z20)np
2ψ0
l(z10)ψl(z20 )(4)
Also, ψl(z) = zJl(z) and χl(z) = zYl(z), where Jland
Ylare the spherical Bessel functions of first and second
kind. The characteristic equation can be used to identify
eigen frequencies. The geometric dispersion is the differ-
ence in resonance frequencies where νl=ckl/2πn2
and ∆νl=νl+1 νl. On the other hand material disper-
sion is related to the Sellmeier equations [42]:
n21 = B1λ2
λ2C2
1
+B2λ2
λ2C2
2
+B3λ2
λ2C2
3
(5)
where nis the refractive index of the material. The
constants are determined by experiments and, for sil-
ica, B1= 0.6961663, B2= 0.4079426, B3= 0.8974794,
C1= 0.0684043, C2= 0.1162414 and C3= 9.896161 [42].
The total dispersion is obtained by summing the geomet-
ric and material dispersion (Fig. 6).The plot shows how
the ZDW can be pushed into the visible range by using
thinner wall and larger bubble diameters. Theoretical
FIG. 6. Dispersion for TM modes in MBR. Reproduced with
permission from [41].
calculations for materials other than silica are available
in [43].
Recently, experimental work confirmed these theoret-
ical predictions by demonstrating, for the first time, a
direct Kerr comb in the visible range [44]. 14 comb
lines were observed around a pump wavelength of 765
nm (Fig.7). The result suggests that comb lines could be
pushed further into the visible range by simply decreasing
the wall thickness and increasing the pump power. The
table summarizes the current state-of-the-art for comb
generation in hollow microresonators.
FIG. 7. Frequency comb generation in a MBR at a center
wavelength of 765 nm. Up to 14 comb lines are excited. Re-
produced with permission from [44].
V. Conclusion
In this brief review, we have discussed recent progress
in the generation of optical frequency combs using
WGMs. The geometrical properties of these devices
enables the required FWM phase matching condition
to be satisfied, leading to frequency comb generation.
To-date, the geometrical dispersion of hollow structure
WGRs has been studied in order to broaden the comb
5
Team Lei Soria Nic Chormaic
Diameter (µm) 136 475 120
Wall Thickness (µm) 3 4 3 4<1.5
Pump Wavelength (nm)1545 1552.4765
Laser Pow (mW ) 3 480 6
Q factor 5 ×107− −
# of comb lines 5 many 14
TABLE I. The current state-of-the-art of frequency comb gen-
eration in MBRs.
span to the visible range. Visible combs from microres-
onators could have applications in ground, and satellite
based, astrophysical and LIDAR measurements as well
as miniaturisation of rubidium and cesium atomic clocks.
Further work on this topic is ongoing worldwide, in
order to increase the functionality of WGRs in frequency
comb applications.
Acknowledgments
This work was supported by Okinawa Institute of Sci-
ence and Technology Graduate University.
[1] M. L. Gorodetsky, A. A. Savchenkov, and V. S. Ilchenko:
Opt. Lett. 21 (1996) 453.
[2] V. S. Ilchenko, M. L. Gorodetsky, X. S. Yao, and
L. Maleki: Opt. Lett. 26 (2001) 256.
[3] T. J. Kippenberg, S. M. Spillane, and K. J. Vahala: Appl.
Phys. Lett. 85 (2004) 6113.
[4] J. Ward and O. Benson: Laser Photon. Rev. 5(2011)
553.
[5] A. Chiasera, Y. Dumeige, P. Fron, M. Ferrari, Y. Jestin,
G. N. Conti, S. Pelli, S. Soria, and G. C. Righini: Laser
Photon. Rev. 4(2010) 457.
[6] V. S. Ilchenko, A. B. Matsko, A. A. Savchenkov, and
L. Maleki: J. Opt. Soc. Am. 20 (2003) 1304.
[7] D. Farnesi, A. Barucci, G. C. Righini, S. Berneschi, S. So-
ria, and G. N. Conti: Phys. Rev. Lett. 112 (2014) 093901.
[8] Y. Wu, J. M. Ward, and S. N. Chormaic: J. Appl. Phys.
107 (2010) 033103.
[9] W. Lee, Q. Chen, X. Fan, and D. K. Yoon: Lab Chip 16
(2016) 4770.
[10] K. Scholten, W. R. Collin, X. Fan, and E. T. Zellers:
Nanoscale 7(2015) 9282.
[11] P. Wang, R. Madugani, H. Zhao, W. Yang, J. M. Ward,
Y. Yang, G. Farrell, G. Brambilla, and S. N. Chormaic:
IEEE Photonics Technol. Lett. 28 (2016) 2277.
[12] W. Yoshiki and T. Tanabe: Opt. Express 22 (2014)
24332.
[13] J. U. Frst, D. V. Strekalov, D. Elser, A. Aiello, U. L. An-
dersen, C. Marquardt, and G. Leuchs: Phys. Rev. Lett.
106 (2011) 113901.
[14] Y. Li, J. Millen, and P. F. Barker: Opt. Express 24
(2016) 1392.
[15] T. J. Kippenberg and K. J. Vahala: Science 321 (2008)
1172.
[16] F. Monifi, J. Zhang, . K. zdemir, B. Peng, Y. Liu, F. Bo,
F. Nori, and L. Yang: Nat. Photonics 10 (2016) 399.
[17] Y. Chen: J. Phys. B: At. Mol. Opt. Phys. 46 (2013)
104001.
[18] J. N. Eckstein, A. I. Ferguson, and T. W. Hnsch: Phys.
Rev. Lett. 40 (1978) 847.
[19] C. Gohle, T. Udem, M. Herrmann, J. Rauschenberger,
R. Holzwarth, H. A. Schuessler, F. Krausz, and T. W.
Hnsch: Nature 436 (2005) 234.
[20] M. R. Holzwarth, T. Udem, T. W. Hnsch, J. C. Knight,
W. J. Wadsworth, and P. S. J. Russell: Phys. Rev. Lett.
85 (2000) 2264.
[21] T. Udem, R. Holzwarth, and T. W. Hnsch: Nature 416
(2002) 233.
[22] Y. K. Chembo: Nanophotonics 5(2016) 214.
[23] P. Del’Haye, A. Schliesser, O. Arcizet, T. Wilken,
R. Holzwarth, and T. J. Kippenberg: Nature 450 (2007)
1214.
[24] M. H. P. Pfeiffer, C. Herkommer, J. Liu, H. Guo, M. Kar-
pov, E. Lucas, M. Zervas, and T. J. Kippenberg: Optica
4(2017) 684.
[25] M. Sumetsky: Opt. Lett. 29 (2004) 8.
[26] M. Li, X. Wu, L. Liu, and L. Xu: Opt. Express 21 (2013)
16908.
[27] Y. Yang, J. M. Ward, and S. N. Chormaic: Proc. SPIE
8960 (2014) 89600I.
[28] Y. Yang, Y. Ooka, R. M. Thompson, J. M. Ward, and
S. N. Chormaic: Opt. Lett. 41 (2016) 575.
[29] I. H. Agha, Y. Okawachi, M. Foster, J. Sharping, and
A. L. Gaeta: Phys. Rev. A 76 (2007) 043837.
[30] I. H. Agha, Y. Okawachi, and A. L. Gaeta: Opt. Express
17 (2009) 16209.
[31] N. Riesen, S. V. Afshar, A. Franccois, and T. M. Monro:
Opt. Express 23 (2015) 14784.
[32] G. Lin and Y. K. Chembo: Opt. Express 23 (2015) 1594.
[33] A. A. Savchenkov, A. B. Matsko, W. Liang, V. S.
Ilchenko, D. Seidel, and L. Maleki: Nat. Photonics 5
(2011) 293.
[34] K. T. Yang, K. Beha, D. C. Cole, X. Yi, P. Del’Haye,
H. Lee, J. Li, D. Y. Oh, S. A. Diddams, S. B. Papp, and
J. K. Vahala: Nat. Photonics 10 (2016) 316.
[35] M. Sumetsky, Y. Dulashko, and R. S. Windeler: Opt.
Lett. 35 (2010) 898.
[36] S. Berneschi, D. Farnesi, F. Cosi, G. N. Conti, S. Pelli,
G. C. Righini, and S. Soria: Opt. Lett. 36 (2011) 3521.
[37] Y. Yang, S. Saurabh, J. M. Ward, and S. N. Chormaic:
Opt. Express 24 (2016) 294.
[38] D. Farnesi, A. Barucci, G. C. Righini, G. N. Conti, and
S. Soria: Opt. Lett. 40 (2015) 4508.
[39] S. Miller, K. Luke, Y. Okawachi, J. Cardenas, A. L.
Gaeta, and M. Lipson: Opt. Express 22 (2014) 26517.
[40] L. Wang, L. Chang, N. Volet, M. H. P. Pfeiffer, M. Zer-
vas, H. Guo, T. J. Kippenberg, and J. E. Bowers: Laser
Photon. Rev. 10 (2016) 631.
[41] N. Riesen, W. Zhang, and T. M. Monro: Opt. Lett. 41
(2016) 1257.
[42] I. H. Malitson: J. Opt. Soc. Am. 55 (1965) 1205.
6
[43] N. Riesen, W. Zhang, and T. M. Monro: Opt. Express
24 (2016) 8832.
[44] Y. Yang, X. Jiang, S. Kasumie, G. Zhao, L. Xu, J. M.
Ward, L. Yang, and S. N. Chormaic: Opt. Lett. 41 (2016)
5266.
... These allow us to engineer properties related to WGMs such as mode field distribution, mode dispersion, and mode spectrum. Such engineered WGMs are particularly useful for various nonlinear optical processes, for example, four-wave parametric oscillation and frequency comb generation [26][27][28][29]. Therefore, an accurate determination of the geometry of a microbubble to precisely characterize its WGMs is an important prerequisite for practical applications of such cavities. ...
Article
Full-text available
Whispering gallery mode (WGM) microbubble cavities are a versatile optofluidic sensing platform owing to their hollow core geometry. To increase the light–matter interaction and, thereby, achieve higher sensitivity, thin-walled microbubbles are desirable. However, a lack of knowledge about the precise geometry of hollow microbubbles prevents us from having an accurate theoretical model to describe the WGMs and their response to external stimuli. In this work, we provide a complete characterization of the wall structure of a microbubble and propose a theoretical model for the WGMs in this thin-walled microcavity based on the optical waveguide approach. Structural characterization of the wavelength-scale wall is enabled by focused ion beam milling and scanning electron microscopy imaging. The proposed theoretical model is verified by finite element method simulations. Our approach can readily be extended to other low-dimensional micro-/nanophotonic structures.
... These allow us to engineer properties related to the WGMs, such as the mode field distribution, the mode dispersion, and the mode spectrum. Such engineered WGMs are particularly useful for various nonlinear optical processes, for example, four-wave parametric oscillation and frequency comb generation [22][23][24][25]. Therefore, an accurate determination of the geometry of a microbubble to precisely characterize its WGMs is an important prerequisite for practical applications of such cavities. ...
Preprint
Full-text available
Whispering gallery mode (WGM) microbubble cavities are a versatile optofluidic sensing platform owing to their hollow core geometry. To increase the light-matter interaction and, thereby, achieve a higher sensitivity, thin-walled microbubbles are desirable. However, a lack of knowledge about the precise geometry of hollow microbubbles prevents us from having an accurate theoretical model to describe the WGMs and their response to external stimuli. In this work, we provide a complete characterization of the wall structure of a microbubble and propose a theoretical model for the WGMs in this thin-walled microcavity based on the optical waveguide approach. Structural characterization of the wavelength-scale wall is enabled by focused ion beam milling and scanning electron microscopy imaging. The proposed theoretical model is verified by finite element method simulations. Our approach can readily be extended to other low-dimensional micro-/nanophotonic structures.
... Developing applications in many areas such as biological imaging, underwater communications and detection, atomic clocks, and quantum technologies all rely on wavelengths in this range [20]. To date, near-visible optically pumped parametric oscillation and Kerr frequency combs have been realized in silica WGRs [21][22][23][24][25][26]. Green and blue light generation through a third-harmonic process with near-IR pumping near 1550 nm also has been observed [27][28][29]. ...
Article
Full-text available
To date there are extensive studies of optical nonlinearities in microcavities in the near and mid-infrared wavelengths. Pushing this research into the visible region is equally valuable. Here, we demonstrate a directly pumped, blue band Kerr frequency comb and stimulated Raman scattering (SRS) at 462 nm in a silica nanofiber-coupled whispering gallery microcavity system. Notably, due to the high optical intensities achieved, photodarkening is unavoidable and can quickly degrade the optical quality of both the coupling optical nanofiber and the microcavity even at very low pump powers. Nonetheless, stable hyperparametric oscillation and SRS are demonstrated in the presence of photodarkening by taking advantage of in-situ thermal bleaching. This work highlights the challenges of silica-based, short wavelength nonlinear optics in high quality, small mode volume devices and gives an effective method to overcome this apparent limitation, thus providing a baseline for optics research in the blue region for any optical devices fabricated from silica.
... In the past few decades, the level of research activity on whispering gallery mode (WGM) resonators has increased rapidly [1][2][3][4][5][6][7]. Concurrently, the number of WGM geometries has also increased. ...
Article
Full-text available
We present a method for making microbubble whispering gallery resonators (WGRs) from tellurite, which is a soft glass, using a CO2 laser. The customized fabrication process permits us to process glasses with low melting points into microbubbles with loaded quality factors as high as 2.3 × 10⁶. The advantage of soft glasses is that they provide a wide range of refractive index, thermo-optical, and optomechanical properties. The temperature and air pressure dependent optical characteristics of both passive and active tellurite microbubbles are investigated. For passive tellurite microbubbles, the measured temperature and air pressure sensitivities are 4.9 GHz/K and 7.1 GHz/bar, respectively. The large thermal tuning rate is due to the large thermal expansion coefficient of 1.9 × 10⁻⁵ K⁻¹ of the tellurite microbubble. In the active Yb³⁺-Er³⁺ co-doped tellurite microbubbles, C-band single-mode lasing with a threshold of 1.66 mW is observed with a 980 nm pump and a maximum wavelength tuning range of 1.53 nm is obtained. The sensitivity of the laser output frequency to pressure changes is 6.5 GHz/bar. The microbubbles fabricated using this method have a low eccentricity and uniform wall thickness, as determined from electron microscope images and the optical spectra. The compound glass microbubbles described herein have the potential for a wide range of applications, including sensing, nonlinear optics, tunable microcavity lasers, and integrated photonics.
... In the past few decades, the level of research activity on whispering gallery mode (WGM) resonators, has increased rapidly [1][2][3][4][5][6][7]. Concurrently, the number of WGM geometries has also increased. ...
Preprint
We present a method for making microbubble whispering gallery resonators (WGRs) from tellurite, which is a soft glass, using a CO$_2$ laser. The customized fabrication process permits us to process glasses with low melting points into microbubbles with loaded quality factors as high as $2.3 \times 10^6$. The advantage of soft glasses is that they provide a wide range of refractive index, thermo-optical and optomechanical properties. The temperature and air pressure dependent optical characteristics of both passive and active tellurite microbubbles are investigated. For passive tellurite microbubbles, the measured temperature and air pressure sensitivities are 4.9 GHz/K and 7.1 GHz/bar, respectively. The large thermal tuning rate is due to the large thermal expansion coefficient of $1.9 \times 10^{-5}$ K$^{-1}$ of the tellurite microbubble. In the active Yb$^{3+}$-Er$^{3+}$ co-doped tellurite microbubbles, C-band single-mode lasing with a threshold of 1.66 mW is observed with a 980 nm pump and a maximum wavelength tuning range of 1.53 nm is obtained. The sensitivity of the laser output frequency to pressure changes is 6.5 GHz/bar. The microbubbles fabricated by this novel method have a low eccentricity and uniform wall thickness, as determined from electron microscope images and the optical spectra. The compound glass microbubbles described herein have potential for a wide range of applications, including sensing, as tunable microcavity lasers, and for integrated photonics.
Article
Full-text available
Octave-spanning, self-referenced frequency combs are applied in diverse fields ranging from precision metrology to astrophysical spectrometer calibration. The octave-spanning optical bandwidth is typically generated through nonlinear spectral broadening of femtosecond pulsed lasers. In the past decade, Kerr frequency comb generators emerged as a novel scheme offering chip-scale integration, high repetition rate, and bandwidths that are only limited by group velocity dispersion. The recent observation of the dissipative Kerr soliton (DKS) operation regime, along with dispersive wave formation, has provided the means for fully coherent, broadband Kerr frequency comb generation with an engineered spectral envelope. Here, by carefully optimizing the photonic Damascene fabrication process, and dispersion engineering of Si 3 N 4 microresonators with a free spectral range of 1 THz, we achieve bandwidths exceeding one octave at low powers ( ∼ 100 mW ) for pump lasers residing in the telecom C band (1.55 μm) as well as in the O band (1.3 μm). Precise dispersion engineering enables emission of two dispersive waves, increasing the power in the spectral ends of the comb, down to a wavelength as short as 850 nm. Investigating the coherence of the generated Kerr comb states, we unambiguously identify DKS states using a response measurement. This allows demonstrating octave-spanning DKS comb states at both pump laser wavelengths of 1.3 μm and 1.55 μm, including the broadest DKS state generated to date, spanning more than 200 THz of optical bandwidth. Octave-spanning DKS frequency combs can be applied in metrology or spectroscopy, and their operation at 1.3 μm enables applications in biological and medical imaging such as Kerr-comb-based optical coherence tomography or dual-comb coherent anti-Stokes Raman scattering.
Article
Full-text available
DNA lasers self-amplify optical signals from a DNA analyte as well as thermodynamic differences between sequences, allowing quasi-digital DNA detection. However, these systems have drawbacks, such as relatively large sample consumption and complicated dye labelling. Moreover, although the lasing signal can detect the target DNA, it is superimposed on an unintended fluorescence background, which persists for non-target DNA samples as well. From an optical point of view, it is thus not truly digital detection and requires spectral analysis to identify the target. In this work, we propose and demonstrate an optofluidic laser that has a single layer of DNA molecules as the gain material. A target DNA produces intensive laser emission comparable to existing DNA lasers, while any unnecessary fluorescence background is successfully suppressed. As a result, the target DNA can be detected with a single laser pulse, in a truly digital manner. Since the DNA molecules cover only a single layer on the surface of the laser microcavity, the DNA sample consumption is a few orders of magnitude lower than that of existing DNA lasers. Furthermore, the DNA molecules are stained by simply immersing the microcavity in the intercalating dye solution, and thus the proposed DNA laser is free of any complex dye-labelling process prior to analysis.
Article
Full-text available
A vapor sensor comprising a nanoparticle-coated microfabricated optofluidic ring resonator (µOFRR) is introduced. A multilayer film of polyether functionalized thiolate monolayer protected gold nanoparticles (MPN) was solvent cast on the inner wall of the hollow cylindrical SiOx µOFRR resonator structure, and whispering gallery mode (WGM) resonances were generated with a 1550-nm tunable laser via an optical fiber taper. Reversible shifts in the WGM resonant wavelength upon vapor exposure were detected with a photodetector. The μOFRR chip was connected to a pair of upstream etched-Si chips containing PDMS-coated separation μcolumns and calibration curves were generated from the peak-area responses to five volatile organic compounds (VOCs). Calibration curves were linear, and the sensitivities reflected the influence of analyte volatility and analyte-MPN functional group affinity. Sorption-induced changes in film thickness apparently dominate over changes in the refractive index of the film as the determinant of responses for all VOCs. Peaks from the MPN-coated µOFRR were just 20-50% wider than those from a flame ionization detector for similar μcolumn separation conditions, reflecting the rapid response of the sensor for VOCs. The five VOCs were baseline separated in < 1.67 min, with detection limits as low as 38 ng.
Article
Full-text available
Optical whispering gallery mode (WGM) resonators have been very attracting platforms for versatile Kerr frequency comb generations. We report a systematic study on the material dispersion of various optical materials that are capable of supporting quality factors above 10⁹. Using an analytical approximation of WGM resonant frequencies in disk resonators, we investigate the effect of the geometry and transverse mode order on the total group-velocity dispersion (GVD). We demonstrate that the major radii and the radial mode indices play an important role in tailoring the GVD of WGM resonators. In particular, our study shows that in WGM disk-resonators, the polar families of modes have very similar GVD, while the radial families of modes feature dispersion values that can differ by up to several orders of magnitude. The effect of these giant dispersion shifts are experimentally evidenced in Kerr comb generation with magnesium fluoride. From a more general perspective, this critical feature enables to push the zero-dispersion wavelength of fluorite crystals towards the mid-infrared (mid-IR) range, thereby allowing for efficient Kerr comb generation in that spectral range. We show that barium fluoride is the most interesting crystal in this regard, due to its zero dispersion wavelength (ZDW) at 1.93 μm and an optimal dispersion profile in the mid-IR regime. We expect our results to facilitate the design of different platforms for Kerr frequency comb generations in both telecommunication and mid-IR spectral ranges.
Article
Full-text available
High quality factor whispering-gallery-mode microresonators are ideally suited for nonlinear optical interactions. We analyze, experimentally and theoretically, a variety of χ(3) nonlinear interactions in silica microspheres, consisting of third harmonic generation and Raman assisted third order sum-frequency generation in the visible. A tunable, room temperature, cw multicolor emission in silica microspherical whispering-gallery-mode microresonators has been achieved by controlling the cavity mode dispersion and exciting nonequatorial modes for efficient frequency conversion.
Article
Full-text available
Silica micro-bubble resonators (MBRs) with cavity quality factor as high as Q = 5 × 10⁷ are fabricated. The total dispersion of MBRs is analyzed. The thin-wall structure opens a new anomalous dispersion window and thus supports the dispersion compensation for hyper-parametric frequency conversion processes. Experimentally, Kerr parametric oscillation is observed in a 136 μm diameter MBR, frequency comb generation is also realized. Meanwhile the same nonlinear process is not allowed in solid silica spheres with size smaller than 150 μm.
Article
In this letter, a packaged add-drop filter composed of a silica microsphere resonator with a diameter of 80 µm and an optical microfiber coupler with a waist diameter of 1 µm is investigated. A one-step fabrication process using UV curable epoxy is shown to stabilize the coupling between the microsphere and the microfiber coupler, which is used for the add and drop ports. The packaged microsphere-microfiber coupler device has a high Q-factor of 1 × 107 in the add–drop filter configuration. This device has excellent features, such as improved ease of fabrication, multiple resonant peaks, and high mechanical stability.
Article
Chaotic dynamics has been reported in many physical systems and has affected almost every field of science. Chaos involves hypersensitivity to the initial conditions of a system and introduces unpredictability into its output. Thus, it is often unwanted. Interestingly, the very same features make chaos a powerful tool to suppress decoherence, achieve secure communication and replace background noise in stochastic resonance - a counterintuitive concept that a system's ability to transfer information can be coherently amplified by adding noise. Here, we report the first demonstration of chaos-induced stochastic resonance in an optomechanical system, as well as the optomechanically mediated chaos transfer between two optical fields such that they follow the same route to chaos. These results will contribute to the understanding of nonlinear phenomena and chaos in optomechanical systems, and may find applications in the chaotic transfer of information and for improving the detection of otherwise undetectable signals in optomechanical systems.
Article
We propose to fabricate a toroidal dielectric cavity from a periodically poled χ(2) nonlinear material (e.g., LiNbO3) to achieve efficient interaction among high-Q whispering-gallery modes. We show that the periodic poling allows for suppression of both material and cavity dispersion. Such a cavity might be a basic element of a family of efficient nonlinear devices operating at a broad range of optical wavelengths.
Article
We have demonstrated the feasibility of Doppler-free two-photon spectroscopy with a train of picosecond standing-wave light pulses from a synchronously pumped mode-locked cw dye laser. The actively controlled mode spectrum provides a means for accurate measurements of large frequency intervals. From a multipulse spectrum of the sodium 3s-4d transition we have determined a new value of the 4d fine-structure splitting, 1028±0.4 MHz.