PreprintPDF Available

Nonlinear Multi-Resonant Cavity Quantum Photonics Gyroscopes Quantum Light Navigation

Authors:
Preprints and early-stage research may not have been peer reviewed yet.

Abstract

We propose an on-chip all-optical gyroscope based on nonlinear multi-resonant cavity quantum photonics in thin film $\chi^{(2)}$ resonators -- Quantum-Optic Nonlinear Gyro or QONG in short. The key feature of our gyroscope is co-arisal and co-accumulation of quantum correlations, nonlinear wave mixing and non-inertial signals, all inside the same sensor-resonator. We theoretically analyze the Fisher Information of our QONGs under fundamental quantum noise conditions. Using Bayesian optimization, we maximize the Fisher Information and show that $\sim 900\times$ improvement is possible over the shot-noise limited linear gyroscope with the same footprint, intrinsic quality factors and power budget.
Nonlinear Multi-Resonant Cavity Quantum Photonics Gyroscopes
for Quantum Light Navigation
Mengdi Sunand Vassilios Kovanis
Bradley Department of Electrical and Computer Engineering, Virginia Tech, Arlington, VA, USA
Marko Lonˇcar
John A. Paulson School of Engineering and Applied Sciences, Harvard University, Cambridge, MA, USA
Zin Lin
Bradley Department of Electrical and Computer Engineering, Virginia Tech, Blacksburg, VA, USA
(Dated: July 22, 2023)
We propose an on-chip all-optical gyroscope based on nonlinear multi-resonant cavity quantum
photonics in thin film χ(2) resonators—Quantum-Optic Nonlinear Gyro or QONG in short. The key
feature of our gyroscope is co-arisal and co-accumulation of quantum correlations, nonlinear wave
mixing and non-inertial signals, all inside the same sensor-resonator. We theoretically analyze the
Fisher Information of our QONGs under fundamental quantum noise conditions. Using Bayesian
optimization, we maximize the Fisher Information and show that 900×improvement is possible
over the shot-noise limited linear gyroscope with the same footprint, intrinsic quality factors and
power budget.
I. INTRODUCTION
Gyroscopes are critical components of an inertial nav-
igation system for augmenting the GPS guidance or sal-
vaging GPS-denied operational environments [1]. In an
optical gyroscope, the rotation rate is measured through
the phase shift between two counter-propagating beams
in an optical loop. This approach was first proposed by
Sagnac in 1913 [2, 3] and soon different types of op-
tical gyroscopes were developed [4, 5]. Careful studies
have been performed on the sensitivity and the quantum
noise of these gyroscopes [6, 7], and remarkable levels
of sensitivity (<0.001/h) have been achieved in state-
of-the-art discrete component optical gyroscopes, includ-
ing fiber optic gyroscopes (FOG) [8–10], ring laser gy-
roscopes (RLG) [11], atom-laser gyroscope [12] and op-
tical cavity gyroscopes [1315]. However, bulky compo-
nents and relatively high power consumption remain ma-
jor roadblocks to further exploiting discrete component
optical gyroscopes. On the other hand, on-chip optical
gyroscopes [16–20] exhibit great potential for fully inte-
grated inertial navigation platforms (free of fragile mov-
ing parts) and can outperform their discrete component
counterparts in size, weight, power consumption, ma-
neuverability, manufacturing scalability, robustness and
the ability to operate in harsh environments. However,
on-chip gyros are yet to reach sensitivity levels smaller
than 1/h. This is due to fundamentally limited optical
path lengths even in ultra-high quality factor resonan-
tors [16], leaving dubious prospects for further improve-
ments via increasing resonator size or quality factors. To
address the challenge of this seemingly intrinsic trade-
off between sensitivity and compactness, novel physics
mengdis@vt.edu
and designs have been investigated, including exceptional
point sensing [21], slow light [22, 23], dispersive enhance-
ments [24, 25], dynamic thermal drift cancellation [15, 26]
and nuclear magnetic resonance [27].
Meanwhile, driven by the emerging trend of quantum
technologies [28–30], quantum light sensors have been
identified as a promising option that can extend the fun-
damental sensitivity limits beyond the shot noise regime
[3133]. These ideas were reinforced by decades of de-
velopment and analysis that lead to the construction of
very large laser interferometers with extreme sensitivity
that is capable of detecting gravitational waves from re-
mote cosmological events. Recently, squeezed light was
used in the LIGO in the US and the VIRGO in Italy
to substantially improve the sensitivity of the observing
runs that happened late in April 2019 [34, 35]. On the
other hand, recent advances in nanofabrication, integra-
tion and packaging of ultra-coherent laser sources [36],
low-loss photonic circuits [37, 38] and highly efficient
photo-detectors [39] have opened up exciting opportu-
nities for realizing fully on-chip quantum devices. Along
this trend, we identify on-chip quantum light gyroscopes,
which combine high sensitivities, low power consumption,
and small form factors, as promising candidates for next-
generation rotation sensing.
In this paper, we theoretically introduce a new type of
on-chip quantum light gyroscope that exploits nonlinear
multi-resonant cavity quantum photonics in integrated
thin film resonators with strong quadratic χ(2) nonlinear-
ities. We call our gyroscope Quantum-Optic Nonlinear
Gyro or QONG in short. Instead of externally inject-
ing quantum states of light into the gyroscope [32], one
of the distinguishing features of our gyroscope is that it
fuses quantum-coherent nonlinear interactions, quantum
light generation and non-inertial signal accumulation in-
side the same sensor-resonator, enabling 900×im-
arXiv:submit/5020283 [quant-ph] 22 Jul 2023
2
provements in gyroscopic sensitivity over the linear shot
noise limit. In our scheme, classical laser light (coherent
state) is injected into a doubly-resonant χ(2) cavity, and
output light is measured at the fundamental (ω1) and
second-harmonic frequencies (ω2= 2ω1). The sensitiv-
ity of the gyroscope is evaluated by Fisher information
(FI) [40, 41], and the latter is maximized by Bayesian
optimization [42]. Various parameter regimes associated
with both fundamental and second harmonic injection
schemes were investigated, which reveal correlated noise
suppression and sensitivity enhancements via paramet-
ric oscillations and critically sensitive three wave mixing
dynamics. We predict that, under quantum noise con-
ditions, a minimum detectable rotation rate (MDR) of
<0.01 /h can be achieved using a thin film lithium nio-
bate (TFLN) ring resonator with a diameter of 20 mm,
intrinsic quality factors Qi2= 106at the second har-
monic wavelength (795 nm), Qi1= 107at the fundamen-
tal wavelength (1590 nm). We discuss the scope, validity
and implications of our approach and results, while the
key sensitivity enhancement factors are summarized in
Table I of Section III D.
II. GYROSCOPIC MODEL
A. Linear resonant gyroscope as a baseline
We first review a basic interferometric scheme probing
the gyroscopic shift of a linear resonant cavity, as outlined
in Fig. 1. We perform a quantum noise analysis similar to
Ref. [43] or Section 4 of Ref. [7]. Two identical counter-
propagating probes (seeded from the same on-chip laser)
are injected into the clockwise (CW) and the counter-
clockwise (CCW) modes of a ring resonator; at the exit,
the two probe fields are set to interfere via balanced ho-
modyne detection [44]. In the absence of rotation, the
CW and the CCW modes are degenerate and the exiting
fields register a vanishing differential photocurrent signal
at the detection setup [7]. Rotational motion induces a
frequency splitting proportional to the rotation rate Ω,
which in turn induces a phase difference between the out-
going CW and CCW probes. Subsequently, interference
of the two probe fields gives rise to a non-zero differen-
tial signal and the underlying can be measured. For
conceptual simplicity, we assume that the frequency of
the probe laser is always locked to the degenerate fre-
quency of the unperturbed gyro [7]. In principle, this
can be achieved by self-injection locking the laser to an
independent rotation-insensitive cavity (such as a high-Q
spiral resonator [45]) having the exact same frequency as
the unperturbed gyro ring. It has been demonstrated
[46] that self-injection locking to a high-Q cavity can
produce an ultra-coherent integrated laser with a sub-
Hertz linewidth; therefore, we can readily approximate
the laser state as a quantum-mechanical coherent state.
In the Heisenberg picture, the ring resonator gyro obeys
the Heisenberg-Langevin equations [47]:
dˆacw
dt =κ
2γ
2+ˆacw +ˆaccw +κˆ
bin
cw +γˆcin
cw
(1)
dˆaccw
dt =κ
2γ
2ˆaccw +ˆacw +κˆ
bin
ccw +γˆcin
ccw
(2)
where ˆacw and ˆaccw are the annihilation operators for
the cavity CW and CCW modes excited by the injec-
tions ˆ
bin
cw and ˆ
bin
ccw respectively. ˆcin
cw and ˆcin
ccw represent
intrinsic loss channels (such as radiative losses). κand
γare the decay rates for the coupling and the intrinsic
losses. We approximate Rayleigh-type back-scattering as
a linear (conservative) coupling βbetween CW and CCW
modes inside the cavity [16]. δis the rotation-induced
resonant frequency shift due to the Sagnac effect. Here
we have assumed single-photon normalization for each
eigenmode so that ˆaˆa, for example, represents the pho-
ton number operator inside the cavity.
We denote ˆ
A=ψ|ˆ
A|ψas the usual notation for
computing the expectation value of a physical observable
ˆ
Awith respect to the quantum state |ψ. In the linear
problem, we will consider coherent states of the same am-
plitude bin in the input waveguides and vacuum states in
the intrinsic loss channels for both CW and CCW light
[7]. The input quantum state of the gyro is then given
by |ψ=|bincw |binccw |0cw |0ccw . The classical coun-
terpart of the input operator ˆ
bin is the input amplitude
of a coherent state in the feeder waveguide, and can be
related to the input power Pby the formula:
|bin|2=ˆ
binˆ
bin=P
ω(3)
The output operators in the waveguides are given by
[7]:
ˆ
bout
cw =ˆ
bin
cw κ1ˆacw (4)
ˆ
bout
ccw =ˆ
bin
ccw κ1ˆaccw (5)
The clockwise and counterclockwise signals are set
to interfere through a directional coupler/beam splitter
with a controllable phase shift ϕ, followed by photode-
tection. The signal incident on the photodetectors and
the photocurrent operators are then given by:
ˆ
b+=ˆ
bout
cw eiϕ/2+iˆ
bout
ccweiϕ/2/2 (6)
ˆ
b=iˆ
bout
cw eiϕ/2+ˆ
bout
ccweiϕ/2/2 (7)
ˆ
i+=ˆ
b
+ˆ
b+(8)
ˆ
i=ˆ
b
ˆ
b(9)
We measure the differential current signal:
3
ˆ
i=ˆ
i+ˆ
i(10)
As a figure of merit, we will investigate the minimum
detectable frequency shift by calculating the ratio be-
tween the standard variation of the measured differen-
tial current and the derivative of the mean value of the
current over the rotation-induced frequency shift, as re-
ported by Dowling in 1998 [12]:
δmin =qˆ
i2⟩−⟨ˆ
i2
ˆ
i
∂δ δ=0
(11)
Since the resonant frequency shift due to the Sagnac
effect is given by δ=2πr
λn0[7], the minimum detectable
rotation rate (MDR) is given by:
min =λn0
2πR qˆ
i2⟩−⟨ˆ
i2
ˆ
i
∂δ δ=0
(12)
where R and n0are the radius and the refractive index
of the micro-ring. λis the wavelength of the input light.
We emphasize that min is a holistic measure that con-
siders the deterministic sensitivity of the noise-averaged
photocurrent with respect to as well as the variance
of the measured current signals due to noise (Both are
critical to correctly characterizing the overall sensitivity
of the gyro; it has been pointed out [48] that an analysis
only of the deterministic sensitivity could often lead to
misleading conclusions).
If we ignore Rayleigh back-scattering β= 0, we can
derive a simple closed-form expression for MDR in the
linear gyroscope (LG):
LG
min =2λn0(κ+γ)2
32πRκN(13)
where N=P
ωis the incident number of photons
per unit time. LG
min is minimized at κ=γ, yielding
MDRLG
min =2cn0
4RNQi, where the intrinsic quality factor is
defined by Qi=ω
γ. Note that this is only an example to
illustrate the sensitivity dependence of the simplest linear
gyroscope without considering Rayleigh back-scattering,
which will be taken into account in the following dis-
cussions. We note that, in our analysis, we only con-
sider fundamental quantum noise: without loss of gen-
erality, we have assumed perfect beam splitters and de-
tectors external to the resonator, while we do consider
realistic losses inside the resonator. This linear quan-
tum result will serve as a baseline comparison for our
later analysis of a new mode of gyroscope that relies on
nonlinear quantum optical effects. Note that the scaling
MDRLG
min 1
NQirecovers the familiar shot noise limit
or the standard quantum limit [12]. In addition, the 1
R
dependence in Equation (13) indicates that a larger ring
radius Roffers better sensitivity, which is one of the com-
mon control knobs of classical linear optical gyroscope.
FIG. 1. The schematic of the linear micro-ring gyroscope.
The input and output light of the micro-ring cavity is injected
at two waveguide ports bin
1(CW) / bin
1(CCW) and bout
1(CW)
/bout
1(CCW). The radiation losses are expressed by the fic-
ticious radiation channel cin
1(CW) / cin
1(CCW). The output
light is measured by homodyne detection.
B. Nonlinear Multi-Resonant Cavity Quantum
Photonics Gyro
We now consider the gyroscopic operation of a doubly
resonant ring resonator with quadratic χ(2) nonlinearities
(Fig. 2). Quadratic nonlinearities are well-known gener-
ators of quantum-coherent correlations such as squeez-
ing and entanglement [49, 50]. Our Nonlinear Multi-
resonant Cavity Quantum Photonics Gyro, or Quantum-
Optic Nonlinear Gyro (QONG) in short, fuses nonlinear
dynamics, quantum correlations, and non-inertial Sagnac
effects in the same sensor-resonator, to maximally lever-
age any possible nonlinear quantum-optical effects for gy-
roscopic sensitivity. Specifically, we investigate the fol-
lowing Heisenberg-Langevin equations:
4
dˆa1,cw
dt =κ1
2+γ1
21ˆa1,cw +1ˆa1,ccw
+χˆa
1,cwˆa2,cw +κ1ˆ
bin
1,cw +γ1ˆcin
1,cw
(14)
dˆa1,ccw
dt =κ1
2+γ1
2+1ˆa1,ccw +1ˆa1,cw
+χˆa
1,ccwˆa2,ccw +κ1ˆ
bin
1,ccw +γ1ˆcin
1,ccw
(15)
dˆa2,cw
dt =κ2
2+γ2
22ˆa2,cw +2ˆa2,ccw
1
2χˆa2
1,cw +κ2ˆ
bin
2,cw +γ2ˆcin
2,cw
(16)
dˆa2,ccw
dt =κ2
2+γ2
2+2ˆa2,ccw +2ˆa2,cw
1
2χˆa2
1,ccw +κ2ˆ
bin
2,ccw +γ2ˆcin
2,ccw
(17)
In these equations, the index j= 1,2 in the field op-
erators aj,ˆ
bj, ˆcj) stands for the fundamental ω1and
second harmonic ω2= 2ω1resonances. κjand γjare
the decay rates of the coupling and the intrinsic loss
channels. Rayleigh scattering rate between CW and
CCW modes at each resonance jis again character-
ized by βj. The nonlinear coupling terms χˆa
1ˆa2indi-
cate a multi-photon process in which one incident pho-
ton with ω2breaks down into two photons of half the
frequency ω1=ω2/2 (parametric down conversion [51]),
or its reverse, 1
2χˆa2
1, indicating that two photons with
ω1combine into one photon with the double frequency
ω2= 2ω1(second harmonic generation [52]). These are
quantum-coherent, energy-conserving, three-wave mix-
ing processes, which preserve the fundamental commuta-
tion relations [53]. The rotation-induced frequency shifts
(δ1,δ2) are different for each resonance, have opposite
polarity between CW and CCW modes, and can be ap-
proximated by δ2= 2δ1(since δ=2πr
λn0[7]). Note that
here the material dispersion of the lithium niobate is ne-
glected due to relatively small index difference (n=2.21
at 1590 nm and n=2.25 at 795 nm). In order to improve
the accuracy of the model, however, the dispersion ef-
fect should be taken into account in future exploration.
The key parameter in this model is the nonlinear modal
coupling strength [47, 54]:
χ=ϵ0
ZZZ 3χ(2)(r)
42u
1(z, r, θ)2u2(z, r, θ)rdrdθdz (18)
where u
1(z, r, θ) and u2(z, r, θ) are the electric field
profiles (in polar coordinates) of the fundamental and the
second harmonic eigenmodes of the gyroscopic resonator.
Given that our sensor is a ring resonator of radius R, it
is instructive to decompose χinto cross-sectional modal
overlap ζand the remaining contributions. Following
[55], we approximate:
χsω2
1ω2
ϵ02πR
ζ
ϵ1ϵ2
3χ(2)
42(19)
ζ=RR u
1(z, r)2u2(z , r)drdz
RR |u
1(z, r)|2drdzqRR|u2(z, r)|2drdz
(20)
It is important to realize that the nonlinear Langevin
equations [47, 56] encode the time evolution of four cou-
pled infinite-dimensional quantum operators; as such, it
is very challenging to obtain an exact solution either an-
alytically or numerically (we note that straightforward
numerical methods using a truncated Fock basis [57] are
not feasible because our system typically involves milli-
watts of optical power amounting to 1016 photons).
However, at milli-watt injection powers, quantum fluc-
tuations can be considered “small signals” compared to
much stronger average field intensities at steady state,
so that each operator can be decomposed into a classi-
cal scalar-valued amplitude and a quantum fluctuation
operator, e.g. ˆa=α+ˆ
δa. The details of calculat-
ing steady state solutions are included in Appendix A.
The classical amplitude represents a steady-state solu-
tion to the mean-field averaged Langevin equations at the
classical (large photon number) limit while the “small-
signal” fluctuation operator approximately obeys the lin-
earized Langevin equations in the vicinity of the steady-
state mean-field solution. Linearizing a nonlinear steady
state to study the fluctuations in its vicinity is com-
monly known as small-signal modeling in electronics en-
gineering [58]. In a similar spirit, the small-signal treat-
ment of quantum fluctuation operators in the Heisenberg-
Langevin picture is a simple but effective approach widely
accepted for steady-state noise analysis in laser and non-
linear quantum optics literature with experimental sup-
port [7, 5962]. Theoretically, it is important to note that
such an approach is justified as long as the steady state
we consider is a hyperbolic fixed point whose neighbor-
hood is a topologically stable manifold that ensures small
fluctuations (Hartman-Grobman theorem [63]). On the
other hand, a more sophisticated phase-space formalism,
which employs quasi-probability distributions, Fokker-
Planck equations and stochastic calculus, can be used to
study more complicated dynamics such as large fluctu-
ations at non-hyperbolic critical points and self-pulsing
(limit-cycle) solutions [64]. Using the small-signal ap-
proximation, we can compute the differential photocur-
rent signals at both the fundamental and the second har-
monic resonances (see also Fig. 2):
ˆ
i1=iA1ˆ
bout
1,cwˆ
bout
1,ccwe1ˆ
bout
1,ccwˆ
bout
1,cwe1(21)
ˆ
i2=iA2ˆ
bout
2,cwˆ
bout
2,ccwe2ˆ
bout
2,ccwˆ
bout
2,cwe2(22)
Here A1and A2are constant factors determined by
the frequencies of the light and the responsivity of the
5
photodetectors. e1and e2are the propagation phase
shifts that each output light experiences, which can be set
to zero here. Here we measure both ˆ
i1and ˆ
i2to extract
maximal information out of the nonlinear wave-mixing
gyro.
The output of our quantum-optic nonlinear gyro
(QONG) is now characterized by a mean vector iand
a covariance matrix i2:
i=ˆ
i1
ˆ
i2(23)
i2= ˆ
i2
1⟩−⟨ˆ
i12ˆ
i1ˆ
i2+ˆ
i2ˆ
i1
2 ˆ
i1⟩⟨ˆ
i2
ˆ
i1ˆ
i2+ˆ
i2ˆ
i1
2 ˆ
i2⟩⟨ˆ
i1 ˆ
i2
2⟩−⟨ˆ
i22!
(24)
Assuming that the joint probability distribution of the
measured photocurrents follow a bi-variate Gaussian, we
can express the Fisher information [41] of our QONG:
I(δ)=(di
)Ti21(di
) (25)
The details of the statistial analysis of the differential
currents are included in Appendix B. Then the sensitivity
is determined by Cramer-Rao bound [65]:
δmin =1
pI(δ)(26)
Similar to the linear gyroscope, here MDR is given by:
min =λn0
2πR δmin (27)
Before we provide further estimation for particular
QONG implementation, we want to offer a few remarks
of our modeling approach:
In this paper, we have stuck to a Langevin descrip-
tion of our quantum gyro, which takes into account
quantum noise through the Langevin fluctuation
operator ˆ
bor ˆcin each coupling or dissipation chan-
nel (with the rates determined by the fluctuation-
dissipation theorem), preserving the fundamental
commutation relations [47]. An equivalent formu-
lation can also be considered in terms of a density
operator ˆρ[66], leading to a Lindblad Master equa-
tion of the form (under vacuum noise conditions):
dˆρ
dt =i
[H, ˆρ] + Xγi[[ˆai,j ,ˆρ],ˆa
i,j ] (28)
H=Xωi,j ˆa
i,j ˆai,j +Xiκi
2(ˆ
b
i,j ˆai,j + ˆa
i,jˆ
bi,j )
+Xiχ
2a
1,j 2ˆa2,j ˆa2
1,j ˆa
2,j )
+Xβiˆa
i,cwˆai,ccw + ˆa
i,ccwˆai,cw(29)
We note that while the Langevin form is widely
utilized in many experimental situations [55], more
sophisticated theoretical analysis, delineating the
open-system quantum dynamics [67], can be per-
formed using the density operator formalism and
the Master equation, which will be the subject
of future investigations. In particular, our sim-
ple perturbative approach restricts our solution to
examine the quantum fluctuations around a sta-
ble hyperbolic fixed point. On the other hand,
non-hyperbolic fixed points and non-steady state
attractors (such as limit cycles) require more so-
phisticated non-perturbative treatment (while their
implications for quantum correlations and sensing
remain unexplored). One such treatment involves
expanding the density operator in a non-diagonal
coherent state basis (so-called positive P represen-
tation), deriving a Fokker-Planck equivalent of the
Lindblad Master equation and simulating the asso-
ciated stochastic dynamics [47]. However, to the
best of our knowledge, Fokker-Planck equations
corresponding to more than two bosonic operators
[64, 66] have not been well studied; our nonlinear
multi-resonant cavity quantum photonic gyro is de-
scribed by 4 coupled Langevin equations and will
lead to an 8+1 dimensional Fokker Planck equa-
tion, which requires substantial computational re-
sources and will be the subject of future investiga-
tions.
In our approach, we have assumed idealized sources
and detectors in order to simplify our gyroscopic
model to physically most crucial components, and
thereby to unveil the fundamental information-
theoretic limits (in the same spirit as the analysis
presented in Ref. [43] or the Section 4 of Ref. [7]).
Future works will develop more detailed models
that can compute commonly accepted experimen-
tal metrics such as the integration-time depen-
dent Allan deviation curve [68], for example, by
incorporating the quantum theory of photodetec-
tion [69, 70], which can explicitly take into account
photo-electron generation rates and detector inte-
gration times.
Last but not least, we note that our present model
focuses on χ(2) processes to delineate their effects
on the gyroscopic sensitivity. A more thorough
gyro model may also consider χ(3) (Kerr-type self
modulation) nonlinearities, which may come into
effect at ultra-high quality factors and are found to
limit the sensitivity of the (otherwise) linear gyro-
scope [7, 16]. While we shall take into account χ(3)
processes in detailed comprehensive models in the
future (see Section IV), we note that Eqs. 14-17 are
fully applicable to material platforms, such as thin
film lithium niobate [54, 71], which possess promi-
nent χ(2). Furthermore, unlike their linear coun-
terparts, resonators with χ(2) can be engineered to
6
exhibit negative Kerr shifts via cascaded second-
order effects [72], which can mitigate the intrinsic
positive Kerr shift; we shall investigate such cancel-
lation schemes in our future works. On the other
hand, we would like to emphasize that χ(3) pro-
cesses, including even the Kerr shift, need not be
treated as a nuisance, but as extra complexities and
additional degrees of freedom that can be optimized
to our advantage (see Section IV). For example, it
has been recently reported that the bistability ef-
fects associated with the Kerr-shift self-modulation
can even enhance sensitivities under appropriate
sensing schemes [73, 74].
FIG. 2. The schematic of the nonlinear micro-ring gyroscope.
The input light is injected at four waveguide ports bin
1(CW)
/bin
1(CCW) and bin
2(CW) / bin
2(CCW). The input/output
and loss channels for the second harmonic light are expressed
by blue arrows. The output light is measured by a joint mea-
surement using the differential currents at both output ports
(bout
1and bout
2). The accuracy of the measurement is deter-
mined by evaluating the quantum Fisher information of the
output light.
C. Thin film lithium niobate as an implementation
platform
Our nonlinear multi-resonant cavity quantum photonic
gyro (or quantum-optic nonlinear gyro QONG in short)
can be implemented in any thin film material platform,
including LiNbO3 [71], AlN [75], SiC [76], GaAs [77], etc,
which has prominent χ(2). In this work we consider thin
film lithium niobate (TFLN) as a particularly promising
platform, as it has gained widespread popularity for re-
alizing quantum-grade ultra-low loss photonic integrated
circuits [78, 79]. Indeed, lithium niobate has been tra-
ditionally employed in quantum optics applications as
a nonlinear medium for generating squeezed light and
entangled photon states [80, 81]. However, traditional
LN crystals are bulky and suffer from relatively lim-
ited strength of light-matter interactions (leading to very
weak nonlinear coupling χ103Hz in Eqs. 19). Only
recently, high quality wafer-scale TFLN becomes widely
available for realizing integrated photonic circuits with
nonlinear and electro-optic functionalities [71]. Associ-
ated with large χ(2), low optical loss and strong nanopho-
tonic confinement [82], TFLN devices offer orders of mag-
nitude enhancements in nonlinear coupling χ106Hz
[54].
Fig. 3 shows the design of a TFLN ring resonator which
can used as an QONG. Note that feeder waveguides
of different dimensions, frequency cutoffs, and disper-
sion characteristics, can be designed to selectively cou-
ple to the fundamental (1590nm) and the second har-
monic (795nm) modes [83], and their coupling rates can
be further tuned by TFLN electro-optics [84]. To realize
strong nonlinear coupling χbetween the two resonances,
two zeroth-order transverse electric eigenmodes (TE00)
can be (quasi-)phase-matched [54] via periodic poling [85]
that achieves crystal domain inversion, leading to peri-
odically varying nonlinear susceptibility χ(2) which com-
pensates wave vector mismatch between the fundamental
and the second harmonic modes kχ=k(ω2)2k(ω1) [54].
Apart from the phase-matched resonator itself, a fully in-
tegrated QONG can be implemented in TFLN, incorpo-
rating flip-chip bonded semiconductor lasers [86] and het-
erogeneously integrated uni-travelling carrier photode-
tectors [87]. Furthermore, we note that TFLN comes
with unique electro-optic control capabilities [71] which
can be used for tuning resonator parameters such as the
coupling rates to the waveguides [88], managing long
term temperature stability, cancelling thermal drifts and
electronic noise [15, 89], and performing signal processing
[90].
FIG. 3. The 3D schematic (not drawn to scale) of quasi-phase-
matching (QPM) achieved by periodic poling. Poled rings are
used to form quasi-phase matched structures. Semiconductor
lasers [86] and photodetectors [87] are also integrated on the
chip. The numerically simulated field profiles of both the
fundamental frequency (1590 nm) and the second harmonic
(795 nm) cavity modes are shown in the inset.
In our gyroscopic model, intrinsic losses γand back-
scattering rates βshould be treated as “fixed” parame-
ters which depend on the experimentally feasible char-
7
acteristics of a particular implementation platform, such
as residual material losses and surface roughness due to
fabrication imperfections. In thin film lithium niobate,
intrinsic quality factors reaching 108have been demon-
strated [91], and we expect proportionate back-scattering
rates with the same order of magnitude. It is impor-
tant to realize that, apart from γand β, almost all other
parameters can be designed, engineered and optimized,
including injection powers, coupling rates, resonator ra-
dius, resonator waveguide cross section and dispersion as
well as quasi-phase matching processes. We will utilize
these parameters as degrees of freedom (DoF) in optimiz-
ing the Fisher Information and hence the minimum de-
tectable rotation (MDR) of our QONG (as compared to
the linear gyro). While any optimization algorithm can
be employed, gradient-free global optimization methods
are most suitable for a relatively low-dimensional prob-
lem like our two-resonance gyro (where about 5–10 DoFs
can be optimized). We will use Bayesian optimization,
a simple but powerful machine-learning-based optimiza-
tion algorithm which requires relatively few function eval-
uations (as compared to other heuristic methods such as
simulated annealing and evolutionary algorithms [92, 93])
and has been observed to be particularly effective for op-
timizing 20 DoFs [42].
III. RESULTS
As noted above, the Fisher Information and the Min-
imum Detectable Rotation (MDR) of the gyroscope is
determined by various parameters and can be optimized
by judiciously adjusting their values. In our design, the
operational wavelengths (of the input/output light) are
fixed at λ1= 1590 nm and at λ2=795 nm. The mate-
rial property of TFLN is taken from the literature [82]:
the refractive index is n= 2.2 while the second order
nonlinear susceptibility is χ(2) = 30 pm/V. The intrin-
sic quality factors are fixed at Qi1= 107for the 1590
nm and Qi2= 106for the 795 nm. The Rayleigh scat-
tering rates for both cavity modes β1= 5.4×104Hz
and β2= 5.4×105Hz are inferred by adjusting the
literature-reported values [16] to the quality factors Qi
of our TFLN platform. We have also fixed the radius of
the resonator at R= 20 mm as well as the cross-sectional
dimensions of the resonator waveguide (1.2µm width
×0.6µm thickness) and the fabrication side wall angle
of 75, leading to a cross-sectional area of 0.8µm2. The
two TE00 modes have phase mismatch of 1.354 µm1,
which can be compensated by poling with a period of
4.64 µm. Based on Eqs. 19, the quasi-phase matched
χ(2) [54], and the numerically simulated modal overlap-
ping factor ζ= 1.18 /µm, the nonlinear coupling strength
is χ= 1.26×106Hz, which is independent of the injection
schemes. The rest of the parameters remain to be deter-
mined, including the injection power at the fundamental
and the second harmonics P1and P2, the quality factors
due to coupling to the waveguides for each cavity mode:
Qc1and Qc2. These four parameters will be determined
by Bayesian optimization. We will investigate gyroscopic
performance under different injection schemes including
(1) coherent state input at the second harmonic (λ2), (2)
coherent state input at the sub-harmonic (λ1), and (3)
coherent state inputs at both second and sub-harmonics
(λ1and λ2).
A. Optical parametric oscillator gyro (coherent
injection at second harmonic)
First, we study the performance of an optical para-
metric oscillator gyroscope under the coherent injection
at the second harmonic frequency. As shown in Fig.
4(a), classical laser light with a wavelength λ2= 795
nm is injected from opposite directions from the waveg-
uide ports bin
2,cw and bin
2,ccw, while no light is injected from
the waveguide port bin
1,cw or bin
1,cw, i.e., P1= 0. The care-
fully phase-matched fundamental and second harmonic
modes facilitate parametric down conversion, in which
one photon with higher frequency (shorter wavelength
λ2breaks down into two photons with half the frequency
(longer wavelength λ1), generating phase-squeezed sig-
nals [47]. First we investigate the mininum detectable
rotation (MDR) as a function of the input power P2and
the coupling factor Qc2. Using Bayesian optimization,
we identify a high sensitivity regime, that is, low MDR,
as shown in the 2D density plot of MDR in terms of
P2and Qc2(Fig. 4b). We mapped over the parame-
ter space where Qc2ranges from 5 ×105to 6 ×105and
P2ranges from 20 mW to 30 mW. At the second har-
monic injection, the selection of P2is determined by the
critical power Pc, below which the steady-state solutions
of the system become unstable. This phenomenon has
been discussed in details by Drummond [56]. Here Pc=
14.05 mW such that P2should be larger than this value.
Within this region, small MDR (0 - 1.7 /h) is observed
(indicated by rainbow colors), showing that high sensi-
tivity is achieved for a sizable parameter range such that
the enhanced sensitivity is not an isolated singularity.
The lowest MDR (Ωmin <0.25 /h) appears in the nar-
row purple band. Outside this region, MDR gradually
increases as the color becomes more greenish and red-
dish, indicating reduced sensitivity. In principle, as the
rotation rate increases, the difference between the CW
and CCW mode becomes more significant. To bench-
mark the gyroscope performance, we compared the sen-
sitivity of our optical parametric oscillator (OPO) gyro
under the second harmonic injection (solid blue line) with
the sensitivity of a standard linear gyroscope (solid red
line) in Fig. 4(c). The optimal coupling factors for the
OPO gyro are Qc1= 1.018 ×105and Qc2= 5.462 ×105,
both discovered by Bayesian optimization. Meanwhile,
the highest sensitivity of the linear gyroscope is found at
Qc1= 9.58 ×106107. Note that, in the limit of van-
ishing β1, the lowest linear gyro MDR is achieved when
κ1=γ1(see Eq. 13). Since Qi1is fixed at 107, the op-
8
timal Qc1for the linear gyro is expected to be close but
not exactly equal to this value, considering the influence
of the small but non-zero back-scattering. In Fig. 4(d),
as the input power is increased from 20 mW to 30 mW,
the MDR of the OPO gyro drops from 0.42 /h down
to near zero and rises back to 1.21 /h with a local
minimum at 23.507 mW, corresponding to the optimal
sensitivity point (Ωmin = 4.465×103/h). To better vi-
sualize this local sensitivity minimum, a zoomed-in curve
is shown in the inset of Fig. 4(d). We also found that
the optimal-sensitivity point is associated with 9.9 dB
phase squeezing at an experimentally feasible value on a
TFLN platform [78]. Meanwhile, the MDR of the linear
gyroscope remains >0.49 /h, which demonstrates that
the OPO gyro is 124.4 ×more sensitive than the lin-
ear gyro under the same injection power, resonator size
and intrinsic quality factors. In order to visualize this
effect, we investigated the mean current values as a func-
tion of the rotation rate at different frequencies in Fig.
4(d). The mean values of the output differential currents
at the subharomonic (< i1>) and the second harmonic
(< i2>) are expressed by the solid blue and red line re-
spectively, while that of the linear gyroscope at the same
power consumption is expressed by the solid green line
for reference. It is shown that as increases from 0 - 100
/h, < i1>increases from 0 - 0.43 nA, < i2>increases
from 0 - 0.4 nA while the output current of the linear
gyroscope increases from 0 - 0.17 nA. This result shows
that the nonlinear eigenmode dispersion, as discussed in
Section II B, produces stronger output signals (differen-
tial currents) at both wavelengths compared to the linear
gyroscope.
B. Coherent injection at the fundamental
frequency
Aside from injecting light at the second harmonic,
we also investigated the scheme of the fundamental fre-
quency (subharmonic) injection. As shown in Fig. 5(a),
the input light at the wavelength λ1= 1590 nm is in-
jected from the waveguide ports bin
1,cw and bin
1,cw while
P2= 0, stimulating intra-cavity up-conversion (two pho-
tons of lower energy are combined to one photon of higher
energy). We study the sensitivity (MDR) in terms of P1
and Qc1, as indicated by the 2D density plot Fig. 5(b),
in which variations in MDR again show up in rainbow
colors. In Fig. 5(b) we mapped over the parameter
space where Qc1ranges from 6 ×106to 7 ×106and P1
ranges from 0.9µW to 1 µW . Note that much lower
power injection is required at this injection scheme com-
pared to coherent injection at the second harmonic, re-
lated to the fact that the steady-state solutions of the
cavity modes are stable only when P1is below the criti-
cal power when the second harmonic injection is absent
[56]. Here the critical power is Pc= 3.24 mW. Only
solutions of P1smaller than this value are stable, result-
ing in orders of magnitude lower power consumption. In
FIG. 4. (a) The schematic of the second harmonic injection
scheme. The input light is only injected at the second har-
monic (blue arrow). (b) 2D density plot of the sensitivity
at second harmonic injection. The sensitivity (MDR) is ex-
pressed as a function of the input power P2and the qual-
ity factor due to coupling loss of the second harmonic cavity
mode Qc2. (c) 1D linear plot of the sensitivity of the second
harmonic injection scheme (blue) and the standard linear gy-
roscope (red) at optimal Qcin terms of the input power. (d)
1D linear plot of the mean differential current of the out-
put light at fundamental frequency (blue), second harmonic
(green) and the standard linear gyroscope (red).
Fig. 5(b), MDR varies from 0 - 5 /h across the entire
parameter space. The lowest MDR is found in a nar-
row band-like region. The mean values of the output
currents are also studied here. To better evaluate the
gyroscope performance, we compared it with the linear
gyroscope at the same power consumption. Figure 5(c)
shows the sensitivity of the fundamental frequency injec-
tion (solid blue line) and the linear gyroscope (solid red
line) in terms of the injection power at fixed optimal Q
factors (Qc1= 6.747 ×106, Qc2= 6.675 ×107), which
are determined by Bayesian optimization, and optimal
Qc= 9.58 ×106in the linear gyro. As P1increases from
0.9µW to 1 µW , MDR drops from 0.46 /h down to
near zero at 0.945 µW , then rises back to 0.49 /h.
Meanwhile, the sensitivity of the linear gyroscope mono-
tonically but slowly decreases from 89.78 /h to
85.17 /h. The lowest MDR is found in a narrow band-
like region. The optimal sensitivity min = 0.093 /h
is found at P1= 0.945 µW . At the optimal point, a
surprisingly high sensitivity improvement of 942.5 ×
is observed (0.093 /h vs 87.62 /h). To better visual-
ize the sensitivity of the fundamental frequency injection
scheme, a magnified plot of the optimal sensitivity region
is shown in the inset of Fig. 5(d), clearly exhibiting the
local sensitivity minimum at 0.945 µW . We note that no
phase squeezing is observed in this scheme (although a
6.03 dB amplitude squeezing was observed at the second
harmonic (795 nm)); rather than squeezing, we posit that
9
the sensitivity improvement is largely due to nonlinear
eigenmode dispersion in the presence of critically sensi-
tive χ(2)-mediated wave mixing processes, not disimilar
to enlarged Sagnac shifts reported in dispersive materi-
als [94]. Most importantly, the fundamental frequency
injection scheme merges high sensitivity and low power
consumption together in a compact form, which shows
great potential in practical applications. Similar to the
analysis in Section III A, the mean differential currents
(< i1>,< i2>and linear) as a function of the rotation
rate are shown in Fig. 5(d) in solid blue, red and green
lines. As increases from 0 - 100 /h, < i1>increases
from 0 - 0.05 pA, < i2>increases from 0 - 0.0062 pA and
the output differential current of the linear gyroscope in-
creases from 0 - 0.0069 pA. The differential currents are
in the pA range, not the nA range, again because of much
weaker input power. It is shown that the sensitivity of
< i1>is almost 10×larger than that of < i2>and that
of the output current in the linear gyroscope, suggesting
that, even though very little second harmonic power is
ultimately extracted, the presence of a non-linearly in-
teracting second harmonic mode critically enhances the
sensitivity of the fundamental mode.
FIG. 5. (a) The schematic of the fundamental frequency (sub-
harmonic) injection scheme. The input light is only injected
at the fundamental frequency (red arrow). (b) 2D density
plot of the sensitivity at the fundamental frequency injection
scheme. The sensitivity (MDR) is expressed as a function of
the input power P1and the quality factor due to coupling loss
of the cavity mode Qc1. (c) 1D linear plot of the sensitivity
of the fundamental frequency injection scheme (blue) and the
standard linear gyroscope (red) at optimal Qcin terms of the
input power. The inset shows the magnified plot of the criti-
cal region. (d) 1D linear plot of the mean differential current
of the output light at fundamental frequency (blue), second
harmonic (green) and the standard linear gyroscope (red).
C. Dual frequency injection
For the sake of completeness, we also studied the dual
injection scheme (Fig. 6a) where coherent light is injected
at both fundamental and second harmonics (λ1= 1590
nm and λ2= 795 nm). Here, we optimize four indepen-
dent parameters: P1,P2,Qc1, and Qc2. The 2D density
plot of the sensitivity (MDR) in terms of P1and P2is
shown in Fig. 6(b). Here Qc1and Qc2are fixed at the
optimal values of 4.353×105and 8.769×106, again discov-
ered by Bayesian optimization. In Fig. 6(b), P1ranges
from 1 mW to 2 mW and P2ranges from 1 mW to 2
mW. Similar to the scheme of fundamental frequency in-
jection, we found that low input power is also necessary
for high sensitivity, which is beneficial for integrated op-
tical gyroscopes. As shown in the figure, MDR ranges
from 0 - 1.4 /h across the entire parameter space. The
region of the lowest MDR is expressed by a narrow purple
band in the figure while outside this region MDR grad-
ually increases. We also compare the MDR of the dual
injection scheme (solid blue line) with the linear gyro-
scope (solid red line), as shown in Fig. 6(c). Note that
in the figure the x-axis denotes the total input power P,
which equals to P1+P2for the dual frequency injection
and P1for the linear gyroscope. Here Qc1,Qc2and P2
are fixed at 4.353×105, 8.769×106and 1.873 mW, which
are the optimal values discovered by Bayesian optimiza-
tion. It is shown that the sensitivity of dual frequency
injection drops from 0.88/h down to near zero as P
increases from 3 mW to 3.373 mW, then rises back to
0.762 /h as Pincreases to 4 mW while sensitivity of
the linear gyroscope drops from 1.555 /h to 1.347 /h
within the same range of injection power. At the opti-
mal power (3.373 mW), a high sensitivity of 0.013 /h
is observed, leading to a substantial sensitivity improve-
ment of 113.1 ×over the linear gyroscope (0.013 /h
vs 1.47 /h). To better visualize this optimal point, a
magnified plot of the optimal sensitivity region is shown
in the inset of Fig. 6(c). To better understand the in-
fluence of rotation, the mean output currents are shown
in Fig. 6(d). As the rotation rate increases from 0
- 100 /h, < i1>,< i2>and the mean differential
current of the linear gyroscope monotonically increases
from 0 to 5.3 pA, 11.1 pA and 24.7 pA respectively. Note
that here neither the sub nor the second harmonic mode
shows stronger output current compared to the linear gy-
roscope. This is because the injection power of the linear
gyroscope equals to the summation of both the sub and
the second harmonic injection. Nonetheless, significant
sensitivity improvement is still observed in this case due
to the combination of the nonlinear coupling and the gen-
eration of the phase-squeezed photons.
D. Discussion
Table I summarizes the maximal sensitivity enhance-
ment factors (over the linear baseline) that can obtained
10
FIG. 6. (a) The schematic of the dual (sub-second harmonic)
injection scheme. The input light is only injected at both
the second (blue arrow) and the fundamental frequency (red
arrow). (b) 2D density plot of the sensitivity at the dual
injection scheme. The sensitivity (MDR) is expressed as a
function of the input power P1and P2. (c) 1D linear plot
of the sensitivity of the dual injection scheme (blue) and the
standard linear gyroscope (red) at optimal Qcin terms of the
input power. The inset shows the magnified plot of the critical
region. (d) 1D linear plot of the mean differential current of
the output light at the fundamental frequency (blue), second
harmonic (green) and the standard linear gyroscope (red).
in multiple operational regimes over a wide range of
critical power requirements. Under the optimal second-
harmonic injection at 23.5 mW, a 9.9 dB quadra-
ture squeezing is predicted, where our nonlinear multi-
resonant cavity quantum photonics gyro (or quantum-
optic nonlinear gyro QONG in short) can be nearly
124.4×more sensitive than an optimized linear gyro with
the same radius, intrinsic quality factor and power bud-
get, allowing for a minimum detectable rotation (MDR)
as small as 0.0044 /h. Alternatively, even larger en-
hancement factors can be obtained at smaller powers
under the fundamental and the dual frequency injection
schemes. The dual frequency injection scheme achieves
near 7 ×sensitivity improvement over the fundamen-
tal frequency injection scheme (0.013 /h vs 0.093 /h),
which could be the result of the extra squeezed photons
generated by the process of parametric down conversion.
In fact, these two latter scenarios do not necessarily pro-
mote squeezing but rely on nonlinear eigenmode disper-
sion due to critically-sensitive three-wave mixing dynam-
ics between the two resonances—an effect that resembles
the enlarged Sagnac shifts observed in suitably dispersive
materials [94]. In either of our fundamental or second-
harmonic injection scheme, we measured the output sig-
nals at both the fundamental and the second harmonic
frequencies in order to fully utilize the input pump power
(which gets converted into both harmonics), setting up
a fair comparison to a linear gyro under the same pump
TABLE I. Optimal sensitivities of various injection schemes
power. For a more conservative comparison, one may ar-
gue for using dual inputs and outputs in the linear case.
Aside from the fact that having to use two different fre-
quency lasers can be disadvantageous, a simple calcula-
tion readily shows that measuring two non-interacting
resonances in a linear gyro can offer only up to 2×
improvement (under the same power budgets)—in fact,
much less than 2 due to the smaller Qi2—highlighting
that nonlinear effects are indeed indispensable for sig-
nificant sensitivity enhancements. Most importantly, the
crucial insight we have drawn from our investigations is to
realize that multiple resonances in a nonlinear resonator
can be engineered to reinforce each other through nonlin-
ear wave mixing, and can be used as powerful degrees of
freedom to optimize sensitivities. This critical realization
suggests an exciting future direction: to generalize our
QONG from just two resonances to many more nonlin-
early interacting resonances (see also Section IV), which
may lead to even better sensitivities and functionalities
(approaching the ultimate Heisenberg limit).
IV. SUMMARY AND OUTLOOK
We have introduced a new type of quantum light gy-
roscopes based on nonlinear multi-resonant cavity quan-
tum photonics (quantum-optic nonlinear gyro or QONG
in short). Specifically, we analyzed and optimized the
quantum-enhanced gyroscopic sensitivity of a doubly res-
onant χ(2) cavity, revealing that, under quantum noise
conditions, 900×enhancement is possible over the
classical shot noise limit. We highlight that our cur-
rent design, which uses two resonances, represents only
an elementary step and a relatively simple example of
an QONG. In future works, we will develop a compre-
hensive QONG inertial sensing paradigm, where a syn-
ergistic amalgamation of both quadratic χ(2) and cu-
bic χ(3) nonlinearities, along with multiple intermixing
resonances, mutually reinforced Sagnac shifts, co-arising
quantum correlations, electro-optics dynamical control
and geometry-induced anomalous dispersion effects, can
unleash extraordinary complexities and freedoms, which
can be fully exploited by state-of-the-art optimization
techniques [9598] in order to identify unprecedented
regimes for gyroscopic operation and sensitivities. A
full incarnation of an QONG can be described by a
Heisenberg-Langevin system of the form (or an equiv-
11
alent density-operator Master equation [64]):
dˆaµ
j
dt =j+µ
j(Ω) κj
2γj
2ˆaµ
j+jˆaν
j
+X
klαβ
f(2)
klαβ ˆaα
k,ˆaβ
l,aα
k),aβ
l)
+X
klmαβθ
f(3)
klmαβθ ˆaα
k,ˆaβ
l,ˆaθ
m,aα
k),aβ
l),aθ
m)
+X
kqκc
jk ˆaµ
jk,in +X
kqγr
jk ˆηµ
jk (30)
for a selected set of carefully phase-matched and
dispersion-engineered resonances {ωj, j = 1, ..., N }.
Here, δµ(Ω), µ {cw,ccw}, is the rotation-dependent
Sagnac shift in the CW or CCW mode at each res-
onance. The functions f(2) and f(3) are polynomials
of the annihilation and creation operators, represent-
ing all possible quantum-coherent three-wave mixing and
four-wave mixing interactions between the selected res-
onances; these processes include sum and difference fre-
quency generations of different orders and combinations
as well as Kerr-variety self-phase and cross-phase mod-
ulation, and even cascaded processes [99]. It is impor-
tant to note that the strengths of different f(2) and f(3)
terms are determined by nonlinear coupling factors [55]
which characterize the field concentration and nonlin-
ear overlaps of the modes of the photonic resonator and
can be computed from nanophotonic simulations. There-
fore, on-chip structural parameters, ranging from a few
simple shape parameters to entire permittivity distribu-
tions, can serve as design degrees of freedom [100], by
which we can engineer and optimize the different non-
linear processes (e.g. their relative contributions). The
outputs of this multi-resonance system are collected by
multiple waveguide ports and are set to passively inter-
fere with each other and/or go through active electro-
optics pulse processing (readily achievable on a TFLN
platform [71]) before arriving at multiple photodetectors
to yield multiple photocurrent signals i={ˆ
i1, ...,ˆ
iM}.
From these multi-variable (vector-valued) measurements,
one can perform deep inferential analysis (such as ad-
vanced Bayesian computing [101]) to deduce the underly-
ing non-inertial motion; the sensitivity of the entire pro-
cess can be characterized by an end-to-end computation
of Fisher Information, which will serve as an optimization
figure of merit. We recognize tremendous opportunity in
analyzing and optimizing such a system with increasing
levels of mathematical and computational vigor, start-
ing from steady-state analysis, small-signal modeling,
classical stochastic simulations, to the non-perturbative
quantum phase-space apparatus involving positive P-
representations, Fokker-Planck equations and stochastic
calculus [47, 64], from few-parameter deterministic global
optimization [96, 102], multi-parameter Bayesian opti-
mization [42] and evolutionary algorithms [103], machine-
learning assisted hybrid optimization [104], and Monte
Carlo gradient computations [98] to billion-voxel topol-
ogy optimization [105] and full end-to-end inverse design
[106] of the entire workflow from the underlying resonator
geometry to multi-variable inferential processes. Exper-
imentally, thin film lithium niobate (TFLN) continues
to offer the most suitable platform which features state-
of-the-art on-chip frequency combs, pulse shaping, fre-
quency shifting and ultra-fast signal processing capabili-
ties [79, 107109].
ACKNOWLEDGMENTS
We thank Charles Roques-Carmes, Steven G. John-
son, Kiyoul Yang and Michael Larsen for informative
discussions. Financial support was provided by generous
gifts from the Virginia Tech Foundation. VK thanks the
late Stavros Katsanevas for pointing out the new runs of
VIRGO based on quantum mechanical states in Novem-
ber of 2018.
Appendix A: Steady state solutions
As discussed in Section II B, the nonlinear coupled
equations are solved by linearization. The classical scalar
valued amplitudes αare obtained from the steady state
analysis of Eqs. 14-17. To evaluate the steady state solu-
tions, the noise terms containing the quantum operators
are omitted. The equations are thus simplified as follows:
f1=κ1
2+γ1
21a1,cw 1a1,ccw
χa
1,cwa2,cw κ1bin
1,cw
(A1)
f2=κ1
2+γ1
2+1a1,ccw 1a1,cw
χa
1,ccwa2,ccw κ1bin
1,ccw
(A2)
f3=κ2
2+γ2
22a2,cw 2a2,ccw
+1
2χa2
1,cw κ2bin
2,cw
(A3)
f4=κ2
2+γ2
2+2a2,ccw 2a2,cw
+1
2χa2
1,ccw κ2bin
2,ccw
(A4)
Next, the steady state solutions are hence obtained by
solving the equations F(f1, f2, f3, f4) = 0. Aside from
the cavity modes an,cw/ccw, the input fields are also ex-
pressed as the steady states bn,cw/ccw . These values are
determined by the input power Pnat both waveguide
ports as discussed in Equation 3. Here we fix bnas
real values, which reduces the phase noise as reported
by Dowling [12]. At different injection schemes, different
arrangements of bnare employed. For example, b2= 0
at when the input light is injected at the fundamental
frequency and b1= 0 at the second harmonic injec-
tion. Though the steady state solutions of a similar sys-
tem has been studied by Drummond [56], in which the
12
analytical solutions are given at each injection scheme.
In our system, however, the rotation-induced frequency
shift δnand the Rayleigh back-scattering βnwhich in-
troduces cross-coupling between the CW and the CCW
modes make it impossible to calculate the analytical solu-
tion, hence we calculated the numerical solutions instead.
Since these equations are nonlinear equations, multiple
solutions are expected at each set of parameters. In or-
der to discover the steady state solutions, linear stability
analysis is performed [110, 111]. By checking the eigen
values of the Jacobian matrix [112] associated with each
set of solution we can determine the stability of these so-
lutions. The Jacobian matrix J of Eqs A1-A4 is obtained
by taking the gradient over a vector of the unknown vari-
ables an,cw/ccw:
J=F|an,cw/ccw (A5)
Note that in Equation A5 the function system Fand
the variables an,cw/ccw are both vectors, hence the gradi-
ent operator generates a matrix Jinstead of a vector.
When the real part of the eigen values of the matrix is
negative, the solution is stable and can be used for fur-
ther calculation.
Appendix B: The algebra of the quantum operators
The system is assumed to be quantum-limited, mean-
ing that the shot noise is considered as the main source
of noise. To this end, it is necessary to investigate the
statistical properties of the output light. As discussed in
Section II B, the output light are expressed by the op-
erators ˆ
bout
n,cw/ccw. As shown in Fig. 2, the output light
is measured by homodyne detection that ˆ
bout
n,cw and ˆ
bout
n,ccw
are coupled with each other before being detected by two
independent photodetectors:
ˆ
b1+ =ˆ
bout
1,cwe1+iˆ
bout
1,ccwe1/2 (B1)
ˆ
b1=iˆ
bout
1,cwe1+ˆ
bout
1,ccwe1/2 (B2)
ˆ
b2+ =ˆ
bout
2,cwe2+iˆ
bout
2,ccwe2/2 (B3)
ˆ
b2=iˆ
bout
2,cwe2+ˆ
bout
2,ccwe2/2 (B4)
Here ϕ1and ϕ2are the propagation phase shifts of the
output light at the sub and the second harmonics, which
can be arbitrarily selected. Here we set them to zero.
The resultant differential current current is given by:
ˆ
i1=Rω1ˆ
b
1+ˆ
b1+ ˆ
b
1ˆ
b1(B5)
ˆ
i2=Rω2ˆ
b
2+ˆ
b2+ ˆ
b
2ˆ
b2(B6)
Here R is the responsivity of the photodetectors, set as
0.58 A/W in our analysis. Setting A1=Rω1and A2=
Rω2, Eqs. B5-B6 can be further simplified as:
ˆ
i1=iA1ˆ
bout
1,cwˆ
bout
1,ccw ˆ
bout
1,ccwˆ
bout
1,cw(B7)
ˆ
i2=iA2ˆ
bout
2,cwˆ
bout
2,ccw ˆ
bout
2,ccwˆ
bout
2,cw(B8)
Following Maleki’s approach [7], the output operators are
linearized as ˆ
bout
n,cw/ccw =bout
n,cw/ccw +ˆ
δbout
n,cw/ccw. Here we
analyze the perturbation terms, such that Eqs. B7- B8
are simplified as:
ˆ
δi1=iA1bout
1,ccw ˆ
δbout
1,cw +bout
1,cw ˆ
δbout
1,ccw
bout
1,ccw ˆ
δbout
1,cw bout
1,cw ˆ
δbout
1,ccw
(B9)
ˆ
δi2=iA2bout
2,ccw ˆ
δbout
2,cw +bout
2,cw ˆ
δbout
2,ccw
bout
2,ccw ˆ
δbout
2,cw bout
2,cw ˆ
δbout
2,ccw
(B10)
Then quadrature basis expansion is performed to sepa-
rate the real and the imaginary parts (Xand Y) of the
operators:
bout
n,cw/ccw =Xout
n,cw/ccw +iY out
n,cw/ccw (B11)
bout
n,cw/ccw =Xout
n,cw/ccw iY out
n,cw/ccw (B12)
ˆ
δbout
n,cw/ccw =ˆ
δXout
n,cw/ccw +iˆ
δY out
n,cw/ccw (B13)
ˆ
δbout
n,cw/ccw =ˆ
δXout
n,cw/ccw iˆ
δY out
n,cw/ccw (B14)
Hence, Eqs. B9-B10 are converted to:
ˆ
δi1= 2A1Xout
1,ccw ˆ
δY out
1,cw Yout
1,ccw ˆ
δXout
1,cw
Xout
1,cw ˆ
δY out
1,ccw +Yout
1,cw ˆ
δXout
1,ccw(B15)
ˆ
δi2= 2A2Xout
2,ccw ˆ
δY out
2,cw Yout
2,ccw ˆ
δXout
2,cw
Xout
2,cw ˆ
δY out
2,ccw +Yout
2,cw ˆ
δXout
2,ccw(B16)
Next we need to determine the statistical properties of
ˆ
δi1and ˆ
δi2. Nevertheless, the output light is in com-
plex quantum states (squeezed vacuum/squeezed coher-
ent) which are difficult to calculate. On the other hand,
these quadrature operators are nothing but linear com-
binations of the input light which is in relatively simple
quantum states (vacuum/coherent). To this end, we cal-
culate the mean values and the variances from the input
states. Rewrite Eqs. B15 and B16 in the form below:
ˆ
δi1=
2
X
n=1
(b(1)
x,n,cw/ccwˆ
bin
X,n,cw/ccw +b(1)
y,n,cw/ccwˆ
bin
Y,n,cw/ccw
+c(1)
x,n,cw/ccwˆcin
X,n,cw/ccw +c(1)
y,n,cw/ccwˆcin
Y,n,cw/ccw)
(B17)
ˆ
δi2=
2
X
n=1
(b(2)
x,n,cw/ccwˆ
bin
X,n,cw/ccw +b(2)
y,n,cw/ccwˆ
bin
Y,n,cw/ccw
+c(2)
x,n,cw/ccwˆcin
X,n,cw/ccw +c(2)
y,n,cw/ccwˆcin
Y,n,cw/ccw)
(B18)
13
In Eqs. B17-B18, ˆ
bin
X,n,cw/ccw,ˆ
bin
Y,n,cw/ccw, ˆcin
X,n,cw/ccw
and ˆcin
Y,n,cw/ccw are the quadrature operators (real and
imaginary parts) of the injection light light ˆ
bin
n,cw/ccw
and the intrinsic loss channels ˆcin
n,cw/ccw for sub/second
(n=1,2) harmonic light, where ˆ
bin
X,n,cw/ccw,ˆ
bin
Y,n,cw/ccw are
in coherent states and ˆcin
X,n,cw/ccw and ˆcin
Y,n,cw/ccw are in
vacuum states. b(1,2)
x,n,cw/ccw,b(1,2)
y,n,cw/ccw,c(1,2)
x,n,cw/ccw and
c(1,2)
y,n,cw/ccw are the corresponding coefficients of these op-
erators at either sub or second harmonic. Assume ψ1/ψ2
and N1/N2are the initial phases and the numbers of the
injected photons of the input light ˆ
b1,cw/ccw and ˆ
b2,cw/ccw,
where N1=|bin
1|2and N2=|bin
2|2. With all these ingredi-
ents, now we can calculate mean values and the variances
of both differential currents. For a coherent state |α >,
the mean values of both quadrature operators and their
squares are given by [113]:
< α|ˆ
X|α >=Ncos ψ(B19)
< α|ˆ
Y|α >=Nsin ψ(B20)
< α|ˆ
X2|α >=4N(cos ψ)2+ 1
4(B21)
< α|ˆ
Y2|α >=4N(sin ψ)2+ 1
4(B22)
and the variances are defined as the mean values of the
square minus the square of the mean values of the quadra-
ture operators:
< α|ˆ
X2|α >=< α|ˆ
X2|α > (< α|ˆ
X|α >)2=1
4
(B23)
< α|ˆ
Y2|α >=< α|ˆ
Y2|α > (< α|ˆ
Y|α >)2=1
4
(B24)
Note that when two quadrature operators do not share
the same eigen vectors, the definitions are different. For
example, the inner product of a real quadrature opera-
tor at the second harmonic and an imaginary quadrature
operator at the fundamental frequency injection is given
by:
< α|ˆ
X2ˆ
Y1|α >=pN1pN2cos ψ2sin ψ2(B25)
< α|ˆ
X2ˆ
Y1|α > < α|ˆ
X1|α >< α|ˆ
Y1|α >= 0 (B26)
Now we can obtain the mean values and the variances of
ˆ
δi1and ˆ
δi2:
<ˆ
δi1>=
2
X
n=1b(1)
x,n,cw/ccwpNncos ψn
+b(1)
y,n,cw/ccwpNnsin ψn
(B27)
<ˆ
δi2>=
2
X
n=1b(2)
x,n,cw/ccwpNncos ψn
+b(2)
y,n,cw/ccwpNnsin ψn
(B28)
<ˆ
δi2
1>=1
4
2
X
n=1[(b(1)
x,n,cw/ccw)2+ (b(1)
y,n,cw/ccw)2
+ (c(1)
x,n,cw/ccw)2+ (c(1)
y,n,cw/ccw)2]
(B29)
<ˆ
δi2
2>=1
4
2
X
n=1[(b(2)
x,n,cw/ccw)2+ (b(2)
y,n,cw/ccw)2
+ (c(2)
x,n,cw/ccw)2+ (c(2)
y,n,cw/ccw)2]
(B30)
As discussed in Section II B, in order to determine the
covariance matrix, it is also necessary to calculate the
correlation between ˆ
δi1and ˆ
δi2following the rule of op-
erator calculation defined above:
<ˆ
δi1ˆ
δi2><ˆ
δi1>< ˆ
δi2>=
2
X
n=1b(1)
x,n,cw/ccwb(2)
x,n,cw/ccw
+b(1)
y,n,cw/ccwb(2)
y,n,cw/ccw
+c(1)
x,n,cw/ccwc(2)
x,n,cw/ccw
+c(1)
y,n,cw/ccwc(2)
y,n,cw/ccw
(B31)
<ˆ
δi2ˆ
δi1><ˆ
δi2>< ˆ
δi1>=<ˆ
δi1ˆ
δi2><ˆ
δi1>< ˆ
δi2>
(B32)
With everything discussed in this section, particularly
Eqs. B27-B32, now we can calculate Eqs. 23-25 to de-
termine the Fisher information and the corresponding
sensitivity of the system.
[1] U. S. G. A. Office, Defense navigation capabilities: Dod
is developing positioning, navigation, and timing tech-
nologies to complement gps (2021).
[2] G. Sagnac, L’ether lumineux demontre par l’effet du
vent relatif d’ether dans un interferometre en rotation
uniforme, Proceedings of the USSR Academy of Sciences
157, 708 (1913).
[3] E. J. Post, Sagnac effect, Reviews of Modern Physics
39, 475 (1967).
[4] H. J. Arditty and H. C. Lef`evre, Sagnac effect in fiber
gyroscopes, Opt. Lett. 6, 401 (1981).
[5] W. W. Chow, J. Gea-Banacloche, L. M. Pedrotti, V. E.
Sanders, W. Schleich, and M. O. Scully, The ring laser
gyro, Rev. Mod. Phys. 57, 61 (1985).
14
[6] J. D. Cresser, W. H. Louisell, P. Meystre, W. Schleich,
and M. O. Scully, Quantum noise in ring-laser gyros.
i. theoretical formulation of problem, Phys. Rev. A 25,
2214 (1982).
[7] A. B. Matsko, W. Liang, A. A. Savchenkov, V. S.
Ilchenko, and L. Maleki, Fundamental limitations of sen-
sitivity of whispering gallery mode gyroscopes, Physics
Letters A 382, 2289 (2018), special Issue in memory of
Professor V.B. Braginsky.
[8] Z. Wang, Y. Yang, P. Lu, R. Luo, Y. Li, D. Zhao,
C. Peng, and Z. Li, Dual-polarization interferometric
fiber-optic gyroscope with an ultra-simple configuration,
Opt. Lett. 39, 2463 (2014).
[9] H. Ma, W. Wang, Y. Ren, and Z. Jin, Low-noise low-
delay digital signal processor for resonant micro op-
tic gyro, IEEE Photonics Technology Letters 25, 198
(2013).
[10] G. A. Sanders, S. J. Sanders, L. K. Strandjord, T. Qiu,
J. Wu, M. Smiciklas, D. Mead, S. Mosor, A. Arrizon,
W. Ho, and M. Salit, Fiber optic gyro development
at Honeywell, in Fiber Optic Sensors and Applications
XIII , Vol. 9852, edited by E. Udd, G. Pickrell, and
H. H. Du, International Society for Optics and Pho-
tonics (SPIE, 2016) p. 985207.
[11] W. Z. Korth, A. Heptonstall, E. D. Hall, K. Arai, E. K.
Gustafson, and R. X. Adhikari, Passive, free-space het-
erodyne laser gyroscope, Classical and Quantum Grav-
ity 33, 035004 (2016).
[12] J. P. Dowling, Correlated input-port, matter-wave in-
terferometer: Quantum-noise limits to the atom-laser
gyroscope, Phys. Rev. A 57, 4736 (1998).
[13] W. Liang, V. S. Ilchenko, A. A. Savchenkov, E. Dale,
D. Eliyahu, A. B. Matsko, and L. Maleki, Resonant mi-
crophotonic gyroscope, Optica 4, 114 (2017).
[14] J. Zhang, H. Ma, H. Li, and Z. Jin, Single-polarization
fiber-pigtailed high-finesse silica waveguide ring res-
onator for a resonant micro-optic gyroscope, Opt. Lett.
42, 3658 (2017).
[15] P. P. Khial, A. D. White, and A. Hajimiri, Nanopho-
tonic optical gyroscope with reciprocal sensitivity en-
hancement, Nature Photonics 12, 671 (2018).
[16] Y.-H. Lai, M.-G. Suh, Y.-K. Lu, B. Shen, Q.-F. Yang,
H. Wang, J. Li, S. H. Lee, K. Y. Yang, and K. Vahala,
Earth rotation measured by a chip-scale ring laser gy-
roscope, Nature Photonics 14, 345 (2020).
[17] L. Ge, R. Sarma, and H. Cao, Rotation-induced evolu-
tion of far-field emission patterns of deformed microdisk
cavities, Optica 2, 323 (2015).
[18] J. Scheuer, Direct rotation-induced intensity modula-
tion in circular bragg micro-lasers, Opt. Express 15,
15053 (2007).
[19] C. Ciminelli, D. D’Agostino, G. Carnicella, F. Dell’Olio,
D. Conteduca, H. P. M. M. Ambrosius, M. K. Smit, and
M. N. Armenise, A high-q inp resonant angular velocity
sensor for a monolithically integrated optical gyroscope,
IEEE Photonics Journal 8, 1 (2016).
[20] M. J. Grant, P.-B. Vigneron, A. Feshali, W. Jin,
N. Abrams, M. Paniccia, and M. Digonnet, Chip-scale
gyroscope using silicon-nitride waveguide resonator with
a Q factor of 100 million, in Optical and Quantum Sens-
ing and Precision Metrology II, Vol. 12016, edited by
J. Scheuer and S. M. Shahriar, International Society for
Optics and Photonics (SPIE, 2022) p. 120160S.
[21] J. Ren, H. Hodaei, G. Harari, A. U. Hassan, W. Chow,
M. Soltani, D. Christodoulides, and M. Khajavikhan,
Ultrasensitive micro-scale parity-time-symmetric ring
laser gyroscope, Opt. Lett. 42, 1556 (2017).
[22] J. Scheuer and A. Yariv, Sagnac effect in coupled-
resonator slow-light waveguide structures, Phys. Rev.
Lett. 96, 053901 (2006).
[23] L. Yan, Z. Xiao, X. Guo, and A. Huang, Circle-
coupled resonator waveguide with enhanced Sagnac
phase-sensitivity for rotation sensing, Applied Physics
Letters 95, 10.1063/1.3243456 (2009), 141104.
[24] D. D. Smith, H. Chang, L. Arissian, and J. C. Diels,
Dispersion-enhanced laser gyroscope, Phys. Rev. A 78,
053824 (2008).
[25] M. S. Shahriar, G. S. Pati, R. Tripathi, V. Gopal,
M. Messall, and K. Salit, Ultrahigh enhancement in ab-
solute and relative rotation sensing using fast and slow
light, Phys. Rev. A 75, 053807 (2007).
[26] M. J. F. Digonnet and J. N. Chamoun, Recent develop-
ments in laser-driven and hollow-core fiber optic gyro-
scopes, in Fiber Optic Sensors and Applications XIII,
Vol. 9852, edited by E. Udd, G. Pickrell, and H. H. Du,
International Society for Optics and Photonics (SPIE,
2016) p. 985204.
[27] M. Larsen and M. Bulatowicz, Nuclear magnetic reso-
nance gyroscope: For darpa’s micro-technology for posi-
tioning, navigation and timing program, in 2014 Inter-
national Symposium on Inertial Sensors and Systems
(ISISS) (2014) pp. 1–5.
[28] G. oth and I. Apellaniz, Quantum metrology from
a quantum information science perspective, Journal of
Physics A: Mathematical and Theoretical 47, 424006
(2014).
[29] C. L. Degen, F. Reinhard, and P. Cappellaro, Quantum
sensing, Rev. Mod. Phys. 89, 035002 (2017).
[30] C. Roques-Carmes, Y. Salamin, J. Sloan, S. Choi,
G. Velez, E. Koskas, N. Rivera, S. E. Kooi, J. D.
Joannopoulos, and M. Soljaˇci´c, Biasing the quantum
vacuum to control macroscopic probability distribu-
tions, Science 381, 205 (2023).
[31] C. M. Caves, Quantum-mechanical noise in an interfer-
ometer, Phys. Rev. D 23, 1693 (1981).
[32] M. R. Grace, C. N. Gagatsos, Q. Zhuang, and
S. Guha, Quantum-enhanced fiber-optic gyroscopes us-
ing quadrature squeezing and continuous-variable en-
tanglement, Phys. Rev. Appl. 14, 034065 (2020).
[33] M. Fink, F. Steinlechner, J. Handsteiner, J. P. Dowl-
ing, T. Scheidl, and R. Ursin, Entanglement-enhanced
optical gyroscope, New Journal of Physics 21, 053010
(2019).
[34] A. et al. (Virgo Collaboration), Increasing the astro-
physical reach of the advanced virgo detector via the
application of squeezed vacuum states of light, Phys.
Rev. Lett. 123, 231108 (2019).
[35] T. et al., Quantum-enhanced advanced ligo detectors
in the era of gravitational-wave astronomy, Phys. Rev.
Lett. 123, 231107 (2019).
[36] W. Jin, Q.-F. Yang, L. Chang, B. Shen, H. Wang, M. A.
Leal, L. Wu, M. Gao, A. Feshali, M. Paniccia, K. J. Va-
hala, and J. E. Bowers, Hertz-linewidth semiconductor
lasers using cmos-ready ultra-high-q microresonators,
Nature Photonics 15, 346 (2021).
[37] I. Krasnokutska, J.-L. J. Tambasco, X. Li, and A. Pe-
ruzzo, Ultra-low loss photonic circuits in lithium niobate
15
on insulator, Opt. Express 26, 897 (2018).
[38] K. Liu, N. Jin, H. Cheng, N. Chauhan, M. W. Puck-
ett, K. D. Nelson, R. O. Behunin, P. T. Rakich, and
D. J. Blumenthal, Ultralow 0.034 db/m loss wafer-scale
integrated photonics realizing 720 million q and 380 µw
threshold brillouin lasing, Opt. Lett. 47, 1855 (2022).
[39] L. Li, H. Lin, Y. Huang, R.-J. Shiue, A. Yadav, J. Li,
J. Michon, D. Englund, K. Richardson, T. Gu, and
J. Hu, High-performance flexible waveguide-integrated
photodetectors, Optica 5, 44 (2018).
[40] M. Zhang, W. Sweeney, C. W. Hsu, L. Yang, A. D.
Stone, and L. Jiang, Quantum noise theory of excep-
tional point amplifying sensors, Phys. Rev. Lett. 123,
180501 (2019).
[41] D. Anderson, M. Shah, and L. Fan, Clarification of
the exceptional-point contribution to photonic sensing,
Phys. Rev. Appl. 19, 034059 (2023).
[42] B. Shahriari, K. Swersky, Z. Wang, R. P. Adams, and
N. de Freitas, Taking the human out of the loop: A re-
view of bayesian optimization, Proceedings of the IEEE
104, 148 (2016).
[43] M. Zhang, W. Sweeney, C. W. Hsu, L. Yang, A. Stone,
and L. Jiang, Quantum noise theory of exceptional point
amplifying sensors, Physical review letters 123, 180501
(2019).
[44] M. S. Stefszky, C. M. Mow-Lowry, S. S. Y. Chua,
D. A. Shaddock, B. C. Buchler, H. Vahlbruch, A. Kha-
laidovski, R. Schnabel, P. K. Lam, and D. E. McClel-
land, Balanced homodyne detection of optical quantum
states at audio-band frequencies and below, Classical
and Quantum Gravity 29, 145015 (2012).
[45] C. Ciminelli, F. Dell’Olio, and M. N. Armenise, High-q
spiral resonator for optical gyroscope applications: Nu-
merical and experimental investigation, IEEE Photonics
Journal 4, 1844 (2012).
[46] N. M. Kondratiev, V. E. Lobanov, A. E. Shitikov, R. R.
Galiev, D. A. Chermoshentsev, N. Y. Dmitriev, A. N.
Danilin, E. A. Lonshakov, K. N. Min’kov, D. M. Sokol,
S. J. Cordette, Y.-H. Luo, W. Liang, J. Liu, and I. A.
Bilenko, Recent advances in laser self-injection lock-
ing to high-q microresonators, Frontiers of Physics 18,
21305 (2023).
[47] P. D. Drummond, Squeezing with nonlinear optics, in
Quantum Squeezing (Springer Berlin Heidelberg, Berlin,
Heidelberg, 2004) pp. 99–139.
[48] H. Wang, Y.-H. Lai, Z. Yuan, M.-G. Suh, and K. Va-
hala, Petermann-factor sensitivity limit near an excep-
tional point in a brillouin ring laser gyroscope, Nature
Communications 11, 1610 (2020).
[49] Y. Zhao, Y. Okawachi, J. K. Jang, X. Ji, M. Lipson,
and A. L. Gaeta, Near-degenerate quadrature-squeezed
vacuum generation on a silicon-nitride chip, Phys. Rev.
Lett. 124, 193601 (2020).
[50] G. Masada, K. Miyata, A. Politi, T. Hashimoto, J. L.
O’Brien, and A. Furusawa, Continuous-variable entan-
glement on a chip, Nature Photonics 9, 316 (2015).
[51] C. Couteau, Spontaneous parametric down-conversion,
Contemporary Physics 59, 291 (2018).
[52] D. A. Kleinman, Theory of second harmonic generation
of light, Phys. Rev. 128, 1761 (1962).
[53] I. Breunig, Three-wave mixing in whispering gallery res-
onators, Laser and Photonics Reviews 10, 569 (2016).
[54] J. Lu, J. B. Surya, X. Liu, A. W. Bruch, Z. Gong, Y. Xu,
and H. X. Tang, Periodically poled thin-film lithium nio-
bate microring resonators with a second-harmonic gen-
eration efficiency of 250,000%/w, Optica 6, 1455 (2019).
[55] X. Guo, C.-L. Zou, and H. X. Tang, Second-
harmonic generation in aluminum nitride microrings
with 2500%/w conversion efficiency, Optica 3, 1126
(2016).
[56] P. Drummond, K. McNeil, and D. Walls, Non-
equilibrium transitions in sub/second harmonic genera-
tion, Optica Acta: International Journal of Optics 27,
321 (1980).
[57] J. R. Johansson, P. D. Nation, and F. Nori, Qutip: An
open-source python framework for the dynamics of open
quantum systems, Computer Physics Communications
183, 1760 (2012).
[58] W. Tang, F. Lee, and R. Ridley, Small-signal modeling
of average current-mode control, IEEE Transactions on
Power Electronics 8, 112 (1993).
[59] L. Gillner, G. Bj¨ork, and Y. Yamamoto, Quantum noise
properties of an injection-locked laser oscillator with
pump-noise suppression and squeezed injection, Phys.
Rev. A 41, 5053 (1990).
[60] S. Machida, Y. Yamamoto, and Y. Itaya, Observation of
amplitude squeezing in a constant-current–driven semi-
conductor laser, Phys. Rev. Lett. 58, 1000 (1987).
[61] Y. K. Chembo, Quantum dynamics of kerr optical fre-
quency combs below and above threshold: Spontaneous
four-wave mixing, entanglement, and squeezed states of
light, Physical Review A 93, 033820 (2016).
[62] S. Pontula, J. Sloan, N. Rivera, and M. Soljacic,
Strong intensity noise condensation using nonlinear dis-
persive loss in semiconductor lasers, arXiv preprint
arXiv:2212.07300 (2022).
[63] P. J. Costa, The hartman–grobman theorem, in Select
Ideas in Partial Differential Equations (Springer Inter-
national Publishing, Cham, 2021) pp. 173–198.
[64] P. D. Drummond and C. W. Gardiner, Generalised p-
representations in quantum optics, Journal of Physics
A: Mathematical and General 13, 2353 (1980).
[65] S. L. Braunstein and C. M. Caves, Statistical distance
and the geometry of quantum states, Phys. Rev. Lett.
72, 3439 (1994).
[66] P. Drummond, K. McNeil, and D. Walls, Non-
equilibrium transitions in sub/second harmonic genera-
tion, Optica Acta: International Journal of Optics 28,
211 (1981).
[67] R. Kapral, Quantum dynamics in open quantum-
classical systems, Journal of Physics: Condensed Matter
27, 073201 (2015).
[68] M. Giglio, P. Patimisco, A. Sampaolo, G. Scamarcio,
F. K. Tittel, and V. Spagnolo, Allan deviation plot
as a tool for quartz-enhanced photoacoustic sensors
noise analysis, IEEE Transactions on Ultrasonics, Fer-
roelectrics, and Frequency Control 63, 555 (2016).
[69] J. H. Shapiro, The quantum theory of optical commu-
nications, IEEE Journal of Selected Topics in Quantum
Electronics 15, 1547 (2009).
[70] H. J. Carmichael, Dissipation in quantum mechanics:
The master equation approach, in Statistical Methods
in Quantum Optics 1: Master Equations and Fokker-
Planck Equations (Springer Berlin Heidelberg, Berlin,
Heidelberg, 1999) pp. 1–28.
[71] D. Zhu, L. Shao, M. Yu, R. Cheng, B. Desiatov,
C. J. Xin, Y. Hu, J. Holzgrafe, S. Ghosh, A. Shams-
Ansari, E. Puma, N. Sinclair, C. Reimer, M. Zhang, and
16
M. Lonˇcar, Integrated photonics on thin-film lithium
niobate, Adv. Opt. Photon. 13, 242 (2021).
[72] C. Cui, L. Zhang, and L. Fan, In situ control of effec-
tive kerr nonlinearity with pockels integrated photonics,
Nature Physics 18, 497 (2022).
[73] J. M. Silver, L. Del Bino, M. T. Woodley, G. N. Gha-
lanos, A. Ø. Svela, N. Moroney, S. Zhang, K. T. Grat-
tan, and P. Del’Haye, Nonlinear enhanced microres-
onator gyroscope, Optica 8, 1219 (2021).
[74] K. J. Peters and S. R. Rodriguez, Exceptional precision
of a nonlinear optical sensor at a square-root singularity,
Physical Review Letters 129, 013901 (2022).
[75] J. Pastrˇak and L. Roskovcov´a, Refraction index mea-
surements on aln single crystals, physica status solidi
(b) 14, K5 (1966).
[76] S. Singh, J. R. Potopowicz, L. G. Van Uitert, and S. H.
Wemple, Nonlinear Optical Properties of Hexagonal Sil-
icon Carbide, Applied Physics Letters 19, 53 (2003).
[77] D. E. Aspnes, S. M. Kelso, R. A. Logan, and R. Bhat,
Optical properties of alxga1x as, Journal of Applied
Physics 60, 754 (1986).
[78] R. Nehra, R. Sekine, L. Ledezma, Q. Guo, R. M. Gray,
A. Roy, and A. Marandi, Few-cycle vacuum squeezing
in nanophotonics, Science 377, 1333 (2022).
[79] M. W. McCutcheon, D. E. Chang, Y. Zhang, M. D.
Lukin, and M. Lonˇcar, Broadband frequency conversion
and shaping of single photons emitted from a nonlinear
cavity, Opt. Express 17, 22689 (2009).
[80] P.-K. Chen, I. Briggs, S. Hou, and L. Fan, Ultra-
broadband quadrature squeezing with thin-film lithium
niobate nanophotonics, Opt. Lett. 47, 1506 (2022).
[81] J. Zhao, C. Ma, M. R¨using, and S. Mookherjea, High
quality entangled photon pair generation in periodically
poled thin-film lithium niobate waveguides, Phys. Rev.
Lett. 124, 163603 (2020).
[82] M. Zhang, C. Wang, R. Cheng, A. Shams-Ansari, and
M. Lonˇcar, Monolithic ultra-high-q lithium niobate mi-
croring resonator, Optica 4, 1536 (2017).
[83] Z.-F. Bi, A. W. Rodriguez, H. Hashemi, D. Duch-
esne, M. Loncar, K.-M. Wang, and S. G. Johnson,
High-efficiency second-harmonic generation in doubly-
resonant χ(2) microring resonators, Opt. Express 20,
7526 (2012).
[84] A. Shams-Ansari, M. Yu, Z. Chen, C. Reimer,
M. Zhang, N. Picqu´e, and M. Lonˇcar, Thin-film lithium-
niobate electro-optic platform for spectrally tailored
dual-comb spectroscopy, Communications Physics 5, 88
(2022).
[85] K. Nakamura, J. Kurz, K. Parameswaran, and
M. M. Fejer, Periodic poling of magnesium-oxide-doped
lithium niobate, Journal of Applied Physics 91, 4528
(2002).
[86] A. Rantamaki, E. Saarinen, J. Lyytik¨ainen, J. Kontio,
J. Heikkinen, K. Lahtonen, M. Valden, and O. Okhot-
nikov, Towards high power flip-chip long-wavelength
semiconductor disk lasers, in Vertical External Cav-
ity Surface Emitting Lasers (VECSELs) V , Vol. 9349,
edited by M. Guina, International Society for Optics
and Photonics (SPIE, 2015) p. 934908.
[87] X. Guo, L. Shao, L. He, K. Luke, J. Morgan, K. Sun,
J. Gao, T.-C. Tzu, Y. Shen, D. Chen, B. Guo, F. Yu,
Q. Yu, M. Jafari, M. Lonˇcar, M. Zhang, and A. Beling,
High-performance modified uni-traveling carrier photo-
diode integrated on a thin-film lithium niobate platform,
Photon. Res. 10, 1338 (2022).
[88] A. Guarino, G. Poberaj, D. Rezzonico,
R. Degl’Innocenti, and P. unter, Electro–optically
tunable microring resonators in lithium niobate, Nature
Photonics 1, 407 (2007).
[89] M. Wang, J. Li, H. Yao, X. Li, J. Wu, K. S. Chiang,
and K. Chen, Thin-film lithium-niobate modulator with
a combined passive bias and thermo-optic bias, Opt.
Express 30, 39706 (2022).
[90] R. Lu, Y. Yang, M.-H. Li, and S. Gong, Ghz low-
loss acoustic rf couplers in lithium niobate thin film,
IEEE Transactions on Ultrasonics, Ferroelectrics, and
Frequency Control 67, 1448 (2020).
[91] A. Shams-Ansari, G. Huang, L. He, Z. Li, J. Holz-
grafe, M. Jankowski, M. Churaev, P. Kharel, R. Cheng,
D. Zhu, N. Sinclair, B. Desiatov, M. Zhang, T. J.
Kippenberg, and M. Lonˇcar, Reduced material loss in
thin-film lithium niobate waveguides, APL Photonics 7,
081301 (2022).
[92] D. Bertsimas and J. Tsitsiklis, Simulated Annealing,
Statistical Science 8, 10 (1993).
[93] A. Eiben, R. Hinterding, and Z. Michalewicz, Parameter
control in evolutionary algorithms, IEEE Transactions
on Evolutionary Computation 3, 124 (1999).
[94] D. D. Smith, H. Chang, L. Arissian, and J. C. Diels,
Dispersion-enhanced laser gyroscope, Phys. Rev. A 78,
053824 (2008).
[95] S. H. Haji and A. M. Abdulazeez, Comparison of op-
timization techniques based on gradient descent algo-
rithm: A review, PalArch’s Journal of Archaeology of
Egypt / Egyptology 18, 2715 (2021).
[96] C. A. Floudas, Deterministic global optimization: the-
ory, methods and applications, Vol. 37 (Springer Science
and Business Media, 2013).
[97] A. G. Baydin, B. A. Pearlmutter, A. A. Radul, and J. M.
Siskind, Automatic differentiation in machine learning:
a survey, Journal of Marchine Learning Research 18, 1
(2018).
[98] S. Mohamed, M. Rosca, M. Figurnov, and A. Mnih,
Monte carlo gradient estimation in machine learning,
Journal of Machine Learning Research 21 (2020).
[99] L. Misoguti, S. Backus, C. G. Durfee, R. Bartels, M. M.
Murnane, and H. C. Kapteyn, Generation of broadband
vuv light using third-order cascaded processes, Phys.
Rev. Lett. 87, 013601 (2001).
[100] S. Molesky, Z. Lin, A. Y. Piggott, W. Jin, J. Vuckovi´c,
and A. W. Rodriguez, Inverse design in nanophotonics,
Nature Photonics 12, 659 (2018).
[101] K. Csill´ery, M. G. Blum, O. E. Gaggiotti, and
O. Fran¸cois, Approximate bayesian computation (abc)
in practice, Trends in ecology and evolution 25, 410
(2010).
[102] R. Horst and H. Tuy, Global Optimization: Determinis-
tic Approaches (Springer Berlin Heidelberg, 2013).
[103] T. ack and H.-P. Schwefel, An overview of evolution-
ary algorithms for parameter optimization, Evolution-
ary Computation 1, 1 (1993).
[104] Q. Wu, Y. Cao, H. Wang, and W. Hong, Machine-
learning-assisted optimization and its application to an-
tenna designs: Opportunities and challenges, China
Communications 17, 152 (2020).
[105] N. Aage, E. Andreassen, B. S. Lazarov, and O. Sigmund,
Giga-voxel computational morphogenesis for structural
design, Nature 550, 84 (2017).
17
[106] Z. Lin, R. Pestourie, C. Roques-Carmes, Z. Li, F. Ca-
passo, M. Soljaˇci´c, and S. G. Johnson, End-to-end
metasurface inverse design for single-shot multi-channel
imaging, Opt. Express 30, 28358 (2022).
[107] C. Wang, M. Zhang, M. Yu, R. Zhu, H. Hu, and M. Lon-
car, Monolithic lithium niobate photonic circuits for
kerr frequency comb generation and modulation, Na-
ture Communications 10, 978 (2019).
[108] M. Yu, D. Barton III, R. Cheng, C. Reimer, P. Kharel,
L. He, L. Shao, D. Zhu, Y. Hu, H. R. Grant, L. Jo-
hansson, Y. Okawachi, A. L. Gaeta, M. Zhang, and
M. Lonˇcar, Integrated femtosecond pulse generator on
thin-film lithium niobate, Nature 612, 252 (2022).
[109] J. Lu, D. N. Puzyrev, V. V. Pankratov, D. V. Skryabin,
F. Yang, Z. Gong, J. B. Surya, and H. X. Tang, Two-
colour dissipative solitons and breathers in microres-
onator second-harmonic generation, Nature Communi-
cations 14, 2798 (2023).
[110] T. Erneux and P. Glorieux, Laser Dynamics (Cam-
bridge University Press, 2010).
[111] A. Gavrielides, V. Kovanis, and T. Erneux, Analytical
stability boundaries for a semiconductor laser subject
to optical injection, Optics Communications 136, 253
(1997).
[112] H. Sayama, Introduction to the modeling and analysis
of complex systems (2015).
[113] M. O. Scully and M. S. Zubairy, Quantum Optics (Cam-
bridge University Press, 1997).
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
The stabilization and manipulation of laser frequency by means of an external cavity are nearly ubiquitously used in fundamental research and laser applications. While most of the laser light transmits through the cavity, in the presence of some back-scattered light from the cavity to the laser, the self-injection locking effect can take place, which locks the laser emission frequency to the cavity mode of similar frequency. The self-injection locking leads to dramatic reduction of laser linewidth and noise. Using this approach, a common semiconductor laser locked to an ultrahigh- Q microresonator can obtain sub-Hertz linewidth, on par with state-of-the-art fiber lasers. Therefore it paves the way to manufacture high-performance semiconductor lasers with reduced footprint and cost. Moreover, with high laser power, the optical nonlinearity of the microresonator drastically changes the laser dynamics, offering routes for simultaneous pulse and frequency comb generation in the same microresonator. Particularly, integrated photonics technology, enabling components fabricated via semiconductor CMOS process, has brought increasing and extending interest to laser manufacturing using this method. In this article, we present a comprehensive tutorial on analytical and numerical methods of laser self-injection locking, as well a review of most recent theoretical and experimental achievements.
Article
Full-text available
Frequency conversion of dissipative solitons associated with the generation of broadband optical frequency combs having a tooth spacing of hundreds of giga-hertz is a topical challenge holding the key to practical applications in precision spectroscopy and data processing. The work in this direction is underpinned by fundamental problems in nonlinear and quantum optics. Here, we present the dissipative two-colour bright-bright and dark-dark solitons in a quasi-phase-matched microresonator pumped for the second-harmonic generation in the near-infrared spectral range. We also found the breather states associated with the pulse front motion and collisions. The soliton regime is found to be typical in slightly phase-mismatched resonators, while the phase-matched ones reveal broader but incoherent spectra and higher-order harmonic generation. Soliton and breather effects reported here exist for the negative tilt of the resonance line, which is possible only via the dominant contribution of second-order nonlinearity.
Article
Full-text available
Integrated femtosecond pulse and frequency comb sources are critical components for a wide range of applications, including optical atomic clocks1, microwave photonics2, spectroscopy3, optical wave synthesis4, frequency conversion5, communications6, lidar7, optical computing8 and astronomy9. The leading approaches for on-chip pulse generation rely on mode-locking inside microresonators with either third-order nonlinearity10 or with semiconductor gain11,12. These approaches, however, are limited in noise performance, wavelength and repetition rate tunability 10,13. Alternatively, subpicosecond pulses can be synthesized without mode-locking, by modulating a continuous-wave single-frequency laser using electro-optic modulators1,14–17. Here we demonstrate a chip-scale femtosecond pulse source implemented on an integrated lithium niobate photonic platform18, using cascaded low-loss electro-optic amplitude and phase modulators and chirped Bragg grating, forming a time-lens system19. The device is driven by a continuous-wave distributed feedback laser chip and controlled by a single continuous-wave microwave source without the need for any stabilization or locking. We measure femtosecond pulse trains (520-femtosecond duration) with a 30-gigahertz repetition rate, flat-top optical spectra with a 10-decibel optical bandwidth of 12.6 nanometres, individual comb-line powers above 0.1 milliwatts, and pulse energies of 0.54 picojoules. Our results represent a tunable, robust and low-cost integrated pulsed light source with continuous-wave-to-pulse conversion efficiencies an order of magnitude higher than those achieved with previous integrated sources. Our pulse generator may find applications in fields such as ultrafast optical measurement19,20 or networks of distributed quantum computers21,22. A femtosecond pulse generator is realized using an electro-optic time-lens system integrated on a lithium niobate photonic chip, capable of tunable repetition rates and wavelengths.
Article
Full-text available
It is essential to bias a thin-film lithium-niobate Mach-Zehnder electro-optic (EO) modulator at the desired operation condition to ensure optimal performance of the modulator. While thermo-optic (TO) control can solve the problem of bias drift, it consumes significant electric power. In this paper, we propose a technique to largely reduce bias power consumption by combining passive bias and TO bias. In our design, waveguide sections with different widths are introduced in the two arms of the MZ modulator to produce a desired phase difference of π/2 rad (the desired operation condition), and local heating with electrode heaters placed on the waveguides is employed to provide compensation for any phase drift caused by fabrication errors and other effects. As the TO control only serves to compensate for small errors, the electric power required is low and the response is fast. To demonstrate our technique experimentally, we fabricate several modulators of the same design on the same chip. Our experimental modulators can operate up to ∼40 GHz with a half-wave voltage of ∼2.0 V over a wide optical bandwidth, and the performances are insensitive to ambient temperature variations. The TO bias powers required range from 1 mW to 15 mW, and the thermal rise and fall times are 47 µs and 14 µs, respectively. Our technique can facilitate the development of practical high-speed EO modulators on the lithium-niobate-on-insulator platform.
Article
Full-text available
Thin-film lithium niobate has shown promise for scalable applications ranging from single-photon sources to high-bandwidth data communication systems. Realization of the next generation high-performance classical and quantum devices, however, requires much lower optical losses than the current state of the art resonator ( Q-factor of ∼10 million). Yet the material limitations of ion-sliced thin film lithium niobate have not been explored; therefore, it is unclear how high the quality factor can be achieved in this platform. Here, using our newly developed characterization method, we find out that the material limited quality factor of thin film lithium niobate photonic platform can be improved using post-fabrication annealing and can be as high as Q ≈ 1.6 × 10 ⁸ at telecommunication wavelengths, corresponding to a propagation loss of 0.2 dB/m.
Article
Quantum field theory suggests that electromagnetic fields naturally fluctuate, and these fluctuations can be harnessed as a source of perfect randomness. Many potential applications of randomness rely on controllable probability distributions. We show that vacuum-level bias fields injected into multistable optical systems enable a controllable source of quantum randomness, and we demonstrated this concept in an optical parametric oscillator (OPO). By injecting bias pulses with less than one photon on average, we controlled the probabilities of the two possible OPO output states. The potential of our approach for sensing sub-photon-level fields was demonstrated by reconstructing the temporal shape of fields below the single-photon level. Our results provide a platform to study quantum dynamics in nonlinear driven-dissipative systems and point toward applications in probabilistic computing and weak field sensing.
Article
Exceptional points, with simultaneous coalescence of eigenvalues and eigenvectors, can be realized with non-Hermitian photonic systems. With the enhanced response, exceptional points have been proposed to improve the performance of photonic sensing. Recently, there are intense debates about the actual sensing advantage of exceptional points. The major concern is that intrinsic noise is also increased at exceptional points. Here, we aim to clarify the contribution of exceptional points for photonic sensing. This is achieved by analyzing the condition to realize divergent quantum Fisher information in linear non-Hermitian photonic systems. We show that the divergence of quantum Fisher information is the result of lasing threshold, instead of exceptional points. However, exceptional points correspond to the condition that lasing threshold is simultaneously achieved across multiple photonic modes. Therefore, exceptional points can further improve the sensitivity on top of lasing threshold. On the other hand, exceptional points alone cannot provide sensing advantage.
Article
One of the most fundamental quantum states of light is the squeezed vacuum, in which noise in one of the quadratures is less than the standard quantum noise limit. In nanophotonics, it remains challenging to generate, manipulate, and measure such a quantum state with the performance required for a wide range of scalable quantum information systems. Here, we report the development of a lithium niobate–based nanophotonic platform to demonstrate the generation and all-optical measurement of squeezed states on the same chip. The generated squeezed states span more than 25 terahertz of bandwidth supporting just a few optical cycles. The measured 4.9 decibels of squeezing surpass the requirements for a wide range of quantum information systems, demonstrating a practical path toward scalable ultrafast quantum nanophotonics.
Article
Exceptional points (EPs)-spectral singularities of non-Hermitian linear systems-have recently attracted interest for sensing. While initial proposals and experiments focused on enhanced sensitivities neglecting noise, subsequent studies revealed issues with EP sensors in noisy environments. Here we propose a single-mode Kerr-nonlinear resonator for exceptional sensing in noisy environments. Based on the resonator's dynamic hysteresis, we define a signal that displays a square-root singularity reminiscent of an EP. However, our sensor has crucial fundamental and practical advantages over EP sensors: the signal-to-noise ratio increases with the measurement speed, the square-root singularity is easily detected through intensity measurements, and both sensing precision and information content of the signal are enhanced around the singularity. Our sensor also overcomes the fundamental trade-off between precision and averaging time characterizing all linear sensors. All these unconventional features open up new opportunities for fast and precise sensing using hysteretic resonators.
Chapter
By generalizing the Hartman–Grobman Theorem for the flows of vector fields, it is seen that a class of nonlinear evolution equations admit a local linearization. Under the restriction that the nonlinear part of the solution operator is compact, the linearization turns out to be global. This local linearization is applied to the periodic Korteweg de Vries, Burgers’, and cubic Schrödinger equations.