You are on page 1of 17

Atomic And Molecular Physics

GROUP MEMBERS : ROLL NO's

Mahnoor Asif : 02

Bushra Yaqoob : 06

Hira Saadat : 10

Aleena Afzal : 18

Farwa Inaam : 30

Sehrish Munawar : 33

Bela Ghumman : 36

Areej Liaqat : 42

Tamania Naeem : 116

Course Code : Phys-401

Semester : 8th(A)

Submitted to : Mam Rahat Batool

GCWUS
All-optical link for direct comparison
of distant optical clocks
Abstract:
To make remote comparisons of two distant ultra-stable optical clocks we developed an all-
optical link system. To compensate for the phase noise accumulated during propagation
through a fiber link, an optical carrier transfer system based on a fiber interferometer was
employed. In a 90-km link, transfer stabilities of 2 ranges10−15 in 1 second and 4 ranges10−18
in 1000 seconds were achieved. In addition, an active polarization control system was
introduced to maintain the transmitted light in an appropriate polarization and, as a result, a
stable and reliable comparison was made. The instability of the all-optical link system, including
those of the erbium doped fiber amplifiers (EDFAs) free of phase-noise compensation, was less
than 2×10−15 at 1 second and 7×10−17 at 1000 secondsThe system was available via an urban
fiber link of 60 km for a direct comparison of two distant 87Sr lattice clocks. This technique will
be essential for the measurement of the reproducibility of optical frequency standards.

1. Introduction:

Comparisons of Clock are necessary for confirmation of reproducibility of frequency standards.


By utilizing satellite connections such as GPS or two-way satellite time and frequency transfer
(TWSTF), it has been possible to adjust the frequencies of microwave atomic clocks. One of
the applications of these procedures is the system of international atomic time (TAI) links built
up by Bureau International des Poidset Mesures (BIPM), and it permits us to know time
difference between clocks isolated by some separation [1]. It is given by BIPM that typical
satellite connections utilizing GPS exact point positioning and TWSTFT have stabilities of optical
frequency standards have quickly expanded, leading to 1×10−16for an averaging time of 1000
seconds [2].This progress has prompted the need for a strategy to analyze frequencies without
degrading stabilities. Many experiments have been performed over a short distance link [6] and
with a spooled fiber [7]. Different studies include ones for transfer of a clock signal along with
data traffic [8, 9], utilizing a high gain amplifier for a long stretch link up to 480 km [10], and
building up a repeater [9]. For observation of relative frequency stability of optical clocks and
cavity stabilized lasers [11–13], many different carrier techniques have been established. We
built up an all-optical connection system that comprises of Ti:sapphire (Ti:S) frequency brushes,
nonlinear crystals for frequency multiplying, fiber amplifiers, a 1.5 μm stable laser, an optical
transporter carrier system, and a functioning/active polarization control system. The system
can perform solid estimations and can work for quite a while free from large polarization
varieties. Its achievability was affirmed in an examination that directly compared two Sr clocks
situated at the National Institute of Information and Communications Technology (NICT) and
the University Of Tokyo (UT). In this paper the details and performance of systems are included
along with direct comparison of distant optical clocks.

2. Overall system:

Figure 1 is an all-optical link-system diagram. A secure optical frequency is sent over an optical
fiber connection from the local site to the remote site. The telecom wavelength at 1.5 μm is
appropriate for transmission over optical fiber connections from a low optical loss viewpoint,
while most optical clock transitions are beyond the range of visible light Therefore, frequency
conversion between the two bands is necessary, and for this reason, the device uses two Ti: S
frequency combs and two periodically poled lithium niobate (PPLN). At both sites the clock
lasers are hooked up to an optical transfer of atoms or ions. At the local site the Ti: S frequency
comb is phaselocked to the laser clock An external-cavity laser diode (ECDL) operating at 1.5
μm is phase-locked to the PPLN-generated frequency comb via its second harmonic generation
(SHG) light. The light emitted from the ECDL is amplified by an EDFA booster and an optical
carrier transfer system, listed below, provides the light to the remote location .At the remote site
an EDFA amplifies the transferred light to feed ample power into the PPLN there. The
polarization state (SOP) is controlled by the control mechanism for polarization, and the light is
increased in frequency. At the remote site the Ti: S frequency comb is phase-locked to the SHG
sun. The beat signal between the laser clock and the closest comb part is the difference frequency
and the relative stability of the two optical clocks,

3. Subsystem:

3.1 Optical carrier transfer system:

Figure 2 displays a graphical representation of the optical carrier transfer network that we
evaluated for the ability to cancel fiber noise. We developed a noise cancelation system based
on a fiber interferometer to counteract the accumulated phase noise during the transmission.
The definition is similar to the scheme shown by Ma et al.[6]. A limited-bandwidth 1.5μm light
to be conveyed is bound to a fiber-pigtailed acousto-optic modulator 1 (AOM1) powered by a
100-MHz voltage regulated oscillator (VCO) by an optical circulator 1 (OC1) and transmitted via
a long-haul connection to the remote site. The transmitted light is connected to the second
fiber pigtailed AOM 2 (AOM2) at the remote location, pushed by a 55-MHz stable reference and
magnified by a uni-directional EDFA. A user is served with portion of the amplified light
(referred to as the "out-of-loop signal"), and the remaining is sent back to the local site via
optical circulator 2 (OC2) and AOM2. AOM2 segregates between the reflected signal at
connectors and splices from the stray simulations. One problem when using the uni-directional
EDFA is that the device compensates for only half of the noise caused by it. As previously
highlighted, we discovered that the left portion of the noise didn't restrict our network
performance. Usually, bi-directional EDFAs and Faraday rotator mirrors (FRMs) are used to
accommodate for optical failure and to reflect part of the light transmitted back to the local
site. Evidently, using a spooled fiber or a dedicated fiber connection, the combination of the bi-
directional EDFA and FRM worked properly in performance tests. Nonetheless, also in terms of
applying a bi-directional EDFA to the NICT-UT fiber connection, unnecessary back-reflections
from several remote site SC / PC connectors caused excessive input to the bi-directional EDFA,
resulting in saturation of the signal. The light returning to the local site is combined the AOM1
and OC1 with a photo detector 1 (PD1). AOM1 and AOM2 used diffraction — 1st and +1st
separately. The return light frequency that moves twice the AOM1 and AOM2 is changed by -90
MHz. The shift in polarization in the moving light is an significant problem within the long-haul
connection. The FRM re-reflection allows the difference of the polarization state (SOP) between
the reference and come back light to be maintained. In this situation isolated from the FRM, a
balanced photo detector differentially senses the return light to reduce the 90-MHz beat
signal's amplitude variance. The beat signal obtained by a direct digital synthesizer (DDS) is
divided-by-50 to 1.8 MHz. The DDS does not only work to extend the phase lock capture range
for control of an immense amount of phase noise and also to make the beat signal amplitude
even. The corresponding divided pulse is combined with a secure 1.8-MHz reference connected
to a Hydrogen maser. The output of the mixer is enhanced, processed, and refed back to the
VCO. Thus, the VCO step is balanced to reduce the noise caused by fiber.

3.2 JGN2plus Optical Fiber link:


NICT works on an optical fiber system test bed titled JGN2+, as framework for analysis &
establishment on IC technology [14]. Particular segments of single mode dark Fibre merged
when a fragment of the test bed coupled with NICT and Otemachi (a business district in tokyo)
which is 45 kilometer away. To identify the condition of noise cancellation procedure, 2
matching connections utilized by us, and then fixed them at Otemachi so that the regional and
distant areas of a ninety -km affiliation are both at NICT. Figure 3 demonstrates an illustrative
picture of link. The optical loss is minus thirty dB for circular-trip link. Fig. 4 portrays the strong
spectral density of phase noise obtrude on the 1.5 micrometer light progressing in the disc like
trip 90-kilometers optical fiber loop. Noise of this type is much greater as compared to other
noises in optical fibril loop [15, 16]. e.g, Terra etal. [10] described phase noise that is around
200 rad2/hz at a Fourier frequency of one Hertz in a 480km link, and it is 30 dB smaller than
that in the JGN2+ association. Huge quantity of noise is may be due to approximately 50
percent of the link b/w NICT & Otemachi being embedded across the underpass line and about
1/3 being freaky in the air. This prediction is therefore by the factor that the phase noise grows
on breezy days and reduces at night-time when the tunnel is out of facility.Extending to the
SOP of the transmitted light alters actively during the transfer for the reason that distortion of
the optical fiber core due to Temperature( T)and Pressure( P) changes. Stable

transport of the freq std. wave needs not only remuneration of a big proportion of phase-
noise but also dynamic polarization charge. NIC technology was affixed to UT by a fiber-link
among entire length of sixty-km by expanding the optical-fiber link from Otemachi to U
technology for fifteenkm. The actual distance b/w the 2 laboratories is twenty four- km. The
optical loss in further 15km is neg15-dB. This great optical-loss is mainly due to the fiber fixers
in the various-connected link. Fig5 represents the potential spectral density of the phase
noises assembled on the 60-km NICT to UT link. They were extracted from evaluations on light
that is moving in the free-ongoing circular-trip NICT-UT-NICT link of 120 k-m. The degree is
approx. similar as that of the 45-km loop, specifying that path among NICT and Otemachi
principally causing the phase noise. We estimated the process in the 90-km connector & make
use of this consideration to check the higher-limit of the system capacity for 60 kilometers
link.

3.3 Performance of the optical carrier transfer system:


When the 1538 nm fiber laser is used as a source of light then the current loop out signal is estimated by
combining the reference light. Phase noise of the outside loop signal is demonstrated in figure 4. In the
figurer and blue curves illustrate the outcomes with or without link equilibrium, individually and these
outcomes are achieved at 15:00 pm during the day. The transmission delay limits the servo loop band
thickness. This limits the servo's loop gain to a limited value. Theoretically the extreme of noise
suppression is denoted as 1/3(2π fτ )2 [4], here f is the Fourier frequency and the τ is the one-way
transmission time. Maximum cancellation at the 90 km link and at 1 Hertz's Fourier frequency is 56 dB.
The dominance ratio is reached to the theoretical limit which is shown in figure 4. The servo bandwidth
is reflected by the collision phase locking at 470 Hz in the static link. Here in figure 6 frequency strength
of the out-loop signal is shown.

Figure. 6.Transferred signal have frequency stabilities in the instablize (red) and stabilized (blue &
green) links. Red and green curves are found through Π-type counter and blue curves are found with a Λ-
type frequency counter. Frequency stabilities displayed in dashed and solid lines are established in the
day (15:00-17:00) and in midnight (1:00-3:00), individually.

An acknowledged fact is that Λ-type frequency counters having a dead time must not create the
suitable slope of τ−1 in the Allan aberration if a signal with white phase noise is calculated [17].
The outcomes attained with the Π -counter which is green and the Λ -type counter which is blue
is displayed. Here the dashed and solid lines indicate the measured stability which ranges in
between 15:00 to 17:00 in the day course and 1:00 to 3:00 in the night. The Π-type and Λ-type
counters have acquired the difference amongst short-term stability. This difference can be
credited to the measurement of bandwidth of every frequency counter. During night, transfer
constant evaluated with the help of Π -type counter with a slope of τ−1, outcomes in a middling
time of 1000 seconds with a constancy of 4 × 10 of18. Whereas, the daylight transfer constancy
has a slope ofτ−1/2 and it can be evaluated through Π-type counter. It becomes the noise factor
of the daytime instable link. In figure 4, the phase noise of the not so stable link has
dependence of about 1 / f 4 in between the range of 1 Hertz and 10 Hertz frequency. As the
fiber noise cancellation tracks the f 2 law, and the slope of the remaining phase noise of the day
is 1 / f 2 in the static link. As a result, the static link displays a τ−1/2 trend, much like a signal
subjugated with the white FM noise. During night, fiber noise shifts to 1 / f 3 dependencies
causing a slope of τ−1 and this huge gap between day and night is credited to the operation of
the metro in Tokyo, which is not in the midnight service. Despite the large phase noise in the
JGN2 Plus link, the transmission constancy acquired 10–18 at night, and that is much stable
than traditional satellite-based links. In accumulation, the practical average aberration of the
transmitted carrier frequency is Δν = (−2.0 ±0.4) mHz for amounts between 1:00 and 3:00 at
night, that is in statistical indecision.

3.4 Bridge between clock transition and telecom wavelength

3.4.1. Active polarization control and reliable measurement:


The change productivity of PPLN is very closely to the SOP of the information light. Consequently,
dynamic polarization control is fundamental for dependable estimations without interference. Our
programmed polarization control framework dependent on a business polarization tracker is intended
to recognize the sign in the information light from wrong segments brought about by reflections from
bring light back at connectors and joins. Figure 7 shows the schematic chart of the polarization control
framework. The out-of-circle signal is coupled to the PPLN through the polarization tracker and the
EDFA. The SOP variety because of the EDFA is additionally remunerated in this framework. The force
of the right part of the SHG light is distinguished as the RF plentiful of the particular beat recurrence
against the Ti:S recurrence brush. The yield SHG light is down-changed over to the RF space by the beat
location, where the beat note with the ideal segment is specifically gotten by a restricted band-pass
channel. The force is identified with a microwave finder (a planar doped hindrance indicator). The
business polarization tracker controls the SOP of the light by driving a fiber squeezer with the goal that
the deviation of the criticism signal is diminished. The following pace is 47π/s, and the run of the mill
SOP recuperation time is 0.7 ms, which is sufficiently quick to offset the SOP variety in the committed
connection. Figure 8 shows the practices of the power vacillations of the beat sign and SHG light with
and without the polarization control framework in the 90-km optical fiber interface. The outcomes show
that the framework essentially diminished the power vacillation because of the polarization variety and
kept the force change of the beat signal inside 2 dB for 16 hours. This epic framework is in this way
equipped for consistent activity without manual polarization changes. Be that as it may, a little force
change stayed despite the fact that the SOP is adequately settled by our polarization control framework.
On the off chance that a power fluctuating signal were sent to a recurrence counter, it may instigate off
base readouts in the recurrence estimation. To notice such a recurrence miscount, the Ti:S recurrence
brush at the remote site is stage bolted to the SHG light, and the in-circle beat signal between the
transmitted sign and the Ti:S recurrence brush is tallied to affirm a steady stage lock. Note that the Ti:S
recurrence brush was once in a while stage opened because of the force variety of the SHG light
previously establishment of the polarization control framework. Because of our strong framework, the
steady force of the last beat signal between the Ti:S recurrence brush and the clock laser at the remote
site causes the recurrence counter to peruse out faultless and dependable qualities

3.4.2. Evaluation of instabilities induced by other components:

To link an optical transition to visible light without a deterioration of function, Move stabilities of clock
lasers bridging components and 1.5 μm to be moved will go lower than the optical clock partial
instability. The most pressing part is EDFA. The length of the fiber components that make up the EDFA or
the amplified noise from spontaneous emissions (ASE) creates more phase noise. A Booster in our
framework before and after the optical carrier communication device EDFA and preamplifier EDFA are
used to offset a significant optical loss in the urban connection. An assessment of their results by means
of independent measurements utilizing a fiber interferometer that had arms with and without the
gadget under test, measurement of relative phase instability with heterodyne beat. Since the
preamplifier combines EDFA with PPLN, so is the fiber-pigtail PPLN rated in conjunction with EDFA. The
aftereffects of two volatility measurements with the Λ-type counter are delineated in Fig. 9 as bends (a)
and (b). The outofloop signal stability and interferometer device noise is shown as curves (c) and (e),
respectively. Since the instability of (a) and (b) are at the same level, we presumed that the SHG process
and PPLN does not reduce the frequency stability. There is another concern Polarization tracker-
generated noise; However, the outcomes showed that we were active the polarization control system
does not reduce frequency or give any frequency offsetting. Therefore, Ti: S is the oscillation frequency
and ECDL at the local site strongly phase locked for reference light, and their effects are minimal.
Keeping in the total instability of the all-optical connection framework takes into account all related
partial instabilities as root sum square of the bends (a), (b) and (c) (curve (d)).

4. Demonstration of direct comparison of two optical clocks

We tested our new all-optical communication network to allow a direct comparison of distant optical
clocks in Tokyo via a 60-km urban fiber link. Comparison was made of the 87Sr Lattice Clocks produced
at NICT and UT. Clock details and the frequency comb Ti: S can be found at [18–22]. From NICT the
frequency information was sent to UT, and the adjustment for fiber noise was performed at NICT. The
beat note representing the differential clock frequency was measured at UT. The total stability, including
those of the clocks and the overall system, is depicted in curve (f) of Fig. 9. The frequency measurement
was done with a Λ-type counter around midnight. The short term stability of 5×10−15 at 1 second was
dominated by the 698-nm clock laser at NICT. Figure 9 shows that no constraints are placed on the
frequency comparison by the all-optical link system. The system with the urban fiber link of 60 km
enabled us to determine the relative stability of optical clocks, 24 km away from each other. In addition,
after correction of the unusual systematic changes, the disparity shrank to less than 0.1 Hz with an
uncertainty of 0.3 Hz due to the lattice clocks, not to the fiber link [21]. In addition, a disparity in
frequency between the clocks at Hz-level is clearly noticeable over the time scale of minutes.

5.Conclusion:
This research paper expresses us that optical clock has attracted distinct consideration due to
its incredible working procedure that is the output of this clock is resulting from an optical
frequency standard. Clock contrasts are crucial for determining dependability of frequency
standards. We set up an all optical link joining two faraway placed optical clocks. By means of
an optical carrier transmission system a 1.5 micrometers light was firmly shifted over a 90-
kilometers urban fiber connection. The transmission steadiness was 2×10-15 in one second and
also 4×10-18 at one thousand seconds. On the way to maintain the strength of second
harmonic light and for the sake of permitting longstanding as well as for good calculations, A
polarization controller scheme for the sake of transmitted light had been established. The
general system unsteadiness, as well as the uncertainties related to two EDFAs (Erbium doped
fiber amplifiers) outside the phase- noise compensated path were 2× 10-15 within one second
also 7 × 10-17 at thousand seconds.
The immediate evaluation between two 87Sr (Strontium) lattice clocks have been understood
in the 10-6 variability with respect to the set up established at this point. The unsteadiness
related to contrast was not restricted through overall unsteadiness of the all-optical
connection. The small-range stability optical clocks have been enhanced to 10-6 level [23]. To
investigate such steady optical clocks, one should utilize tracing laser operation as little noise
immense gain amplifier rather than single directing EDFA (Erbium doped fiber amplifiers) used
in a present system. Furthermore, a soundless optical fiber connection will ease additional
steady transmissions.

References and links:


1. G. Panfilo and E. Felicitas Arias, “Algorithms for international atomic time,” IEEE Trans.
Ultrason. Ferroelectr.

Freq. Control 57(1), 140–150 (2010).

2. C. W. Chou, D. B. Hume, J. C. J. Koelemeiji, D. J. Wineland, and T. Rosenband, “Frequency


comparison of two

high-accuracy Al+ optical clocks,” Phys. Rev. Lett. 104, 070802 (2010).

3. S. M. Foreman, K. W. Holman, D. D. Hudson, D. J. Jones, and J. Ye, “Remote transfer of


ultrastable frequency

references via fiber networks,” Rev. Sci. Instrum. 78, 021101 (2007).

4. P. A. Williams, W. C. Swann, and N. R. Newbury, “High-stability transfer of an optical


frequency over long

fiber-optic links,” J. Opt. Soc. Am. B 25(8), 1284–1293 (2008).

5. M. Musha, F.-L. Hong, K. Nakagawa, and K. Ueda, “Coherent optical frequency transfer over
50-km physical

distance using a 120-km-long installed telecom fiber network,” Opt. Express 16(21), 16459–
16466 (2008).

6. L. S. Ma, P. Jungner, J. Ye, and J. L. Hall, “Delivering the same optical frequency at two
places: accurate cancellation of phase noise introduced by an optical fiber or other time-varying
path,” Opt. Lett. 19(21), 1777–1779

(1994).
7. N. R. Newbury, P. A. Williams, and W. C. Swann, “Coherent transfer of an optical carrier
over 251 km,” Opt.

Lett. 32(21), 3056–3058 (2007).

8. F. Kefelian, O. Lopez, H. Jiang, C. Chardonnet, A. Amy-Klein, and G. Santarelli, “High-


resolution optical frequency dissemination on a telecommunications network with data traffic,”
Opt. Lett. 34(10), 1573–1575 (2009).

9. O. Lopez, A. Haboucha, F. Kefelian, H. Jiang, B. Chanteau, V. Roncin, C. Chardonnet, A.


Amy-Klein, and G.

Santarelli, “Cascaded multiplexed optical link on a telecommunication network for frequency


dissemination,”

Opt. Express 18(16), 16849–16857 (2010).

10. O. Terra, G. Grosche, and H. Schnatz, “Brillouin amplification in phase coherent transfer of
optical frequencies

over 480 km fiber,” Opt. Express 18(15), 16102–16111 (2010).

11. A. D. Ludlow, T. Zelevinsky, G. K. Campbell, S. Blatt, M. M. Boyd, M. H. G. de Miranda,


M. J. Martin, J. W.

Thomsen, S. M. Foreman, J. Ye, T. M. Fortier, J. E. Stalnaker, S. A. Diddams, Y. Le Coq, Z. W.


Barber, N. Poli,

N. D. Lemke, K. M. Beck, and C. W. Oates, “Sr lattice clock at 1×10−16 fractional uncertainty
by remote optical

evaluation with a Ca clock,” Science 319, 1805–1808 (2008).

12. O. Terra, G. Grosche, K. Predehl, R. Holzwarth, T. Legero, U. Sterr, B. Lipphardt, and H.


Schnatz, “Phasecoherent comparison of two optical frequency standards over 146 km using a
telecommunication fiber link,”

Appl. Phys. B 97, 541–551 (2009).

13. A. Pape, O. Terra, J. Friebe, M. Riedmann, T. Wubbena, E. M. Rasel, K. Predehl, T. Legero,


B. Lipphardt,

H. Schnatz, and G. Grosche, “Long-distance remote comparison of ultrastable optical


frequencies with 10−15

instability in fractions of a second,” Opt. Express 18(20), 21477–21483 (2010).


14. Japan Gigabit Network 2 plus [Online]. Available:

http://www.jgn.nict.go.jp/jgn2plus archive/english/index.html.

15. M. Kumagai, M. Fujieda, S. Nagano, and M. Hosokawa, “Stable radio frequency transfer in
114 km urban optical

fiber link,” Opt. Lett. 34(19), 2949–2951 (2009).

16. M. Fujieda, M. Kumagai, S. Nagano, “Coherent microwave transfer over a 204-km telecom
fiber link by a

cascaded system,” IEEE Trans. Ultrason. Ferroelectr. Freq. Control 57(1), 168–174 (2010).

17. P. Lesage, “Characterization of frequency stability: bias due to the juxtapositon of time-
interval measurements,”

IEEE Trans. Instrum. Meas. 32(1), 204–207 (1983).

18. M. Takamoto, F. L. Hong, R. Higashi, and H. Katori, “An optical lattice clock,” Nature 435,
321–324 (2005).

19. T. Akatsuka, M. Takamoto, and H. Katori, “Optical lattice clocks with non-interacting boson
and fermions,” Nat.

Phys. 4, 954–959 (2008).

20. M. Takamoto, T. Takano, and H. Katori, “Frequency comparison of optical lattice clocks
beyond the Dick limit,”

Nat. Photonics 5, 288–292 (2011).

21. A. Yamaguchi, M. Fujieda, M. Kumagai, H. Hachisu, S. Nagano, Y. Li, T. Ido, T. Takano,


M. Takamoto, and

H. Katori, “Direct comparison of distant optical lattice clocks at the 10−16 uncertainty,” Appl.
Phys. Express 4,

082203 (2011).

22. S. Nagano, H. Ito, Y. Li, K. Matsubara, and M. Hosokawa, “Stable operation of femtosecond
laser frequency

combs with uncertainty at the 10−17 level toward optical frequency standards,” Jpn. J. Appl.
Phys. 48, 042301

(2009).
23. Y. Y. Jiang, A. D. Ludlow, N. D. Lemke, R. W. Fox, J. A. Sherman, L. S. Ma, and C. W.
Oates, “Making optical

atomic clocks more stable with 10−16-level laser stabilization,” Nat. Photonics 5, 158–161
(2011).

You might also like