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

Principle and performance analysis of coherent tracking sensor based on local oscillator beam nutation

Open Access Open Access

Abstract

By integrating coherent tracking sensor functionality into a coherent communication receiver, a dual-function receiver with coherent boresight error sensing is developed for tracking in coherent free-space optical communication systems. The sensor principles are analyzed according to optical interference theory, and the boresight error detection algorithm and small signal linear model are derived. Analysis of local-oscillator beam nutation on system performance shows that the best nutation half-angle is 0.5–1μrad, the noise equivalent angle is less than 0.02μrad, and the communication sensitivity degradation is 0.2–0.6dB. This technology opens new avenues for free-space optical communication system design.

© 2014 Optical Society of America

1. Introduction

Free-space optical (FSO) communication has become a major technical trend for satellite communication due to its high data rate. With high receiving sensitivity and immunity to the sun, coherent communication is the first choice for long-distance high-data-rate optical links such as low-earth orbit (LEO) to geosynchronous-earth orbit (GEO) or GEO to optical-ground station (OGS) links. Many current optical communication systems are based on coherent technologies, such as the German in-orbit validation laser communication terminals for the LEO-LEO link between the TerraSAR-X observation satellite and the near-field infrared experiment laser communication terminal (NFIRELCT) (5.625Gbps), the LEO-GEO link between the AlphaSat and the Sentinel 1A and 2A satellites (1.8Gbps) [16], and the developing American laser communications relay demonstration (LCRD) (2.88Gbps) [7], and the Japanese next generation optical communication terminals (2.5Gbps) [8].

Coherent terminals tend to apply coherent tracking technologies, and coherent tracking sensors are integrated into the receiver to simplify the optical path and detector assembly. Coherent tracking technology is immune to background light, and the noise equivalent angle (NEA) is smaller than that of incoherent tracking technology due to the nature of the coherence. In coherent systems, the receiving sensitivity is limited by shot noise, and coherent conditions make receiving highly selective. Low-level background radiation such as starlight has a negligible effect on the detecting process. Under high-level background radiation such as sun, when the system is appropriately filtered, the sensitivity degrades, but the signal is not interrupted as it would be with incoherent systems. The integrated coherent tracking sensor uses light for communication to obtain the boresight error. Due to the differences between the communicating bandwidth and the noise equivalent bandwidth (NEB) of the tracking loop, the signal-to-noise ratio (SNR) of tracking sensor is very high. When the SNR is sufficient for communication, the NEA is very small.

At present, there are two primary coherent tracking sensor technologies: coherent tracking sensors based on binary detectors and coherent tracking sensors based on fiber nutation. The binary-detector-based coherent tracking systems are applied on the laser communication terminal (LCT) from the German Aerospace Research Establishment (DLR) [9] and were verified by in-orbit experiments on the TerraSAR-X and NFIRE satellites. The fiber-nutation-based coherent tracking systems were developed by the Massachusetts Institute of Technology [1013] and adopted by NASA’s lunar laser communications demonstration (LLCD) project [14, 15] and the RUAG Space’s OPTEL laser terminal family [16, 17]. In this paper, we propose a third type of coherent tracking sensor, which utilizes the nutating local-oscillator beam to obtain the boresight error from the received communication light and is integrated in the coherent communication receiver.

The working principle of the binary-detector-based coherent tracking sensor is similar to a quadrant detector (QD). Each of the balanced detectors in the in-phase branch of the optical hybrid output is split into two equal parts by a linear dead zone. The two linear dead zones in the detectors are perpendicular. The boresight error is obtained by measuring and calculating the normalized difference of the interfering light photocurrent from each binary detector [9]. As the number of individual active parts increases, the detector requires large number of amplifiers, adders, and subtractors in the subsequent processing circuitry, increasing the complexity of the circuit. Combining the two signals from two different parts of the binary detector causes a little more noise than a single active area detector, leading to minor receiver sensitivity losses.

The fiber-nutation-based coherent tracking sensor obtains the boresight error by synchronously demodulating the fiber nutation and the signal envelope fluctuation of the mixed light between the received light being coupled into fiber and the local oscillator (LO) [1013]. Due to the difference between the space light field and the fiber waveguide mode, there is a theoretical minimal coupling loss of about 0.86dB. Additionally, fiber nutation introduces loss, so the total loss of the receiving sensitivity is more than 1dB. Furthermore, the fiber may fracture due to material fatigue, resulting in low reliability when the fiber is used as a nutation device. In addition, the fiber device needs to be radiation hardened when it is used in space, which is a hindrance to the volume and weight.

In this paper, we present a LO-beam nutation-based coherent tracking sensor. In this technology, the boresight error is calculated by synchronously demodulating the LO nutation and the signal envelope fluctuation of the mixed light between received light and the LO. Similarly to binary-detector-based coherent tracking sensors, with surplus power, the LO light field can be spatially reshaped to match to the received light field to improve the coherent efficiency with acceptable increase in collimator’s size and mass, and laser’s power consumption. A beam deflector will be added to the LO optical path in the coherent receiver, and an envelope detector must be added to the receiver circuits. The beam deflection device could be either acousto-optic or electro-optic, providing a substantial increase in reliability compared to the fiber nutation detector. The single element detectors used in this coherent tracking sensor will have smaller active area than that of binary-detector-based tracking sensor, which would be favorable for increasing the communication rate and decreasing the circuit noise.

2. Principle of coherent tracking sensor based on local oscillator beam nutation

Based on the principles of optical interference, when the focus of the received light does not perfectly overlap with the focus of LO beam, the coherence efficiency decreases. Then, the photocurrent from the detector and the voltage signal after the transimpedance preamplifier will both decrease. When the focus of the LO beam moves in circular route on the detector active area and the nutating center coincides with the focus of the received light, the coherent signal envelope is constant. Otherwise, the signal envelope periodically fluctuates. The terminal controller (TECO) determines the boresight error by synchronizing the periodic coherent signal envelope fluctuation and the position of the LO focus. For example, consider the BPSK receiver based on a decision-directed phase-locked loop. Figure 1 presents the principle block diagram of a coherent tracking sensor based on LO beam nutation that is integrated into this BPSK receiver.

 figure: Fig. 1

Fig. 1 Principle block diagram of coherent tracking sensor based on local oscillator beam nutation, which is integrated in a BPSK receiver.

Download Full Size | PDF

The photocurrent generated from a photodiode can be written as

iD(t)DI(r,t)ds=D[ELO(r,t)+ES(r,t)][ELO(r,t)+ES(r,t)]*ds=PLO(t)+PS(t)+2PLO(t)PS(t)Re{γ(t)}

Where I(r,t) is the intensity distribution in the interference plane, ELO(r,t) andES(r,t) are the complex electrical field amplitude distribution of the LO beam and received light, respectively, PLO(t) and PS(t) are the power of the LO beam and received light, respectively. For simplicity, all interference results are calculated along the entrance pupil of the telescope. The incident angle in the following analysis is also along this plane. The value for the inner position of the optical system can be obtained by multiplying the magnification of the optical system.

The complex degree of coherence γ(t) in Eq. (1) is defined as

γ(t)=DELO(r,t)ES*(r,t)dsD|ELO(r,t)|2dsD|ES(r,t)|2ds

From Eq. (1), we see that the real part of the complex degree of coherence affects the amplitude of the receiving signal. The complex degree of coherence cannot be measured directly; however, it can be obtained from the real part measurement in two special cases: (1) the amplitude distribution of the two light fields is uniform or (2) the phase distribution is the same. In the performance analysis of coherent free-space optical communication, we focus only on the real part of the complex degree of coherence. It is defined as the coherence efficiencyη(t):

η(t)=Re{γ(t)}=D|ELO(r,t)||ES(r,t)|cos[Δωt+Δϕ(r,t)]dsD|ELO(r,t)|2dsD|ES(r,t)|2ds

2.1 Envelope detector for coherent signal

The far-field light transmitted from long distances is only partially received by the telescope, which we consider to be far-field light sampling by the telescope aperture. The receiving light in range of the sample aperture can be treated as a plane wave, and the field transformed through the telescope optics is treated as a beam-compressed plane wave. The initial output of the LO is Gaussian, which can be reshaped to a plane wave after reverse-Gaussian-profile filtering. Therefore, both the amplitude distributions of the LO and the received light fields are uniform. Using Eq. (3), the coherence efficiency of paths I and Q can be written as

ηI(t)=1SDcos[Δωt+Δϕ(r,t)]ds=1Sa2(t)+b2(t)cos[Δωt+atanb(t)a(t)]
ηQ(t)=1SDcos[Δωt+Δϕ(r,t)π2]ds=1Sa2(t)+b2(t)sin[Δωt+atanb(t)a(t)]
where

a(t)=Dcos[Δϕ(r,t)]dsb(t)=Dsin[Δϕ(r,t)]ds

Here, S is the area of the integration domain, Δω is the heterodyne frequency of the LO to the received light, and Δφ(r,t) is the phase difference between the LO and the received light.

The direct current is cancelled in balanced photodetectors, while the alternating current is transformed to a voltage signal by the transimpedance amplifier. The voltage signals in the I and Q paths can be written as

vI(t)=2RrLηI(t)(1k)PLOPS(t)+nI(t)
vQ(t)=2RrLηQ(t)kPLOPS(t)+nQ(t)

Where R is the responsibility of the photodetectors, and rL is the resistance of the transimpedance. Because the LO output power is constant, the symbol PLO(t) is modified to PLO. The additive noises in the I and Q paths are nI(t) and nQ(t), respectively, and in coherent systems, the additive noises are primarily formed from shot noise. The power spectral density (PSD) can be written as [1825]:

SNI(f)=SNQ(f)=2eRPLOrL2

We can determine the amplitude envelope of the coherent signal by multiplying the amplification factors k and 1-k by the square of the signal in the I and Q paths and summing the results:

v(t)kvI2(t)+(1k)vQ2(t)=s(t)+n(t)

We only consider the zero and first order terms in the Zernike polynomials for the phase difference. Without loss of generality, the phase difference distribution can be written as

Δϕ(r,t)=Δϕ0(t)+xtan[φ(t)]Δϕ0(t)+xφ(t)

Where φ(t) is the angle between the LO wavefront and the received light wavefront, and x is the abscissa. Here, Δϕ0(t) is the zero order phase difference of the LO and received light. Then, Eq. (6) can be written as

a(t)=ScosΔϕ0(t)2J1[2πRDλφ(t)]2πRDλφ(t)b(t)=SsinΔϕ0(t)2J1[2πRDλφ(t)]2πRDλφ(t)

In order to simplify the expression, we set ψ(t) = 2πRDφ(t)/λ, where RD is the radius of the telescope entrance pupil. In Eq. (12), J1(∙) is the first order Bessel function. Substituting Eqs. (4)(8) and Eq. (12) into Eq. (10),the voltage envelope of the signal s(t) becomes

s(t)=4R2rL2k(1k)PLOPS(t){2J1[ψ(t)]/[ψ(t)]}2

The noise n(t) is then

n(t)=8RrLkk(1k)PLOPS(t)J1[[ψ(t)]]/[[ψ(t)]]cos(Δωt+Δϕ0(t))nI(t)+...+8RrL(1k)k(1k)PLOPS(t)J1[[ψ(t)]]/[[ψ(t)]]sin(Δωt+Δϕ0(t))nQ(t)+...+knI2(t)+(1k)nQ2(t)4RrLk(1k)PLOPSmax{kcos(Δωt+Δϕ0(t))nI(t)+(1k)sin(Δωt+Δϕ0(t))nQ(t)}

Where PSmax is the maximum received power of the communication link, and the noise PSD can be written as

SN(f)8R2rL2k(1k)PLOPSmaxSNI(f)=16eR3rL4k(1k)PLO2PSmax

2.2 The algorithm for boresight error detection and its equivalent linear model

The LO beam nutates around the boresight of the optical system. As shown in Fig. 2, φN is the half-nutation angle, and α(t) is the azimuth boresight error for the received light and angle of arrival (AOA). We suppose that a negligible change occurs in the boresight error during a half-circle nutation of the LO beam and analyze the azimuthal component of the boresight error. The process for determining the elevation is identical.

 figure: Fig. 2

Fig. 2 Schematic diagram of the intersection angle between wavefronts of LO and received light in nutation process. Point C.T. is the AOA coordinate of the received light from counter terminal.

Download Full Size | PDF

When the tracking has been established, the boresight error is small. Therefore, the angle between the LO and the received light wavefront can be approximated as

φ(t)φNα(t)φ(tT/2)α(tT/2)+φN

By substituting Eq. (16) into Eq. (13) and Eq. (10), the voltage envelope of the coherent signal can be approximated to the first-order as

v(t)4k(1k)R2rL2PLOPS(t){2J1{2πRDλ[φNα(t)]}2πRDλ[φNα(t)]}2+n(t)4k(1k)R2rL2PLOPS(t){[2J1(ψN)ψN]2SF(φN)α(t)}+n(t)
Where ΨN = 2πRDφN/λ. In a similar way, we find
v(tT/2)4k(1k)R2rL2PLOPS(tT/2){[2J1(ψN)ψN]2+SF(φN)α(tT/2)}+n(tT/2)
where the slope factor SF(φN) is

SF(φN)=d[2J1(ψ(t))ψ(t)]2/dφ(t)|φN=16πRDλJ1(ψN)ψN[J0(ψN)ψN2J1(ψN)ψN2]

Similarly, when the tracking has been established, the varing range of the received power is very small. And the sampling frequency 2/T is much higher than the pointing error frequency, where T is the period of nutation. Therefore, PS(t) ≈PS(t – T/2), and we can write

v(t)v(tT/2)4k(1k)R2rL2PLOPS(t)SF(φN)[2α(t)T2dα(t)dt]+n(t)n(tT/2)
v(t)+v(tT/2)4k(1k)R2rL2PLOPS(t){2{2J1[ψN][ψN]}2SF(φN)T2dα(t)dt}

Here, (T/2)[dα(t)/dt] is the small first-order quantity of α(t – T/2), and so

SF(φN)T2dα(t)dt<<SF(φN)α(t)

Furthermore, because SF(φN)α(t) is the small first-order quantity of{2J1{(2πR/λ)[φNα(t)]}/{(2πR/λ)[φNα(t)]}}2, we find

SF(φN)α(t)<<{2J1[ψN][ψN]}2

So that

SF(φN)T2dα(t)dt<<2{2J1[ψN][ψN]}2

Therefore, in Eq. (21), SF(φN)(T/2)[dα(t)/dt] is negligible.

The optical-axis error can be calculated as follows:

eα(t)=v(t)v(tT/2)v(t)+v(tT/2)SF(φN){2J1[ψN][ψN]}2[α(t)T4dα(t)dt]+n(t)

Where n’(t) is

n(t)[n(t)n(tT/2)]/[8k(1k)R2r2PLOPSmax]

The PSD of n’(t) can be written as

SN(f)e/[2k(1k)RPSmax]

Finally, the linear transfer function of the LO-beam nutation-based coherent tracking sensor can be approximated as

GS(s)=Eα(s)A(s)=SF(φN){2J1[ψN][ψN]}2(T4s+1)
where A(s) is the Laplace transform of the optical axis error α(t).

Suppose that the nutation period of LO beam is T = 30 μs. Then, the nutation frequency is approximately 33.3kHz. The normalized frequency response characteristics of the LO-beam nutation-based coherent tracking sensor are shown in Fig. 3.

 figure: Fig. 3

Fig. 3 Normalized frequency response characteristics of coherent tracking sensor based on LO beam nutation.

Download Full Size | PDF

Figure 3 shows that when the angular frequency of boresight error falls below 104rad/s, there is little deviation of the magnitude gain or the phase delay. Because the vibration spectrum primarily falls in this range, we can approximate the sensor normalized gain by 1. Then, the phase lag is less than 10°. Therefore, the frequency response characteristic in this range is appropriate for engineering applications. When the angular frequency of boresight error is higher than 104rad/s, the frequency response characteristics change significantly. The sensor gain increases, while the phase lag quickly increases to −90°. This demonstrates that the amplitude-frequency characteristics of the subsequent devices should be cut off after 104rad/s, and a large phase lag should be avoided.

3. Impact of nutation angle on system performance

3.1 Impact of nutation angle on tracking error

The noise generated from the detectors and electric circuits will enter the tracking loop, as shown in Fig. 4. The tracking error caused by these noises is the noise equivalent angle (NEA).

 figure: Fig. 4

Fig. 4 Linear model of a tracking loop.

Download Full Size | PDF

Here, Θα(f) is the azimuthal component of AOA, and GFPA(f) is the transfer function of fine point assembly (FPA).

The PSD of the NEA can be determined through a linear system analysis:

SNEA(f)=SN(f)|GL(f)/GS(f)|2

Where GL(f) is the closed-loop transfer function of the tracking loop in the frequency domain:

GL(f)=GS(f)GFPA(f)/[1+GS(f)GFPA(f)]

Here, SN’(f) is a constant, and the NEA can be written as

NEA=0SNEA(f)df=SN(f)Bn

Where Bn is the noise equivalent bandwidth.

Bn=0|GL(f)/GS(f)|2df

We show an example of the relationship between the half-nutation angle and NEA in Fig. 5. The primary calculation parameters are defined as follows: the receiving optical power is −55dBm, the aperture of the telescope is 150mm, the wavelength of the signal and LO beam is 1μm, and the power splitting ratio of the I path to the Q path is 99:1. The frequency characteristics of a supposed GL(f) are shown in Fig. 6. The −3dB bandwidth of FPA in this loop is 1500Hz. The FPA is controlled by a digital PID controller, and the update rate of the loop is 20kHz. The parameters of the PID are automatically tuned by the software.

 figure: Fig. 5

Fig. 5 The relationship between half-nutation angle and NEA.

Download Full Size | PDF

 figure: Fig. 6

Fig. 6 The frequency response characteristic of the tracking loop.

Download Full Size | PDF

From Fig. 5, when the half-nutation angle is larger than 0.5μrad, we find that the NEA is less than 0.02μrad. Therefore, the NEA can be ignored in a physical system. As the nutation angle increases, the coherence efficiency sharply decreases. Thus, by satisfying the tracking precision, the nutation angle should be minimized. In next section, we analyze the impact of the nutation angle on communication receiver sensitivity.

3.2 The impact of nutation angle on communication receiver sensitivity

The SNR of the homodyne BPSK coherent receiver without phase locking error can be written as

SNR=|vI|2|nI|2=4R2rL2(1k)PLOPS(t){2J1[ψ(t)]/[ψ(t)]}2SNIB

Where B is the single sideband noise equivalent bandwidth of the communication signal. Therefore, the sensitivity degradation caused by nutation is 10log{{2J1[ψ(t)]/ψ(t)}2}. We show the relationship between the half-nutation angle and receiving sensitivity degradation is shown in Fig. 7.

 figure: Fig. 7

Fig. 7 The relationship between the half-nutation angle and the receiving sensitivity degradation.

Download Full Size | PDF

Figure 7 shows that the receiving sensitivity quickly degrades as the nutation angle is increased. Combined with our earlier analysis, we find that the best half-nutation angle range is 0.5–1μrad. In this range, the receiving sensitivity degradation is approximately 0.2–0.6dB, and the NEA can be ignored.

4. Conclusion

We present a coherent tracking sensor based on LO beam nutation in the coherent optical communication receiver. The results of the theoretical inference and simulation studies show that when the half nutation angle lies between 0.5 and 1μrad, the NEA is less than 0.02μrad, and the degradation of the receiving sensitivity ranges from 0.2 to 0.6dB. The proposed LO nutation-tracking system is a new approach with comparable performance to existing systems and with potential implementation advantages, depending upon the application. The LO-beam nutation-based coherent tracking sensor has a simple structure, which is easy to realize. The sensor is integrated into a typical coherent receiver by adding a beam deflector to the LO optical path in the coherent receiver, and an envelope detector to the receiver circuits. The LO-beam nutation-based coherent tracking sensor provides a new choice for coherent free-space optical communication system designs.

References and links

1. S. Seel, H. Kampfner, F. Heine, D. Dallmann, G. Muhlnikel, M. Gregory, M. Reinhardt, K. Saucke, J. Muckherjee, and U. Sterr, “Space to ground bidirectional optical communication link at 5.6 Gbps and EDRS connectivity outlook,” in Aerospace Conference(IEEE, 2011), pp. 1–7. [CrossRef]  

2. K. Böhmer, M. Gregory, F. Heine, H. Kämpfner, R. Lange, M. Lutzer, and R. Meyer, “Laser Communication Terminals for the European Data Relay System,” Proc. SPIE 8246, 82460D (2012). [CrossRef]  

3. R. Lange, F. Heine, H. Kampfner, R. Czichy, R. Meyer, and M. Lutzer, “Optical inter-satellite communication operational,” in Military Communications Conference, (IEEE, 2010), pp. 2290–2294.

4. B. Smutny, H. Kaempfner, G. Muehlnikel, U. Sterr, B. Wandernoth, F. Heine, U. Hildebrand, D. Dallmann, M. Reinhardt, A. Freier, R. Lange, K. Boehmer, T. Feldhaus, J. Mueller, A. Weichert, P. Greulich, S. Seel, R. Meyer, and R. Czichy, “5.6 Gbps optical intersatellite communication link,” Proc. SPIE 7199, 719906 (2009). [CrossRef]  

5. B. Smutny, R. Lange, H. Kämpfner, D. Dallmann, G. Mühlnikel, M. Reinhardt, K. Saucke, U. Sterr, B. Wandernoth, and R. Czichy, “In-orbit verification of optical inter-satellite communication links based on homodyne BPSK,” Proc. SPIE 6877, 687702 (2008). [CrossRef]  

6. B. Gütlich, R. Meyer, S. Philipp-May, and A. Pagels-Kerp, “German Roadmap on Optical Communication in Space,” in Advanced Solid-State Lasers Congress, P. W. McManamon and O. Steinvall, eds. (Optical Society of America, Paris, 2013), paper LM1B.2. [CrossRef]  

7. B. L. Edwards, K. Wilson, D. Israel, J. Moores, and A. Fletcher, “Overview of the laser communications relay demonstration project,” in Proceedings of SpaceOps, 1261897 (2012).

8. S. Yamakawa, T. Hanada, and H. Kohata, “R&D status of the next generation optical communication terminals in JAXA,” inSpace Optical Systems and Applications, (IEEE, 2011), pp. 389–393.

9. D. Giggenbach, A. Schex, and B. Wandernoth, “Prototype of a coherent tracking and detection receiver with wideband vibration compensation for free-space laser communications,” Proc. SPIE 2699, 186–191 (1996). [CrossRef]  

10. E. A. Swanson and R. S. Bondurant, “Using fiber optics to simplify free-space lasercom systems,” Proc. SPIE 1218, 70–82 (1990). [CrossRef]  

11. T. E. Knibbe, “Spatial tracking using an electro-optic nutator and a single-mode optical fiber,” (Massachusetts Institute of Technology, 1993).

12. E. A. Swanson and R. S. Bondurant, A Space-Based Optical Communication System Utilizing Fiber Optics (Massachusetts Institute of Technology Lincoln Laboratory, 1989).

13. E. A. Swanson and R. S. Bondurant, “Fiber-based receiver for free-space coherent optical communication systems,” in Optical Fiber Communication Conference(Optical Society of America, 1989) paper THC5.

14. J. W. Burnside, S. D. Conrad, A. D. Pillsbury, and C. E. DeVoe, “Design of an Inertially Stabilized Telescope for the LLCD,” Proc. SPIE 7923, 79230L (2011). [CrossRef]  

15. D. M. Boroson, J. J. Scozzafava, D. V. Murphy, B. S. Robinson, and H. Shaw, “The lunar laser communications demonstration (LLCD),” in IEEE Space Mission Challenges for Information Technology, (IEEE, 2009), pp. 23–28.

16. G. Baister, T. Dreischer, E. R. Grond, L. Gallmann, B. Thieme, and K. Kudielka, “The OPTEL terminal development programme–enabling technologies for future optical crosslink applications,” in AIAA Space Conference and Exposition, pp. 1–8 (2003).

17. K. Kudielka, A. Marki, G. C. Baister, and C. Haupt, “Acquisition and tracking strategies of the OPTEL 25 GEO terminal,” in Lasers and Electro-Optics Society(IEEE, 2002), pp. 395–396.

18. Y. Yamamoto, “Receiver performance evaluation of various digital optical modulation-demodulation systems in the 0.5-10 µm wavelength region,” IEEE J. Quantum Electron. 16(11), 1251–1259 (1980). [CrossRef]  

19. T. Okoshi, K. Emura, K. Kikuchi, and R. T. Kersten, “Computation of bit-error rate of various heterodyne and coherent-type optical communication schemes,” J. Opt. Commun. 2(3), 89–96 (1981). [CrossRef]  

20. Y. Yamamoto and T. Kimura, “Coherent optical fiber transmission systems,” IEEE J. Quantum Electron. 17(6), 919–935 (1981). [CrossRef]  

21. F. Favre, L. Jeunhomme, I. Joindot, M. Monerie, and J. Simon, “Progress towards heterodyne-type single-mode fiber communication systems,” IEEE J. Quantum Electron. 17(6), 897–906 (1981). [CrossRef]  

22. T. Okoshi and K. Kikuchi, “Heterodyne-type optical fiber communications,” J. Opt. Commun 2(3), 82–88 (1981). [CrossRef]  

23. T. Okoshi, “Heterodyne and coherent optical fiber communications: recent progress,” IEEE Trans. Microw. Theory Tech. 30(8), 1138–1149 (1982). [CrossRef]  

24. R. Hooper, J. Midwinter, D. Smith, and I. Stanley, “Progress in monomode transmission techniques in the United Kingdom,” J. Lightwave Technol. 1(4), 596–611 (1983). [CrossRef]  

25. T. Okoshi, “Recent progress in heterodyne/coherent optical-fiber communications,” J. Lightwave Technol. 2(4), 341–346 (1984). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (7)

Fig. 1
Fig. 1 Principle block diagram of coherent tracking sensor based on local oscillator beam nutation, which is integrated in a BPSK receiver.
Fig. 2
Fig. 2 Schematic diagram of the intersection angle between wavefronts of LO and received light in nutation process. Point C.T. is the AOA coordinate of the received light from counter terminal.
Fig. 3
Fig. 3 Normalized frequency response characteristics of coherent tracking sensor based on LO beam nutation.
Fig. 4
Fig. 4 Linear model of a tracking loop.
Fig. 5
Fig. 5 The relationship between half-nutation angle and NEA.
Fig. 6
Fig. 6 The frequency response characteristic of the tracking loop.
Fig. 7
Fig. 7 The relationship between the half-nutation angle and the receiving sensitivity degradation.

Equations (33)

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

i D ( t ) D I ( r , t ) d s = D [ E L O ( r , t ) + E S ( r , t ) ] [ E L O ( r , t ) + E S ( r , t ) ] * d s = P L O ( t ) + P S ( t ) + 2 P L O ( t ) P S ( t ) Re { γ ( t ) }
γ ( t ) = D E L O ( r , t ) E S * ( r , t ) d s D | E L O ( r , t ) | 2 d s D | E S ( r , t ) | 2 d s
η ( t ) = Re { γ ( t ) } = D | E L O ( r , t ) | | E S ( r , t ) | cos [ Δ ω t + Δ ϕ ( r , t ) ] d s D | E L O ( r , t ) | 2 d s D | E S ( r , t ) | 2 d s
η I ( t ) = 1 S D cos [ Δ ω t + Δ ϕ ( r , t ) ] d s = 1 S a 2 ( t ) + b 2 ( t ) cos [ Δ ω t + a tan b ( t ) a ( t ) ]
η Q ( t ) = 1 S D cos [ Δ ω t + Δ ϕ ( r , t ) π 2 ] d s = 1 S a 2 ( t ) + b 2 ( t ) sin [ Δ ω t + a tan b ( t ) a ( t ) ]
a ( t ) = D cos [ Δ ϕ ( r , t ) ] d s b ( t ) = D sin [ Δ ϕ ( r , t ) ] d s
v I ( t ) = 2 R r L η I ( t ) ( 1 k ) P L O P S ( t ) + n I ( t )
v Q ( t ) = 2 R r L η Q ( t ) k P L O P S ( t ) + n Q ( t )
S N I ( f ) = S N Q ( f ) = 2 e R P L O r L 2
v ( t ) k v I 2 ( t ) + ( 1 k ) v Q 2 ( t ) = s ( t ) + n ( t )
Δ ϕ ( r , t ) = Δ ϕ 0 ( t ) + x tan [ φ ( t ) ] Δ ϕ 0 ( t ) + x φ ( t )
a ( t ) = S cos Δ ϕ 0 ( t ) 2 J 1 [ 2 π R D λ φ ( t ) ] 2 π R D λ φ ( t ) b ( t ) = S sin Δ ϕ 0 ( t ) 2 J 1 [ 2 π R D λ φ ( t ) ] 2 π R D λ φ ( t )
s ( t ) = 4 R 2 r L 2 k ( 1 k ) P L O P S ( t ) { 2 J 1 [ ψ ( t ) ] / [ ψ ( t ) ] } 2
n ( t ) = 8 R r L k k ( 1 k ) P L O P S ( t ) J 1 [ [ ψ ( t ) ] ] / [ [ ψ ( t ) ] ] cos ( Δ ω t + Δ ϕ 0 ( t ) ) n I ( t ) + ... + 8 R r L ( 1 k ) k ( 1 k ) P L O P S ( t ) J 1 [ [ ψ ( t ) ] ] / [ [ ψ ( t ) ] ] sin ( Δ ω t + Δ ϕ 0 ( t ) ) n Q ( t ) + ... + k n I 2 ( t ) + ( 1 k ) n Q 2 ( t ) 4 R r L k ( 1 k ) P L O P S max { k cos ( Δ ω t + Δ ϕ 0 ( t ) ) n I ( t ) + ( 1 k ) sin ( Δ ω t + Δ ϕ 0 ( t ) ) n Q ( t ) }
S N ( f ) 8 R 2 r L 2 k ( 1 k ) P L O P S max S N I ( f ) = 16 e R 3 r L 4 k ( 1 k ) P L O 2 P S max
φ ( t ) φ N α ( t ) φ ( t T / 2 ) α ( t T / 2 ) + φ N
v ( t ) 4 k ( 1 k ) R 2 r L 2 P L O P S ( t ) { 2 J 1 { 2 π R D λ [ φ N α ( t ) ] } 2 π R D λ [ φ N α ( t ) ] } 2 + n ( t ) 4 k ( 1 k ) R 2 r L 2 P L O P S ( t ) { [ 2 J 1 ( ψ N ) ψ N ] 2 S F ( φ N ) α ( t ) } + n ( t )
v ( t T / 2 ) 4 k ( 1 k ) R 2 r L 2 P L O P S ( t T / 2 ) { [ 2 J 1 ( ψ N ) ψ N ] 2 + S F ( φ N ) α ( t T / 2 ) } + n ( t T / 2 )
S F ( φ N ) = d [ 2 J 1 ( ψ ( t ) ) ψ ( t ) ] 2 / d φ ( t ) | φ N = 16 π R D λ J 1 ( ψ N ) ψ N [ J 0 ( ψ N ) ψ N 2 J 1 ( ψ N ) ψ N 2 ]
v ( t ) v ( t T / 2 ) 4 k ( 1 k ) R 2 r L 2 P L O P S ( t ) S F ( φ N ) [ 2 α ( t ) T 2 d α ( t ) d t ] + n ( t ) n ( t T / 2 )
v ( t ) + v ( t T / 2 ) 4 k ( 1 k ) R 2 r L 2 P L O P S ( t ) { 2 { 2 J 1 [ ψ N ] [ ψ N ] } 2 S F ( φ N ) T 2 d α ( t ) d t }
S F ( φ N ) T 2 d α ( t ) d t < < S F ( φ N ) α ( t )
S F ( φ N ) α ( t ) < < { 2 J 1 [ ψ N ] [ ψ N ] } 2
S F ( φ N ) T 2 d α ( t ) d t < < 2 { 2 J 1 [ ψ N ] [ ψ N ] } 2
e α ( t ) = v ( t ) v ( t T / 2 ) v ( t ) + v ( t T / 2 ) S F ( φ N ) { 2 J 1 [ ψ N ] [ ψ N ] } 2 [ α ( t ) T 4 d α ( t ) d t ] + n ( t )
n ( t ) [ n ( t ) n ( t T / 2 ) ] / [ 8 k ( 1 k ) R 2 r 2 P L O P S max ]
S N ( f ) e / [ 2 k ( 1 k ) R P S max ]
G S ( s ) = E α ( s ) A ( s ) = S F ( φ N ) { 2 J 1 [ ψ N ] [ ψ N ] } 2 ( T 4 s + 1 )
S N E A ( f ) = S N ( f ) | G L ( f ) / G S ( f ) | 2
G L ( f ) = G S ( f ) G F P A ( f ) / [ 1 + G S ( f ) G F P A ( f ) ]
N E A = 0 S N E A ( f ) d f = S N ( f ) B n
B n = 0 | G L ( f ) / G S ( f ) | 2 d f
S N R = | v I | 2 | n I | 2 = 4 R 2 r L 2 ( 1 k ) P L O P S ( t ) { 2 J 1 [ ψ ( t ) ] / [ ψ ( t ) ] } 2 S N I B
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.