You are on page 1of 59

Stimulated Brillouin scattering in optical bers

Andrey Kobyakov,1,* Michael Sauer,1 and Dipak Chowdhury2


1

Corning Incorporated, Science and Technology Division, SP-TD-01-1, Corning, New York 14831, USA Corning European Technology Centre, 7bis avenue de Valvins B.P. No. 3, F-77211 Avon, France author: kobyakova@corning.com

*Corresponding

Received June 19, 2009; revised November 3, 2009; accepted November 3, 2009; published December 17, 2009 (Doc. ID 113125) We present a detailed overview of stimulated Brillouin scattering (SBS) in single-mode optical bers. The review is divided into two parts. In the rst part, we discuss the fundamentals of SBS. A particular emphasis is given to analytical calculation of the backreected power and SBS threshold (SBST) in optical bers with various index proles. For this, we consider acousto-optic interaction in the guiding geometry and derive the modal overlap integral, which describes the dependence of the Brillouin gain on the refractive index prole of the optical ber. We analyze Stokes backreected power initiated by thermal phonons, compare values of the SBST calculated from different approximations, and discuss the SBST dependence on the ber length. We also review an analytical approach to calculate the gain of Brillouin ber ampliers (BFAs) in the regime of pump depletion. In the high-gain regime, ber loss is a nonnegligible effect and needs to be accounted for along with the pump depletion. We provide an accurate analytic expression for the BFA gain and show results of experimental validation. Finally, we review methods to suppress SBS including index-controlled acoustic guiding or segmented ber links. The second part of the review deals with recent advances in ber-optic applications where SBS is a relevant effect. In particular, we discuss the impact of SBS on the radio-overber technology, enhancement of the SBS efciency in Raman-pumped bers, slow light due to SBS and SBS-based optical delay lines, Brillouin ber-optic sensors, and SBS mitigation in high-power ber lasers, as well as SBS in multimode and microstructured bers. A detailed derivation of evolutional equations in the guided wave geometry as well as key physical relations are given in appendices. 2009 Optical Society of America OCIS codes: 190.5890, 250.4480, 290.5830, 290.5900, 060.2320, 060.4370.

Notation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Symbols Used in the Text. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . List of Acronyms Used. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1. Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .


Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 1943-8206/10/010001-59/$15.00 OSA

3 3 4 4
1

2. Key Physical Concepts of Inelastic Light Scattering. . . . . . . . . . . . . . . . 2.1. Coupled SBS Equations for Evolution of Guided Optical Power... 2.2. Noise Initiation of SBS. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3. SBS Threshold. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3a. Denitions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3b. Equations for SBS Threshold Power. . . . . . . . . . . . . . . . . . . . 2.3c. Threshold Dependence on Fiber Length. . . . . . . . . . . . . . . . . 2.3d. Origin of the Numerical Factor 21 in the Common SBST Formula. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4. Brillouin Fiber Ampliers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3. Fibers with Enhanced SBS Threshold. . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1. Index-Controlled Acoustic Guiding. . . . . . . . . . . . . . . . . . . . . . . . . 3.2. Segmented Fibers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3. Other Approaches to Suppress SBS. . . . . . . . . . . . . . . . . . . . . . . . . 4. SBS in Fiber-Optic Applications. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1. Radio-over-Fiber Technology. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2. SBS in Raman-Pumped Fibers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3. Slow Light and Optical Delay Lines. . . . . . . . . . . . . . . . . . . . . . . . . 4.4. SBS-Based Fiber-Optic Sensors. . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.5. High-Power Fiber Lasers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5. Other Fiber-Optic Issues. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.1. SBS in Multimode Fibers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2. SBS in Microstructured Fibers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Appendix A: Derivation of Evolutional Equations for Signal and Stokes Waves. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . A.1. Stationary Solution for the Density Variation. . . . . . . . . . . . . . . . . . . A.2. Equations for Evolution of Optical Power. . . . . . . . . . . . . . . . . . . . . . Appendix B: Approximate Analytical Solution for the SBST in the Short and Long Fiber Approximation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Acknowledgments. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References and Notes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

7 9 12 14 14 14 15 17 17 20 20 23 26 27 27 29 32 34 35 36 36 37 38 38 39 41 45 45 45

Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001

Stimulated Brillouin scattering in optical bers


Andrey Kobyakov, Michael Sauer, and Dipak Chowdhury

Notation
Symbols Used in the Text
The most frequently used symbols in this review are dened below, and corresponding units are given in square brackets. Dimensionless quantities are denoted [ ]. The SI unit system is used throughout the review. Quantities related to the optical power P can also be expressed in dBm 10 log10P mW. Italic subscripts denote running indices. Vector quantities are indicated in bold roman. The magnitude of a vector is given by the corresponding italic letter, e.g., B = B. Acronyms used in the symbol table are dened in the acronym list below.
Symbol Unit m1 m1 m1 W1 m1 W1 m1 W1 [] [m] [] s1 s1 s1 kg/ m3 [W] [] [rad/s] m 2 m 2 [m/s] [] [] [] [m/W] [J s] [J/K] [m] [] [] [] [] [] [W] Denition Loss coefcient of the optical ber Loss coefcient at the wavelength of Raman pump Peak SBS efciency for bers with a single dominant acoustic mode Peak SBS efciency corresponding to the mth acoustic mode Raman efciency Dimensionless SBS efciency Pump wavelength Fraction of the Stokes power relative to the pump power in the SBST denition Brillouin frequency shift for the mth acoustic mode, m = p S B for all m Pump frequency, p = 2p Stokes frequency, S = 2S Mean value of the material density Thermal noise power in the BGS bandwidth mth acoustic mode prole Acoustic phonon frequency (single dominant acoustic mode), = p S Optical effective area Acousto-optic effective area corresponding to the mth acoustic mode Speed of light in vacuum Optical mode prole Gain coefcient for the Stokes power in the noise initiated process Gain coefcient in the Brillouin amplier, PSz = 0 = GampPseed Peak Brillouin gain for the mth acoustic mode Plancks constant Boltzmanns constant Fiber length Lorentzian prole of the Brillouin gain as a function of frequency Effective refractive index of the ber Spontaneous emission factor in the SBS process Number of photons emitted in the backward direction Numerical aperture of the ber Pump (forward propagating) power

R m R p m p S 0 m Aeff ao Am c fr G Gamp gm h k L L n nsp N NA Pp

Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001

P0 PS Pseed Pth p12 r S T vA wm z

[W] [W] [W] [W] [] [m] [] [K] [m/s] s1 [m]

Input pump power, P0 = Ppz = 0 Stokes (backward propagating) power Input Stokes power into Brillouin amplier, Pseed = PSz = L SBST Component of the electrostriction tensor Radial direction in the cylindrical coordinate system Number of segments in a concatenated ber link Absolute temperature Acoustic velocity in the medium FWHM of the BGS peak corresponding to the mth acoustic mode Coordinate along the ber length

List of Acronyms Used


Brillouin ber amplier Brillouin gain spectrum dispersion-compensating ber differential phase-shift keying (modulation format) erbium-doped ber amplier error vector magnitude full width at half-maximum multimode ber MachZehnder modulator nonreturn to zero (modulation format) ordinary differential equation photonic crystal ber passive optical network radio frequency signal-to-noise ratio stimulated Brillouin scattering SBS threshold power undepleted pump approximation wireless local area network BFA BGS DCF DPSK EDFA EVM FWHM MMF MZM NRZ ODE PCF PON RF SNR SBS SBST UPA WLAN

1. Introduction
Molecular scattering became a subject of intensive research in the 1920s and 1930s. Today, scattering from optical phonons (quantized states of the lattice vibration) is known as the Raman process, while interaction of light with acoustic phonons is named after Lon Brillouin, who theoretically predicted light scattering from thermally excited acoustic waves in 1922 [1]. Besides investigations by Raman in India and Brillouin in France, molecular scattering was studied by Landsberg and Mandelshtam in Russia, Smekal in Austria, and Wood in the United States. Priorities of discoveries made at that time as well as the appropriateness of credits given are still being debated (see, e.g., [2,3], for a historical discussion). Brillouin scattering is one of the most prominent optical effects. In a spontaneous process, a photon from an incident light wave is transformed into a scattered photon and a phonon. The scattered wave is downshifted in frequency. It is called a Stokes wave after George Stokes, who found the frequency downshift in the process of luminescence in the 19th century. Typically, the scattering cross secAdvances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 4

tion of the Stokes light is quite low, but in optical bers light can propagate tens of kilometers without signicant attenuation. This makes (stimulated) Brillouin scattering a noticeable and often undesirable effect in optical bers. The scattered light has a certain angular distribution, but the ber geometry selects only two preferred directionsforward and backward. As will be discussed below, forward Brillouin scattering in optical bers is very weak. Therefore, the Stokes wave propagates mainly in the direction opposite to the input, or pump, optical wave. At a particular level of the pump power, the process becomes stimulated, i.e., strongly dependent on the pump power. This is characterized by efcient energy conversion from the input light to the backscattered wave. The most prominent origin of stimulated Brillouin scattering (SBS) is a physical phenomenon called electrostriction (see, e.g., [4]), which manifests itself in a variation of the mediums density by action of light. The backscattered Stokes light interferes with the input pump light and generates an acoustic wave through the effect of electrostriction. Effectively, the propagating light creates a moving density grating from which it scatters in the backward direction. Thus, the frequency downshift of the Stokes wave can also be explained by the Doppler effect. The light scattering mechanism is schematically shown in Fig. 1. With the increased intensity of the Stokes wave the interference pattern becomes more pronounced, and the acoustic wave increases in amplitude. The forward propagating acoustic wave acts as a Bragg grating, which scatters even more light in the backward direction.

Figure 1
input optical wave (pump) spontaneously backscattered (Stokes) wave input + reflected interference

acoustic wave due to electrostriction

stronger Stokes wave due to reflection from moving grating stronger interference

Spontaneous (top) and stimulated (bottom) Brillouin scattering. Backscattered (Stokes) light (blue) from acoustic noise interferes with the input (pump) wave (black). The interference pattern is shown in red. The abscissa of the curves is the coordinate along the medium length. The ordinate is the amplitude of the optical waves (black and blue curves) and the intensity of the interference (red). The amplitude of the acoustic wave is proportional to the optical intensity. The acoustic wave generated as a result of electrostriction further stimulates backscattering, which in turn enhances the interference between the pump and the Stokes waves and reinforces the acoustic wave.
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 5

stimulated scattering

spontaneous scattering

Although spontaneous Brillouin scattering was predicted in 1922, the stimulated process, when the acoustic wave is created by the light beam itself, was only observed in 1964 [5]. SBS is a nonlinear process, i.e., its efciency depends on the input power. The input signal power at which the Stokes wave power increases rapidly and may even be comparable with the input power is called the threshold power or simply the SBS threshold (SBST). Various fundamental and applied aspects of SBS were studied in the past. For example, the electrostrictive contribution to the intensity-dependent refractive index was investigated both theoretically [610] and experimentally [1114]. A detailed model of a temporal response of the bers refractive index is presented in [10,15]. The effect of the refractive index prole on the Brillouin gain spectrum (BGS) [1621] and on the magnitude of the Brillouin gain coefcient [2228] was also the subject of numerous studies. The effect of Ge doping on the acoustic damping coefcient of silica bers was studied in [29]. A very large Brillouin gain coefcient was found in chalcogenide glasses [30,31]. Other important topics related to Brillouin scattering include polarization properties of the scattered light and acousto-optic polarization coupling [3234], an interplay between SBS and nonlinear four-wave mixing [3538] or cross-phase modulation [39], as well as instabilities caused by the four-wave-mixingSBS interaction [40,41], multicascaded SBS supported by Rayleigh backscattering [42], SBS in distributed Er-doped ber ampliers (EDFAs) [4345] and in Raman ampliers [4650], the effect of spectrally broadened pump on scattering efciency [51,52], SBS in ber Bragg gratings [53,54], and dynamic behavior of SBS [5557], to mention only a few. The list of application areas where SBS becomes relevant is even more extensive. One of the most prominent examples is ber-optic telecommunications, where, for example, SBS may manifest itself through the electrostrictive interaction between solitons in optical bers [5864]. The impact of SBS on digital intensity-modulated signals is reviewed in [65,66], while SBS in amplitude-modulated cable television systems was studied in [6769]. Research in SBS remains an actively developing area of nonlinear optics with hundreds of papers published annually. It is certainly beyond the scope of a single review to appropriately cover all the aspects of SBS. Therefore, in this work we restrict ourselves to SBS in optical bers and in particular focus on several of the most recent advances in the eld. An extensive bibliography on the earlier work can be found in [7072]. A detailed description of the physics of Brillouin scattering including the quantum-mechanical treatment can be found in, e.g., [73,74], whereas topics related to ber-optic telecommunication aspects are covered in [75]. Our main goal in this review is twofold. First, we are going to discuss the specics of SBS in optical bers. One key difference from SBS in crystals comes from the extended spatial interaction between the pump and the Stokes waves in optical bers. As a result, one needs to account for optical loss in noise initiation of SBS or in Brillouin amplication. Another difference from the bulk interaction is due to the guided nature of both optical and acoustic waves in the ber. It was recently observed that the guiding of longitudinal acoustic modes by the ber core is a very important effect. The dependence of the acoustic mode prole on the radial variation of the refractive index due to doping in the ber core directly affects the Brillouin gain magnitude. Fibers with different index proles have different BGS and therefore different SBS thresholds [22,25,76]. Understanding the acousto-optic interaction in a cylindrical guided wave geometry alAdvances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 6

lows control of the Brillouin gain by index prole design. It should be emphasized that guiding of both optical and acoustic modes is critical for an accurate theoretical description of the scattering process. For completeness, mathematical details of the derivation of the governing equations are presented in Appendix A. The second goal in our work is to give an overview of novel technology areas where SBS plays an important role. These, for example, include highpower ber lasers, slow light, and optical delay lines for optical memory or SBS in Raman ampliers and in radio-over-ber transmission. We also briey discuss SBS in multimode and photonic crystal bers (PCFs) The review is organized as follows. In Section 2 we outline the physics of the scattering process and introduce key concepts and parameters. We analyze the coupled evolutional equations derived in Appendix A to calculate the SBST and gain of Brillouin ber ampliers (BFAs). Approaches to enhance the SBST of optical bers are discussed in Section 3, while several applications based on SBS are reviewed in Section 4. Other ber-optic issues such as SBS in multimode or microstructured bers are considered in Section 5. Section 6 concludes the paper.

2. Key Physical Concepts of Inelastic Light Scattering


As was mentioned above, Brillouin scattering is caused by modulation of the dielectric permittivity of the medium. The key distinction of Brillouin scattering from other types of molecular scattering is the acoustic type of the relevant lattice vibrations of the medium. For the acoustic mode, the direction of vibration of the two neighboring atoms is the same, i.e., atoms oscillate with a small relative phase shift. The type of dispersion relation of the acoustic mode (Fig. 2) determines several key features of the scattered light such as a small relative frequency shift 105. The polarization induced by modulation of the refractive index contains terms describing oscillations at the sum (anti-Stokes) and the difference (Stokes) frequencies. The anti-Stokes emission is much weaker than the Stokes emission. In addition, it requires a seed wave at the sum frequency. Moreover, during interaction with the incident wave, the anti-Stokes wave is attenuated, i.e., its energy is transferred to the incident pump wave [72,74]. In a quantum mechanical formulation, the process can be viewed as annihilation of the incident photon and creation of a scattered photon and an acoustic phonon. The energy and the momentum of the interacting quanta must be conserved, which results in a relation for the frequencies of the pump photon p, the Stokes photon S, and the acoustic phonon ,

S = p .
The momentum conservation requires

S = p B

for the corresponding wave vectors. These relations can be shown in the dispersion diagram (Fig. 2) on the coordinate plane , [73]. Conservation laws (1) and (2) then require a closed vector diagram for dispersion vectors S, p, and B.
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 7

Figure 2

(a)
p S

light line acoustic phonon


p

(b)
S

B
p

(a) Generation of backscattered light illustrated by a dispersion vector diagram [73], where participating photons and a phonon are represented by a vector on dispersion curves shown in the , coordinate system. Dispersion of the acoustic phonon can be approximated by linear relation (3). Dispersion of the pump photon is the light line = c. The right-hand half of the rst Brillouin zone is shown. A backward propagating (projections of vectors S and p on the abscissa have opposite directions) wave S , S originates from the interaction of a pump photon p , p and the acoustic phonon , B. The phonon dispersion curve originating at point (0,0) entails a strong dependence that is due to the momentum conservation shown in diagram (b), which is used to illustrate the scattering efciency in the backward direction = ; see Eq. (4).

The efciency of Brillouin scattering has a strong angular dependence. This happens because of the form of the dispersion relation of the acoustic phonon [Fig. 2(a)], which can be approximated by a straight line, v AB , 3

near the center of the rst Brillouin zone. In Eq. (3), vA is the acoustic velocity in a medium. The value of B, in turn, depends on the angle between the wave vectors of the pump and the Stokes waves [see Fig. 2(b)]. Substituting the value of B obtained from the triangle relations of Fig. 2(b) into Eq. (3), we obtain for the Brillouin frequency shift 2vA

pn
c

sin

where we used the approximation S p = pn / c, which is due to the small relative frequency shift of the scattered phonon, p,S. As can be seen from Eq. (4), the frequency shift depends on the scattering angle and is maximum for backward scattering = . For forward scattering = 0, the Brillouin frequency shift approaches zero 0. However, the guided wave geometry of the ber can lead to light interaction with transverse acoustic phonons, which makes forward scattering possible. This mechanism was rst studied in [77] and is now referred to as guided acoustic wave Brillouin scattering.
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 8

In Section 2.1 we present a detailed discussion of equations that describe the evolution of the guided pump and Stokes power caused by SBS in single-mode optical bers. We then study the noise initiation of the backward propagating Stokes power in Section 2.2 and introduce the concept of SBST in Section 2.3. To initiate SBS, one can also launch a small amount of optical power at the Stokes frequency from the opposite end of the ber. Such a system will then act as an amplier. BFAs are discussed in Section 2.4.

2.1. Coupled SBS Equations for Evolution of Guided Optical Power


Coupled SBS equations for the pump (forward propagating) and Stokes (backward propagating) light are extensively covered in the literature (see, e.g., monographs [70,71,73] and textbooks [72,74]). However, in all mentioned references and numerous research articles the derivation of evolutional equations is based on the plane wave approach; i.e., guiding properties of participating waves are not accounted for. The plane wave approximation is also used in the ber-optic textbooks [75,78,79] discussing SBS. It will be shown below why the plane wave approximation provides good accuracy for step-index bers. It is wellknown that acoustic waves can be guided in a solid cylinder [77,8084] and in a double-clad (berlike) structure [8589]. Peral and Yariv [90] considered both acoustic and optical guiding to describe the SBS process. However, their analysis did not take into account the radial variation of mechanical properties of glass due to the refractive index variation in the core region. A more rigorous analysis accounting for the acoustic guiding by the doped core region was performed in [2225,28].

Strictly speaking, in deriving evolutional SBS equations the plane wave approach is applicable only for a bulk medium. In optical bers, acoustic guiding effects play a crucial role and need to be accounted for to accurately describe the scattering process.

Below we analyze a set of coupled ordinary differential equations (ODEs) for the spatial evolution of guided optical powers of the input pump Pp and the backreected Stokes PS wave (for a detailed derivation see Appendix A): dPp dz

= mL P pP S P p ,

dPS dz where

= mL P pP S + P S ,

Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001

m =

gm
ao Am

is the peak SBS efciency for the acoustic mode and is the optical loss coefcient of the ber. The peak efciency m is inversely proportional to the acousto-optic effective area (see Appendix A):
ao Am =

f 2r mrf 2r

2 2 m r ,

where fr and mr are radial proles of the fundamental optical and the mth acoustic modes of the ber, respectively. Angular brackets denote averaging over the transverse cross section of the ber. Each acoustic mode is responsible for a spectral feature in the BGS (Fig. 3). Unlike the optical effective area, dened as [75,79] Aeff = f 2r2 f 4r 9

and conventionally used to characterize SBS in optical bers, the quantity given by Eq. (8) actually determines the total Brillouin gain. As we will see in Section 3, the acousto-optic effective area approximately equals Aeff for ber proles that are close to step index. This explains why approximating Aao with Aeff gives good agreement with experimental data when step-index, standard single-mode bers are ao used. However, for nonuniform ber proles, the parameter Am rather than Aeff determines the strength of the acousto-optic interaction and is responsible for different SBS thresholds in optical bers with different index proles. For example, counter-

Figure 3

index profile
n r

acoustic velocity
vA r PS
noise-initiated Stokes wave

pump

Pp

g1
power

g2
w2
p- 2

w1
p- 1

frequency

Schematic representation of spectra of the pump (red) and the Stokes (blue) waves in an optical ber with a non-step-index prole. In this example, the BGS has two peaks due to excitation of two dominant acoustic modes with frequency shifts 1 and 2. Refractive index nr and acoustic velocity vAr proles of the ber are schematically shown in the top diagram to emphasize the guided nature of optical and acoustic waves in the ber.
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 10

intuitive results such as a higher SBS threshold for bers with a smaller effective area were obtained [25,76].

The overlap integral between the optical and acoustic modes is responsible for different SBS thresholds of optical bers with different index proles.

A related metric for the acousto-optic modal overlap was adopted in [25]. A diao mensionless overlap integral Im introduced there is the ratio of the optical and the ao ao acousto-optic effective areas Im = Aeff / Am . This quantity (or, effectively, the acoustooptic effective area) has been used in the full numerical modal analysis [27] to calculate BGS of various bers. Results obtained for F-doped step-index bers showed very good agreement with measured data. The concept of the acousto-optic modal overlap was shown to be also applicable to ErYr doped bers. Good agreement between measured and calculated BGS was demonstrated in [91]. A 2D nite-element analysis [92,93] with the acousto-optic effective area used to calculate Brillouin gain corresponding to different acoustic modes demonstrated accurate prediction of BGS resonances for standard single-mode and PANDA (polarization-maintaining and absorption-reducing) bers. Similar analysis has been used to employ L01 and L03 acoustic modes in w-shaped triple layer bers for strain and temperature sensing [94]. The numerator of Eq. (7) is the peak Brillouin gain of the mth acoustic mode gm, 4 n 8p 2 12 c3 p 0 mw m 10

gm =

where n is the effective refractive index of the ber, p12 is the respective component of the electrostriction tensor, m and wm are the frequency shift and the FWHM width of the mth line in the BGS, respectively, p is the pump wavelength, c is the speed of light, and 0 is the mean value of the material density of the ber. A typical value of m for most germania-doped bers is 11 GHz. For most bers, it varies only slightly (by 0.5 GHz) for different acoustic modes. Hence, without loss of accuracy, one can use m = B for all acoustic modes in Eq. (10). The nonlinear term in Eqs. (5) and (6) is multiplied by the spectral prole of the Brillouin gain L, which has the Lorentzian shape L = wm/22 p + m2 + wm/22 , 11

where p = c / p is the pump frequency. In the discussion of noise-initiated Brillouin scattering, we will assume the presence of a dominant acoustic mode ( = 1 = g1 / Aao 1 m, m = 2 , 3 , 4. . .), which is typical for single-mode bers with a quasi-rectangular index prole. This assumption will be relaxed in Section 3 when we will be discussing high-SBST bers. Finally, several additional terms can be introduced in Eq. (7) to account for polarization effects or the nite spectral line width wlas of the input signal,
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 11

m =

ao Am K 1+


wlas wm

gm

12

For a polarization-scrambled pump, the polarization factor K = 3 / 2 differs from 2 because of different Brillouin gains for spontaneous photons with a polarization identical with and orthogonal to the pump [34]. However, for high pump powers the gain is much larger for copolarized Stokes and pump waves. At larger Brillouin gains, SBS acts like an almost perfect polarization reector with a degree of polarization close to 100% [32]. The term in parentheses in Eq. (12) shows that for a pump laser with a large spectral width wlas wm the Brillouin gain coefcient is reduced [18,66].

2.2. Noise Initiation of SBS


To evaluate the critical input power, or SBST, one has to calculate the total backreected power at the ber input, z = 0. This is a difcult task because the scattering process starts spontaneously from noise; i.e., the boundary value of the Stokes wave PSL, with L being the ber length, is undetermined in ODEs (5) and (6). There are two standard models of noise initiation of SBS. In the framework of the localized, nonuctuating source model, one assumes the presence of the Stokes seed wave launched at the rear side of the lossless medium [71] or at the point where loss exactly compensates for the Brillouin gain (the so-called single-photon approach discussed in [49,95]). According to a more rigorous distributed, uctuating source model, the backscattered wave originates from thermally excited spontaneous phonons [49,96101]. In [97,98,100], where this model was applied for short bulk media, the loss of the medium was ignored. However, the contribution of loss is a very important effect in optical bers because the pumpStokes interaction occurs over long distances and both the pump and the Stokes waves can be attenuated by orders of magnitude. For example, on propagation in a 50 km-long standard single-mode optical ber the input power is reduced to a level of 10% from its original value. In [99,101] the contribution of loss was accounted for to calculate the power spectral density of the Stokes light. The initiation of Brillouin scattering is accompanied by a low pump-to-Stokes conversion efciency, and the undepleted pump approximation (UPA) can be applied to solve the system of ODEs (5) and (6). Assuming the UPA, one can neglect the nonlinear term in Eq. (5) to nd an approximate solution for the pump evolution as Pp = P0ez, which can be substituted into Eq. (6) to yield dPS dz = L P 0P Se z + P S , 13

where is the gain parameter corresponding to the dominant acoustic mode. Equation (13) can be rewritten for the photon occupation number as [95] dN dz = LezP0N + nsp + N , 14

where nsp = 1 + ehB/kT 11 kT / hB and k and h are Boltzmanns and Plancks constants, respectively; T is the ber temperature.
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 12

Solution of Eq. (14) with the boundary condition NL = 0 (i.e., without initial spontaneous Stokes photons) gives the SBS gain for a single polarization component as G = N0 nsp = expL1 eL

+ e L 1

, 15

where the dimensionless parameter

g 1P 0 Aao 1

P0

16

can be viewed as a normalized input power or the strength of the SBS process. The next step in obtaining the total generated Stokes power PS0 is the integration of Eq. (15) over the whole frequency spectrum. Assuming full polarization scrambling of the pump, one obtains PS0 = 2nsp

hGd =

2kT

Gd .

17

The same expression as Eq. (15) for the spectral density of the backscattered light was obtained in [99,101]. However, integration over the frequency to calculate the total Stokes power was believed to be possible only numerically. The analytical form of the integral in Eq. (17) was found in [102]. The result of integration in Eq. (17) can be written as P S 0 = 4 3 eq/2 q 1 +


e L 2 q q I0 2 I1 2 q = 1 e L , = kTpw1 21

1 e L I 1


q 2

18

where Il are the modied Bessel functions of the order l, 19

and 20

is the effective noise power per pump polarization state in the bandwidth corresponding to the BGS of the dominant acoustic mode. For standard single-mode ber (see Table 1) at room temperature and a pump wavelength of p = 1.55 m it amounts to 0.7 nW.
Table 1. Parameters of the BGS for Several Single-Mode Optical Fibersa
Fiber Fiber I Fiber II Fiber III
a

1 / km
0.046 0.046 0.046

1 [GHz]
10.87 10.66 10.8 [110]

w1 [MHz] 20 30 25

m1 W1
0.14

Fiber I is standard single-mode ber, ber II is Corning LEAF ber, ber III is Corning SMF-28e+ optical ber with NexCor technology.

Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001

13

2.3. SBS Threshold


2.3a. Denitions Although the SBS threshold power (SBST) is variously dened in the literature, most denitions share the same conceptual approach. Namely, the output Stokes power is compared with some fraction of the maximum signal power. Because of the exponential dependence of the Stokes power on the input pump power near the threshold, the exact value of is not critical (Fig. 4). Denitions with = 1 [75,95] or = 0.01 [103106] are common. The idea of all these rather qualitative criteria is that the Stokes power begins to increase rapidly and starts to approach the input power so that higher-order Stokes waves can be generated [107]. For generality, in the derivations below we consider a free parameter and use the value = 0.01 to calculate the SBST. Thus, the SBST condition is PS0 = P0 , 21

where the left-hand side of Eq. (21) nonlinearly depends on the input pump power P0 according to Eq. (18) [q is proportional to P0, as follows from Eqs. (19) and (16)]. 2.3b. Equations for SBS Threshold Power The solution of Eq. (21) for P0 is the sought SBST power. It is more convenient, however, to express the right-hand side of Eq. (21) through q, using (19) and (16), and to solve Eq. (21) for q. The threshold power is then q Pth = where 1 e L Leff = is the effective length of the ber. 23 , 22

Leff

Figure 4
backreflected power PS(0), dBm 20 10 0 10 20 30 40 10 5 0 5 10 pump power P0, dBm 15 20
=0.001 =0.01 =0.1

Denition (21) of the SBST for several values of parameter . Measured (dots) dependence PS0 versus the input pump power P0 is shown together with the calculated UPA curve (blue). The abscissa of the intersection point is the threshold power. For standard single-mode ber, the UPA is valid for 0.1.
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 14

A good approximation for Eq. (18) can be obtained if asymptotic expansions of the modied Bessel functions [108] I 0 x ex

2x


1 1+ 8x
L

I 1 x

ex

2x


3 1 8x .

24

are used. The threshold equations (21) and (18) reduce to

3/2e1e

1 eL

e L +

1 2

25

Nearly the same equation is obtained if the steepest descent method (see, e.g., [109], pp. 477484) is used to perform integration in Eq. (17) [49,102]. The only difference compared with Eq. (25) is the second term in the parentheses, which becomes 1 / . For not very long bers (L 40 km, 0.2 dB/ km), the term 1 / 2 in (25) can be neglected, and the equation can be rewritten as q3/2eq = 2 e L 1 e L . 26

The approximate analytical solution of Eq. (26) is (see Appendix B)

q 1+


3 2 ln 3

27

where

= ln

e L 1 e L .

28

Next, Eqs. (27) and (28) can be substituted into Eq. (22) to obtain the value of the SBST. 2.3c. Threshold Dependence on Fiber Length Figure 5 shows the measured dependence of the Stokes power PS0 on the input power P0 for various ber lengths. Expressions derived in the previous section can be used for studying the SBST dependence on the ber length as well as for comparing the accuracy of different approximations. Results for standard single-mode ber are shown in Fig. 6(a), where the calculated and measured SBST power Pth is plotted as a function of the ber length. Parameters used in calculations are shown in Table 1. As can be seen from Fig. 6(a), all approximations, including the analytical solution (27), work well up to a ber length of L 40 km. After that distance, the accuracy of the short-ber approximation decreases because the term eL in Eq. (25) becomes comparable with the term 1 / 2, which was discarded in the short-ber approximation (26). For higher ber loss (for example, for operation at shorter wavelengths) the short-ber approximation remains accurate for L 20 km [Fig. 6(b)]. The asymptotic expansion of the modied Bessel functions works well for all ber lengths and attenuation coefcients. The error of the approxiAdvances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 15

Figure 5
10

reflected power, dBm

0 10 20 30 40 6 8 10 12 14 16
25.3 km 20 km 15 km 10 km 5 km 2 km

input power, dBm

18

20

Reected Stokes power PS0 measured as a function of the input power P0 for various ber lengths. The threshold power can be obtained from denition (21), which is shown by a dashed line = 0.01.

Figure 6

(a) SBST Pth in standard single-mode ber. Pth is calculated exactly from Eqs. (21) and (18) as well as by using several approximations: the asymptotic approximation of Bessel functions I0,1, Eq. (25); the short-ber approximation (26); and analytical formulas (27) and (28). Measured values of SBST obtained from Fig. 5 are shown by lled circles. (b) Same parameters as in (a) but the ber loss is increased to = 0.5 dB/ km.
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 16

Figure 7
error of analytical formula, dB 2 1.5 1 0.5 0 0
=0.2 dB/km =0.5 dB/km

20

40 60 fiber length L, km

80

100

Difference in SBST (decibels) between the exact value calculated from Eqs. (21) and (18) and the value obtained from analytical formulas (27) and (28).

mation does not exceed 0.3 dB. The accuracy of analytical formula (27) is quantied in Fig. 7 for both values of the loss coefcient considered in Fig. 6. As expected, the accuracy of the analytical approximation decreases with the increased ber length and/or ber loss (i.e., the product L). 2.3d. Origin of the Numerical Factor 21 in the Common SBST Formula To complete the discussion on calculation of the SBST power, we briey review the widely used approximation (see, e.g., [31,75,95,111114]) Pth = 21 Aeff gB , 29

where gB = g1 is the peak Brillouin gain for the dominant acoustic mode. First, as was already mentioned, the optical effective area Aeff should be replaced with the acousto-optic effective area Aao 1 . Formula (29) was rst derived by Smith [95] in 1972. The assumed ber loss there was 20 dB/ km, and terms with eL in Eq. (15) could be safely discarded. After use of the steepest descent method in (17), Eq. (25) was obtained with the term eL + 1 / 2 replaced with 1 / . As a result, the threshold equation took the form

5/2e =

gB
Aeff

30

where = 1 was assumed and power in only a single polarization was considered. Next, with p = 1.06 m, 1 = 16.6 GHz, w1 = 50 MHz, [95] one obtains = 1.8 nW, and for = 5 105 cm1, gB = 3 109 cm/ W, and Aeff = 107 cm2, , also taken from Smiths paper, the right-hand side of Eq. (30), which we denote 1 + 5 / 2ln / 5 / 2 gives equates to 19.1 107. Then from Eq. (B.4) 21.1, and from the denition of [Eq. (16)] one obtains the equivalent of Eq. (29). We also note that several authors [50,103,104,106,115] suggested using a smaller numerical factor such as 17 or 18 in Eq. (29).

2.4. Brillouin Fiber Ampliers


SBS can be used for efcient narrowband amplication when the seed Stokes wave is input from the rear (opposite to the pump) end of the ber. BFAs have
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 17

applications in microwave photonics, radio-over-ber technology, and beroptic sensing. For example, BFAs can be used to achieve gain in signal conversion in microwave photonic systems [116,117] or in the realization of a shapeadjustable narrowband optical lter [118,119]. The same principle can also be applied to carrier depletion for increasing the modulation depth of the microwave signal [120]. Optical carrier ltering in radio-over-ber systems was shown to signicantly increase the dynamic range [121] and decrease the loss [122] of microwave ber-optic links. In addition, BFAs proved to be useful for generation of millimeter-wave signals [123,124]. Most recently, BFAs were exploited as tunable slow-light delay buffers [125]. Unlike the case of spontaneous Brillouin scattering, the system of ODEs (5) and (6) for BFAs has well-dened boundary conditions: Pp0 = P0 and PSL = Pseed. Such a mathematical problem is known as the two-point boundary value problem. For systems of nonlinear ODEs, it is typically addressed numerically (see, e.g., [126]). The exact solution to the boundary value problem of Eqs. (5) and (6) is known only for lossless media [72,75,127], which cannot be used for BFAs, where typically wave interaction occurs over tens of kilometers and loss is a signicant effect. Several attempts have been undertaken to nd a general analytical solution to SBS equations in a lossy medium [101,128]. A conserved quantity lnPpPS

Pp PS = const,

31

of the system of ODEs (5) and (6) reduces the problem to a single equation [128], which, however, can only be integrated numerically. In another approach [101], it was proposed to reverse the sign of the loss term in one of the ODEs to make the approximate set of equations integrable. This results in a system of two transcendental equations to be solved numerically with no closed-form solution possible. Interestingly, a quite straightforward solution to the coupled ODEs can be obtained for copropagating scattered waves when the sign of both terms on the right-hand side of Eq. (6) is reverted [129]. As another simplication, the UPA can be used. However, the above-mentioned applications typically require pump powers above the SBS threshold [116,120,122,124] so that the pump becomes depleted and the UPA strongly overestimates the Brillouin gain. A straightforward method to improve the accuracy of the UPA is a perturbative calculation of a rst-order correction to the UPA solution for the Stokes wave. Such an approach works well for a ber Raman amplier [130], which is described by a similar set of ODEs. One can show, however, that for SBS the perturbative correction to the UPA contains an exponential integral function that quickly diverges with increasing pump power.

If the perturbation approach in Brillouin amplication is applied to loss rather than to the nonlinear term, a closed-form analytical solution to the coupled BFA equations can be obtained

Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001

18

The closed-form approximate analytical expression for Brillouin gain in the depleted pump regime is [131] Gamp = P0 Pseed

+ ln1 /u u

e L ,

32

where u = P0L, = lnPseedL, and the BFA gain is dened as the ratio of the amplied and the launched Stokes power, Gamp = PS0 / Pseed. It is assumed that the wavelength of the seed Stokes wave corresponds to the peak BGS frequency, i.e., L 1 in Eqs. (5) and (6). For comparison, the BFA gain calculated from the UPA is GUPA amp = exp L + u

1 e L

33

Figure 8 compares the BFA gain obtained analytically from Eqs. (32) and (33), numerically from integration of Eqs. (5) and (6), and experimentally by using the setup shown in Fig. 9. A very good agreement between predictions of the analytical formula (32) and the measured gain can be seen in the high gain regime. The accuracy of Eq. (32) decreases with increased ber length but remains good for BFAs shorter than 20 km. This is a typical length for many applications [116,118,120,123]. Figure 8 also shows that for the weak-pump regime, the UPAbased estimation for the Brillouin gain, Eq. (33), can be used. However, above some critical power P Pcr + 2 + 4 / 2L Eq. (32) should be used instead.

Figure 8

Measured and calculated gain of a 10 km long BFA based on standard singlemode ber. The power of the input Stokes wave is (a) Pseed = 33 dBm and (b) Pseed = 18 dBm.
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 19

Figure 9
FBG VOA 10.877 GHz 1549.5 nm Tunable Laser OA circulator VOA coupler MZM OA Electrical Spectrum Analyzer PM PM fiber PM

PM

Experimental setup for BFA measurements: MZM, MachZehnder modulator; PM, powermeter; VOA, variable optical attenuator; OA, optical amplier; FBG, ber Bragg grating used to select the upper modulation sideband as a Stokes seed signal.

With the increased Stokes seed power, the ampliers gain begins to saturate earlier, and the maximum gain decreases compared with the case with smaller Pseed [cf. Figs. 8(a) and 8(b)]. In the experimental setup (Fig. 9), a tunable laser with p = 1549.5 nm was used together with an EDFA to generate up to 70 mW of pump power. The seed wave was obtained by using a MachZehnder modulator (MZM). A ber Bragg grating in the reecting regime was used to select a sideband of the MZM output. Powermeters were used to monitor the input pump power P0, the transmitted pump power PpL, the launched seed power Pseed, and the amplied Stokes power PS0. The electrical spectrum analyzer was used to determine the Brillouin frequency shift of the ber and, correspondingly, the modulation frequency for the MZM (10.88 GHz for standard single-mode ber).

3. Fibers with Enhanced SBS Threshold


Some ber-optic applications such as those discussed in the previous section require high Brillouin gain to amplify the Stokes signal. In the majority of cases, however, SBS is an undesired effect, since it prevents launching maximum optical power into the ber. In these cases, it is desirable to decrease the Brillouin gain coefcient or, equivalently, increase the SBS threshold of the ber. In this section, we review several approaches that result in enhanced SBST of the ber.

3.1. Index-Controlled Acoustic Guiding


From Eqs. (7), (8), and (10) one can conclude that one of the strongest control parameters is provided by the acousto-optic effective area, Eq. (8). Even though the optical mode prole might change only slightly from one single-mode ber prole to another, acoustic modes and corresponding acousto-optic effective areas can vary signicantly. The SBS threshold in turn will depend on the ber index prole. To calculate the acoustic mode prole for a given index prole nr one has to solve the system of equations (A.12) and (A.13) for the acoustic mode
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 20

prole r and the eigenfrequency . The longitudinal sound velocity prole can be obtained from the following empirical relation [132134]: vAr = 59441 0.078%r , 34

where %r = 100nr nclad / nclad and nclad is the refractive index of the bers cladding. The optical mode prole can be obtained by solving Eq. (A.21). The results of calculations for bers I, II, and III are shown in Fig. 10. The obtained modal proles can then be used to calculate the corresponding acousto-optic effective areas by using Eq. (8). Results for ve single-mode germania-doped bers are given in Table 2, where we also list the calculated and measured values of the SBST power Pth for 20 km long bers.

Figure 10

Proles of the optical fr and rst three m = 1 , 2 , 3 acoustic mr modes for ao three single-mode optical bers. The corresponding numerical values for Am and Pth are listed in Table 1. (a) Fiber I, (b) ber II, (c) ber III.
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 21

Table 2. Calculated Acousto-optic and Optical Effective Areas m2 and Calculated and Measured SBST Pth [dBm]a
Fiber Fiber I Fiber II Fiber III Fiber IV Fiber V
a

Aao 1 91.5 124.4 178.5 108.2 112.0

Aao 2 3928 274.8 206.9 751.0 161.8

Aao 3 4921 842 1539 1641 2272

Aeff 84.4 73.5 85.4 88.0 61.7

Pcalc th 8.1 9.6 11.2 8.9 9.1

Pmeas th 8.1 9.7 11.5 9.2 9.6

For single-mode optical bers from Table 1 and two Ge-doped single-mode specialty bers (bers IV, V). All bers are 20 km long.

Calculation of the SBST has been performed by summing up reected power contributions from different acoustic modes, i.e., extending Eq. (17) to the case of multiple acoustic modes: P S 0 =

m=1

2T

Gmd = PS0 ,

35

where Gm is the gain coefcient of Eq. (15) corresponding to the mth acoustic mode. Accurate values of gm are difcult to obtain theoretically. We therefore make use of relative calculations of the SBS threshold. From the measured value of Pth for a reference ber I, which has only one dominant acoustic mode and thus a single peak in the BGS, we obtain the value g1,ref from Eq. (26) [also using Eqs. (19) and (16)]. Substitution of Eq. (15) into Eq. (35) with the subsequent integration by the steepest descent method results in the following transcendental equation for the SBST:

1 eL m =1

e L

exprmkx1 eL

rmk

=x

3/2

2g1

Aao 11

36

ao where rmk = mk / 1,ref w11Aao 11 / wmkAmk and the index k denotes the ber type (see Table 2), while the rst index m denotes, as before, the number of the acoustic mode; x = Pthg1,ref / Aao 11. The solution of Eq. (36) for x gives the SBST power for bers IIV from Table 2 as

Pcalc th =

Aao 11 g1,ref

37

We therefore assumed that the relative strength of the SBS interaction due to each acoustic mode is determined by the ratio rmk. Measured SBST powers were obtained from the corresponding reected versus input power curves (Fig. 11). The experimental setup is essentially the same as shown in Fig. 9, except no Stokes seed power is input to the ber. All studied bers have approximately the same attenuation coefcient = 0.2 dB/ km. As can be seen from Table 2, Aao 1 for ber I is smaller than that for ber II. This is a consequence of a weaker overlap between acoustic and optical modes. It leads to a smaller Brillouin gain coefcient and consequently to a higher SBST power of ber II despite its smaller optical effective area compared with ber I. Among the ve
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 22

Figure 11
20 fiber I fiber II fiber III

reflected power, dBm

10 0 10 20 30 40 0

input power, dBm

8 10 12 14 16 18 20

Measured reected versus input power for the rst three bers in Table 2. Each ber length is 20 km. Denition (21) with = 0.01 is shown as a dashed line.

considered single-mode bers, ber III has the largest acousto-optic effective area and thus the highest SBST. Even though its second acoustic mode 2r has a comparable overlap with the optical mode, the SBST power remains high because of the exponential dependence of Brillouin gain on the input power [see Eq. (25)]. A particular prole design can be used to increase the SBST. The increase in the threshold value depends on the ber type, since a change in the index prole also affects other important ber parameters such as dispersion, loss, and bending sensitivity. For standard single-mode ber (ITU G.652 compatible), an up to 3 dB threshold increase is possible [22]. This value can be even higher for some specialty bers such as highly nonlinear bers [135]. Finally, we note that other dopants (e.g., Al, F) used to modify the refractive index of pure silica may differently affect the acoustic guiding. For example, F-doped silica has a smaller refractive index than pure silica and is therefore used as the cladding material. Since the acoustic velocity in F-doped silica is also reduced [19], bers exploiting F doping have different acoustic properties compared with Ge-doped bers. A recent experimental study [136] demonstrated that just using F doping does not increase the SBST of the ber.

3.2. Segmented Fibers


As was shown in Subsection 2.3c, the SBST of an optical ber increases with decreased ber length. This fact can be used to make bers with increased SBST by changing the Brillouin frequency shift m along the ber length [137139]. In doing so, one hinders the exponential growth of the backreected power. Indeed, if BGS before and after some location z1 along the span length L 0 z1 L do not overlap, the amount of Stokes power PSz1 generated in the segment z1 , L is not further amplied but is attenuated in the segment 0 , z1 (see Fig. 12). SBS in nonuniform bers has been studied experimentally [76,111,137140] and theoretically [140,141]. The analysis of Subsection 2.2 can be extended to nonuniform bers consisting of ber pieces with different BGS. For a ber span consisting of S segments (the ith segment has length zi zi1) with nonoverlapping BGS, the total Stokes power can be found from Eq. (17), where contributions from each ber segment are summed [140]:
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 23

Figure 12

index profile fiber A


n r PS,A
pump

index profile fiber B


n r PS,B
independent Stokes waves in each fiber piece

Pp

z1

power

p- 1,B

p- 1,A

frequency

Schematic of SBS in a segmented ber consisting of two ber pieces with different BGS. The pump spectrum is shown in red; the Stokes spectra of bers A and B have a single peak (one dominant acoustic mode) and are only weakly overlapping; so the generation of Stokes light starts anew from noise in each ber piece.

P S 0 =

2kT

G,iezi1d .
i=1

38

In Eq. (38) G,i is the Brillouin gain in the ith segment, and the loss coefcient is assumed equal for all segments. The exponential factor in Eq. (38) accounts for attenuation of the Stokes power generated in the ith segment in all subsequent segments for backward propagating power. With the steepest descent integration, one can obtain the threshold condition as 2 3/2 =
i=1 S

ezi1 ezi zi expti

tiezi1 ezi

39

1 with being the largest gain coefcient among all segments, where ti = i / so that ti = 1 corresponds to the segment with the largest Brillouin gain coefcient. The results of the theoretical analysis performed for two-segment bers S = 2 are shown in Fig. 13 together with measured SBSTs. The SBST is plotted as a function of the length of the rst segment z1 for a constant L = 20 km total link length (i.e., the second segments length is 20 z1 km). Figure 13(a) shows results for concatenation of ber I and ber II, while ber links of Fig. 13(b) consist of pieces of ber I and ber III. A good agreement between theory and experiment is seen for both congurations. The SBST of a segmented link increases compared with the threshold value of the highest-SBST ber of the same length. The SBST increase is about 2 dB for both ber combinations. Similar results have been reported for concatenation of standard single-mode ber with zero-water-peak pure-silicacore ber [142].
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 24

Concatenation of just two bers with nonoverlapping BGS in the optimum span design may result in a several decibel SBST increase compared with a uniform ber of the same length.

Figure 13 also shows that at a particular length of the rst segment z1 the SBST increase is maximum. This optimum length of the rst span can be estimated as [140]

zopt 1 =

ln

t1 + t2 t 1 + t 2e L

40

The BGS of segmented ber links have features of both ber types (Fig. 14). Thus, an overlap of the BGS of ber I with a small [40 dB less than the main peak, see Fig. 14(a)] peak of ber II reveals a spectrum with nearly equal peaks (pink curve in Fig. 14(b)). Note that the BGS of concatenated ber links depends on the direction of the power launch, i.e., links ber I + ber II have a different BGS from that of ber II + ber I even though both segments have equal lengths LI = LII. The difference in BGS becomes more pronounced (especially at the high pump power) if the BGS structure of each individual segment is substantially different, as is the case of ber I and ber III (Fig. 15). Finally, it is interesting to see how the number of segments in the link affects the achievable SBST increase. Figure 16 shows that if the total length of the segmented link is small, increasing the number of segments beyond S = 2 only marginally increases the SBST. If, however, the segmented link is long, a larger number of segments might be needed to maximize the SBST (Fig. 17). As was mentioned above, the segment sequence strongly affects the increase in SBST. Practical implementation of a high SBST multisegment ber link for optical communication was reported in [143].

Figure 13

SBST of two-segmented bers plotted as a function of the length z1 of the rst segment (where the pump power is input) of a ber link consisting of two concatenated single-mode bers. The length of the second segment is 20 z1 km, so that the total link length is constant and equals L = 20 km; (a) concatenation of ber I and ber II, (b) concatenation of ber I and ber III. Theoretical curves are obtained from Eq. (39); scattered data show measured SBST values.
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 25

Figure 14

(a) BGS of bers I and II measured with the electrical spectrum analyzer. The length of each ber is 20 km. The input power is Pp = 8.1 dBm for ber I and Pp = 9.6 dBm for ber II, corresponding to the SBST value for a given ber length; (b) BGS of concatenated spans of 10 km ber I + 10 km ber II (point A in Fig. 13(a)) and 10 km ber II + 10 km ber I (point B in Fig. 13(a)). The input power is 9.5 dBm for both congurations.

3.3. Other Approaches to Suppress SBS


Apart from the discussed index control of the ber core and using segmented bers, a number of approaches can be employed to reduce the SBS efciency. One common method consists in broadening of the pump laser spectrum by using

Figure 15

BGS of concatenated spans of the total length of 20 km for several values of the input power Pp: (a) 5 km ber I +15 km ber III (point C in Fig. 13(b)), (b) 15 km ber III + 5 km ber I (point D in Fig. 13(b)).

Figure 16

SBST of a 20 km long ber link consisting of various numbers S of equal-length segments of alternating ber I and ber II: (a) ber I + ber II + ber I + ; (b) ber II + ber I + ber II + .
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 26

Figure 17

Increase in SBST versus total link length consisting of various numbers M of pairs of equal-length segments of alternating ber I and ber II; (a) ber I + ber II + ber I + ; (b) ber II + ber I + ber II +.

phase modulation [144]. Then, according to Eq. (12) the Brillouin gain decreases. This method is widely used in passive optical networks (PONs), e.g., for delivering a cable TV signal. A number of approaches are based on variation of ber parameters such as strain, temperature, or core radius along the ber length. For example, a strain distribution in the process of ber cabling expanded the Brillouin gain bandwidth from 50 to 400 MHz, which increased the SBST by 7 dB [145]. In another study, a stair ramp strain distribution resulted in 8 dB SBST increase in a 580 m dispersion-shifted ber [146]. The effective Brillouin gain was reduced by 3.5 dB by making a core radius nonuniform along the ber length [147]. In such an approach, one utilizes the dependence of the acoustic resonance frequency on the core radius. SBST in a short, highly nonlinear ber was increased threefold by applying a temperature distribution with a 140 C temperature gradient [148]. Another method of SBST enhancement consists in changing the dopant concentration along the ber length [111,149]. The SBST increases nonlinearly with the dopant concentration and amounts, e.g., to 10 dB for = 0.15% when GeO2 is used as a dopant [149]. As a further design technique for SBS suppression, an acoustic antiguide structure was proposed in [19]. This structure can be formed by doping the ber cladding with F, which results in a decreased velocity of the acoustic wave in this area relative to the ber core. However, it was found that this technique is limited by cladding acoustic modes, which propagate along the corecladding interface. Several other techniques are used to increase the SBST of ber lasers. These are discussed in Subsection 4.5.

4. SBS in Fiber-Optic Applications


4.1. Radio-over-Fiber Technology
Transmission of radio signals over optical bers has long been recognized as an efcient method of RF signal distribution over longer distances (see, e.g., [150] for an overview). While such ber-radio systems are often designed as point-topoint links [151], there has been increasing interest in exploring access network architectures where wireless services are distributed to subscribers homes from a central ofce (see Fig. 18 for a typical scenario). These services may comprise 2G/3G cellular, WiMAX (Worldwide Interoperability for Microwave Access), wireless local area network (WLAN), or other wireless signals. Such distribuAdvances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 27

Figure 18

Network Network Access Access Point Point (splitter) (splitter)

Feeder Feeder ~20 ~20 km km Distribution Distribution Central Central Office Office
Local Local Convergence Convergence Point Point

Drop Drop

An example of an access network architecture. Optical signal is distributed from the central ofce through a feeder ber and then transmitted further from the local convergence point to network access points and premises.

tion systems are typically based on PON architectures with high optical splitting ratios close to the subscribers and up to 20 km signal distribution over singlemode ber. In order to overcome high optical splitting ratios (typically 32 to 64 ) and maintain sufcient power levels at the receiving end, very high optical launch power levels are required. However, SBS plays an increasing role at higher power levels and can limit the performance of such networks and degrade the signal quality. Recently, there have been investigations of PON systems for radio signal distribution that target applications like 3G cellular and WiMAX service distribution [152154], and SBS has been found to be a key limiting factor. A high-SBST ber was shown to be benecial for performance of hybrid bercoaxial and ber-to-thehome access networks [155]. Cost modeling also showed that deployment of bers with a high SBST in ber-to-the-home access networks can reduce material and labor expenditures by more than 20% [156]. The performance of radio-over-ber links can be characterized by the errorvector magnitude (EVM) [157], which is the quadrature amplitude modulation (or QAM) constellation averaged SNR2. According to the IEEE 802.11a/g WLAN standard, the EVM should be kept below 5.4% rms for highest data rate transmission. The dependence of EVM on the input power is shown in Fig. 19 for RF signal transmission over 20 km of ber I and ber III. At a very low input power the link performance is noise limited. An increase in the input power improves the EVM until SBS starts to deteriorate the signal. With the increased amount of backreected power, the signal quality quickly degrades, and the EVM increases. A clear advantage of the ber with enhanced SBST can be seen by comparing the corresponding EVM curves for ber I and ber III. The higher SBS threshold of ber III allows for low EVM 4 % rms for optical powers up to 17 dBm. For experimental analysis of the RF signal quality of 802.11a/g WLAN packets after transmission over a PON system structure, a setup shown in Fig. 20 was used. A distributed feedback laser signal was fed into an MZM via a polarization controller (PC). The modulator was biased at quadrature and modulated with an 802.11 signal (orthogonal frequency-division multiplexing or OFDM, 64Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 28

Figure 19
8 2.4 GHz: fiber I 2.4 GHz: fiber III 5.8 GHz: fiber I 5.8 GHz: fiber III

EVM, % rms

0 2

input power, dBm

10

12

14

16

18

Measured EVM as a function of the optical input power for two values of the carrier RF and two ber types: ber I and ber III. The ber length is 20 km.

quadrature amplitude modulation) generated by a vector signal generator. Channels at both 2.4 and 5.8 GHz were used to represent 802.11g and 802.11a signals, respectively. The resulting optical signal was amplied with an EDFA and fed into a variable attenuator to control the ber launch power. The maximum launch power was about +17 dBm. After transmission, the signal was attenuated by an additional 16 dB, representing the loss of an optical 32 splitter. The signal then was received by a standard photoreceiver, and the EVM was measured by a vector signal analyzer. As was discussed in the previous section, an additional increase in the SBST can be achieved by using concatenated ber spans. As can be seen from Fig. 21, the EVM can be improved up to 1% rms and the signal power increased by 2 dB by using high-threshold concatenated spans [154].

4.2. SBS in Raman-Pumped Fibers


Raman amplication is widely used in telecommunication systems [158160]. Raman ampliers do not require an additional medium to amplify the signal, because the ber itself serves as an amplifying medium. Raman amplication has a distributed nature. Therefore, the noise gure of the amplier can be made very low to improve the optical SNR compared with EDFA-based systems. This advantage in the noise performance becomes especially important for high-bitrate transmission systems that have an elevated SNR requirement. Apart from the optical SNR improvement, Raman amplication enables a signicant increase in the transmission bandwidth of the system [158].

Figure 20
fiber 20 km PC DFB VSG MZM
OA

Rx VOA 16 dB

VSA

Experimental setup for RF transmission over a single-mode ber. DFB, distributed feedback laser; VSA, vector signal analyzer; VSG, vector signal generator.
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 29

Figure 21
4 3.5 20 km fiber III 15 km fiber III + 5 km fiber I

EVM, %rms

3 2.5 2 1.5 1 2 4 6 8 10 12 14 16 18

input power, dBm

Measured EVM versus optical input power for the concatenated span 15 km ber III +5 km ber I for channel frequency of 5.8 GHz. Results for the uniform, 20 km long ber III are shown for comparison.

Stimulated Raman scattering has many similarities to SBS. However, unlike Brillouin scattering, the Raman effect is due to light interaction with optical rather than acoustic phonons; i.e., molecular vibrations replace the acoustic wave in the scattering process. For typical crystals, the frequency of oscillations of neighboring crystal planes as they move toward each other lies in the infrared, and therefore that branch of dispersion is called optical. The dispersion curve of this optical mode does not originate from the point ( = 0, = 0; see Fig. 2) and thus has a nonzero frequency for = 0 (zero group velocity for the oscillation mode). The dispersion curve of the optical phonon is at near the center of the rst Brillouin zone. Therefore, the frequency of scattered light only weakly depends on the angle between wave vectors of input and scattered light waves. That is why Raman scattering is almost equally efcient in forward and backward directions, and one can use both forward (copropagating with the signal) and backward (counterpropagating) schemes of Raman pumping. For Raman scattering, the frequency offset of the Stokes wave is much larger than in the Brillouin scattering and amounts to 13 THz in glass. Therefore, for a telecom signal at 1550 nm the pump wavelength lies in the range of 1400 1450 nm. The Raman gain can be made quite high 20 dB so that the amplied signals power can approach the SBS threshold. This is especially true for forward Raman pumping. The bandwidth of a Raman amplier is much larger than the Brillouin frequency shift. Therefore, the Raman pump will amplify not only the signal but also the Stokes wave, which then will experience gain from both SBS and stimulated Raman scattering so that a Raman pump will affect the SBST condition. As far as Raman pumping efciency is concerned, several studies indicate that SBS, in turn, causes saturation of the Raman gain [4648,161]. Figure 22 shows how the SBS efciency required in order to achieve the threshold decreases in the presence of Raman pump. The dependence shown in Fig. 22 can be approximated as [49]

th

17
R Leff

41

R , is given by where the effective length due to Raman pumping, Leff

Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001

30

Figure 22
15 10 5 0 0

th

Dimensionless SBS efciency th dened by Eq. (16) corresponding to the SBST f versus the forward Raman pump efciency R [Eq. (43)] in an 80 km span of standard single-mode ber (ber I). The curve is calculated by using Eqs. (41) and (42). b Only forward Raman pump is present, R = 0.

R Leff =

f b RL Rz exp z + R 1 e Rz + R e e 1dz

42

with R being the ber loss coefcient at the wavelength of the Raman pump. The normalized efciencies of forward and backward Raman pumping are
f R =

RP R 0 R

b R =

RP R L R

43

respectively; R is the Raman gain coefcient (in m1 W1), and PR0 and PRL are forward and backward pump powers, respectively. The combined action of SBS and stimulated Raman scattering occurs in dispersion compensating bers (DCFs) which may be Raman-pumped to simultaneously compensate for dispersion and loss in a ber-optic link [162]. DCFs differ from conventional transmission bers in having a shorter length, larger attenuation, and smaller effective area that makes them more sensitive to ber nonlinearities. SBS in passive DCFs was studied experimentally in [112]. The b SBST in forward Pf th and backward Pth pumped DCFs can be approximated as [49] Pf th 12 + R exp RL + L
R Leff

44

Pb th

15
R Leff

45

Measurements of SBST in pumped DCFs [163,164] and backward-pumped small-core highly nonlinear bers [164] indicate a strong decrease in the SBST power with increased Raman gain. A schematic of the experimental setup to measure the SBST of DCFs under various pumping conditions is shown in Fig. 23. Measurements were performed by using a continuous wave single-channel external cavity laser source with a linewidth of 100 kHz at 1550 nm. The signal was amplied before being launched into the DCF. The SBST was measured for unAdvances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 31

Figure 23
forward Raman pump

PC DFB
1550 nm

VOA

circulator

OA
PM

DCF

backward Raman pump

Experimental setup for measurement of SBST in Raman-pumped DCFs. DFB, distributed feedback laser; PM, powermeter; PC, polarization controller.

pumped, forward, and backward-pumped DCFs with two Raman pumps centered at 1440 and 1460 nm. For theoretical calculations, this scheme was approximated by a single pump at 1450 nm. Measurements were performed with Raman pump powers of 190, 280, and 380 mW, which correspond to onoff Raman gains of approximately 10, 15, and 20 dB, respectively. The results of the measurements are shown, together with results of calculations using Eqs. (44) and (45), in Fig. 24. As can be seen from the plot, the threshold power in forward-pumped bers is less than that for backward pumping with the same Raman pump power. This implies that SBS is more detrimental in forward-pumped bers than in backward-pumped bers. Theoretical predictions for the SBST are in good agreement with measurements. Measurements of [164] imply that a decrease in SBST due to the presence of a backward Raman pump depends only on the net Raman gain (the ber loss is offset) and does not depend on the ber type. This trend can be seen from Eqs. (45) and (42). Indeed, for backward pumping the difference in the SBST due to R onoff Raman gain is Ponoff dB = log101 eL / Leff , does not depend on the Brillouin efciency , and for short bers only weakly depends on the ber loss .

4.3. Slow Light and Optical Delay Lines


An interesting eld for the application of SBS that resulted in intensive research in recent years is the generation of slow light, where the group velocity of light

Figure 24
5

SBST, dBm

10

forward pump backward pump 0

Raman pump power, mW

100

200

300

400

SBST in a Raman-pumped DCF. Scattered data show measured values, dashed curves are calculated by using Eqs. (44) and (45). The DCF length is L = 10.5 km.
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 32

propagation in a medium is signicantly lower than its phase velocity [165]. This is achieved through increase in the group refractive index ng 1 by modifying the dispersion of the optical waveguide. The change in the group refractive index can occur because of a narrow spectral resonance of a medium. Narrow resonances due to the SBS process are good candidates for changing the group velocity of a pulse. The pulse delay Tm due to the mth acoustic mode resonance is given by [166,167] Tm =

mP pL
2wm

46

For the Stokes pulse, m is positive (gain) and Tm 0; i.e., the pulse is delayed relative to its propagation time in a nonresonant, passive medium. As was mentioned above, m 0 for the anti-Stokes pulse (i.e., the pulse is attenuated), and the fast-light regime is realized. The advantage of SBS versus other resonant techniques such as electromagnetically induced transparency or coherent population oscillation is the opportunity to control the pulse delay all-optically by varying the pump power [see Eq. (46)]. However, the Stokes pulse has to be i) centered precisely at the resonance and ii) have a bandwidth smaller than wm. The use of SBS in optical bers is especially attractive since it easily leads to the application of slow light in standard telecom operating windows, moderate pump powers, use of optical bers as a transmission medium, easy connection to standard telecom equipment, and operation at room temperature [167169]. Different types of bers were considered for generation of slow light. A comparative analysis of slica, bismuth oxide, tellurite, and As2Se2 chalcogenide bers can be found in [166]. More details on slow and fast light in optical bers can be found in a recent review [167]. Slow light can be used for multiple applications in optical communications and signal processing, optical buffering and storage, jitter compensation, synchronization of data, microwave photonics (e.g., phased array antennas), etc. The demonstrations of slow-light generation using SBS in optical bers in 2005 [168,170] triggered an intensive wave of research in this eld. The ability to obtain optical delays that can be controlled by the pump signal is very attractive for the design of future optical systems. It was shown that 2 ns pulses (wavelength 1.55 m) can be stored for up to 12 ns via SBS in a highly nonlinear ber. The stored pulses were then retrieved by a short intense read pulse having the same frequency as the pump pulse [171]. First demonstrations of SBS-assisted slow light used a single pump wavelength counterpropagating with the signal wavelength in an optical ber. If both wavelengths are separated by the SBS acoustic wave frequency, i.e., Brillouin frequency shift (11 GHz in standard optical ber), a change in refractive index is obtained and the signal is slowed down. A side effect is the SBS gain that the signal experiences, leading to an increase in the signal strength at the same time. Other effects like potential signal distortion through increased dispersion need to be well controlled in order to avoid signal distortion beyond practical limits [172]. Since the goal of slow-light delay is to control the timing of optical signals, the data signal needs to be detectable without major degradation, and error-free operation needs to be maintained. An issue for practical applications is the very limited SBS gain bandwidth of only approximately 25 MHz in standard optical bers. With such a small operating bandwidth, high-data-rate signals of 10 Gbit/ s and more cannot be effectively supAdvances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 33

ported. Initial results showed that pulses of 100 ns could be slowed down by up to 32 ns [170], and 63 ns pulses by up to 25 ns [168] without signicant pulse distortion, demonstrating the potential of slow light by SBS in optical bers. In order to overcome the limited operating bandwidth of single-pump systems, several investigations focused on the demonstration of multipump systems for slow-light generation. Since the total SBS power adds linearly for each pump wavelength, a superposition of the SBS gain spectra for multiple pump signals at slightly different wavelengths will lead to a broadening of the resulting gain bandwidth. Multiple techniques have been devised to design multipump generators [173,174]. Comb lasers are an interesting way to generate multipeak pumps because of the ability to provide broad and uniform pump signal generation [175]. A signicant breakthrough for broadband slow-light generation occurred in 2006 when a single pump laser was directly modulated with a noise signal, leading to a very uniform SBS gain spectra of 325 MHz bandwidth [176]. Now, pulses of only 2.7 ns duration could be delayed by more than a pulse width without signicant distortion. This technique was later improved and used to generate a 12 GHz wide SBS spectrum that allowed even 10 Gbit/ s signals to be delayed [177]. This experiment also explored the limits of single-pump SBS slow-light generation due to the limited Brillouin frequency shift of 11 GHz. When the SBS gain becomes broader than the Brillouin frequency shift, a boundary for slow-light generation by this technique is reached. Further investigations led to the extension of the gain bandwidth by separating the pumps at twice the Brillouin gain shift. This led to an effective gain bandwidth approaching 25 GHz [178]. Reaching bandwidths that allow 40 Gbits/ s and eventually 100 Gbits/ s data transmission delay will be important for future systems, and bandwidths of 50 GHz will be needed. Overall, there has been signicant research to demonstrate the slow-light effect on real data transmission and proof that error-free transmission with signal delay can be obtained [174,179]. While most investigations focus on NRZ signals, differential phase-shift keying (DPSK) with its constant envelope behavior was also investigated [180]. A delay of 42 ps on a 10.7 Gbit/ s NRZ-DPSK signal was demonstrated error-free with a power penalty 10 dB. In comparing NRZ-DPSK and return-to-zero (RZ)-DPSK, it is shown that RZ-DPSK signicantly outperforms NRZ-DPSK by 2 dB in power penalty at the same pump power.

4.4. SBS-Based Fiber-Optic Sensors


An SBS-based distributed sensor relies on the temperature and strain dependence of the Brillouin frequency shift [181187]. A schematic diagram of the SBS-based sensing setup is shown in Fig. 25. A pulsed source is required in order to obtain positional information about stress or temperature variation by means of time-delay analysis. The spatial resolution of a sensor is determined by the pulse length and is usually limited to 1 m because of the nite response time of the scattering process [188]. It is worth mentioning that the longitudinal stress is difcult to measure accurately with other techniques. A longitudinal strain of 2 103 results in about a 100 MHz shift of the BGS maximum that is easily detectable [188]. However, a stable, narrowband tunable laser is required for reliable operation of the scheme. For example, a 50 m spatial resolution with a 104 strain sensitivity was achieved with a Nd:YAG ring laser (1 mW
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 34

Figure 25
pump,
p

circulator

Stokes,

p- B

pulsed laser
SA

cw laser

sensing fiber

Schematic diagram of the SBS-based sensing setup. SA is the spectrum analyzer that detects pump and Stokes signals.

power, 5 kHz spectral line width) [189]. As for application areas, a distributed Brillouin sensor was used in civil engineering to identify and localize buckling [190] and wall-thinning defects [191] in steel pipes. Another interesting example is health monitoring of dams. A distributed temperature sensor has been used to monitor concrete setting temperatures of a large dam in Switzerland [183]. The ber cable was installed during the concrete pouring. A spatial resolution of 1 m and a temperature sensitivity of 1 C was demonstrated [183]. From the temperature variation during the setting chemical process one could determine the concrete density and identify microcracks. More details on Brillouin ber sensors can be found in a recent paper [192].

4.5. High-Power Fiber Lasers


A detrimental effect of SBS in optical bers can be observed in high-power ber laser and amplier designs. Because of the natural goal of generating high power levels [193], the SBS threshold is easily reached in many designs, and techniques for SBS suppression are often required. While ber lasers have unique features such as excellent beam quality, high efciency, very high output powers of up to hundreds of watts, choice of lasing frequency (within limits), and ease of thermal management, SBS is often the main limitation on achievable output power for these devices. It therefore is of high interest to increase the SBS threshold in ber laser designs [194]. A widely used technique for SBS suppression is the modulation of the pump laser source [68,144]. This leads to laser linewidth broadening and therefore to a signicant rise in SBS threshold by 10 dB or more. Multiple techniques can be considered: direct modulation of a semiconductor laser with a low-frequency signal, external phase modulation, or a combination of linewidth enhancing techniques. Besides pump linewidth broadening, the transmissiongain medium, i.e., the ber, can be inuenced to produce an increased SBS threshold. Higher-order modes can be used to essentially increase the effective area of the ber, easily raising the SBS threshold signicantly. In order to achieve singlemode designs, mode converters then need to be added [195]. The use of largemode-area bers also has been proposed [196] but is limited by operation in the fundamental ber mode, which puts design restrictions on this approach. The application of thermal gradients to the ber was also extensively investigated. Either the absorption of the pump signal itself [197,198] or some form of external gradient [199] was considered to increase the SBST. However, thermal management through either method is limited because too much heat is generated in the ber core or too much external heat is applied. Nevertheless, several hundreds watts of power could be demonstrated by using such techniques.
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 35

Another technique includes the inuence of the acousto-optic mode eld overlap in the ber medium by design. It has been demonstrated that the Ge doping of the core can be manipulated in such a way that the optical and acoustic elds in the ber have signicantly reduced overlap compared with standard stepindex bers and therefore raise the SBS threshold [22]. However, this technique is difcult to implement in large-mode-area bers with low numerical aperture (NA) where the dopant concentration is relatively small [194]. A technique to overcome this limitation was investigated in [200], where a combination of Ge and Al doping was proposed. This way, the spatial separation between the acoustic and optical elds was increased and the SBS threshold could be raised by 6 dB.A similar design led to an 11 dB SBS increase [201]. Using such designs, very high output powers of up to 500 W with narrow linewidth signals could be achieved [202].

5. Other Fiber-Optic Issues


As was shown above, even if a ber is single-mode with respect to optical modes, it can be multimode with respect to acoustic modes, and it is the overlap between the fundamental optical and the acoustic modes that determines the SBS efciency. The situation becomes even more involved if either the number of optical modes increases or the shape of the optical and acoustic modes drastically changes, for example, if modes become more tightly localized. The former situation takes place in large-core multimode bers (MMFs), while the latter is typical for the small-core PCFs. Below, we briey review SBS peculiarities in these ber types.

5.1. SBS in Multimode Fibers


A rigorous treatment of multiple modes (both optical and acoustic) in a largearea glass cylinder with a graded-index prole seems to be very complicated. However, good agreement with measured data for the SBST is achieved if the nominal core area of a MMF, r2 c , where rc is the radius of the ber core, is re2 placed with some effective area reff with the effective mode radius reff rc [203,204]. Then, the standard threshold equation (30) was used to obtain the SBST value for MMFs. For a multimode graded-index ber with a core diameter of 50 m, cladding diameter of 125 m, and length of 4.4 km, the effective mode diameter was found to be 20.6 m, which explained the experimentally determined SBST of 100 mW. Using the plane wave approach resulted in 4 times overestimation of the threshold power [203]. It was also found that the gain coefcient for lower-order modes strongly exceeds (by a factor of 16) the gain of the higherorder modes [205]. An analytical estimate for the SBST increase in MMFs compared with a singlemode ber was derived in [206]. A step-index prole of the MMF was assumed. The ratio of the SBST power in a MMF to that in a single-mode ber Pth can be written as PMMF th Pth where
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 36

Y = arctan Y ,

47

Y=

vAncore nclad pw

48

w is the FWHM of the BGS, and ncore and nclad are the core and cladding refractive indices, respectively. It was assumed that the mode with the maximum gain induces the SBS process. For small-numerical-aperture bers, Eq. (48) can be expressed through the NA parameter as vANA2 2ncorepw

Y=

49

For most commercial MMFs, the numerical aperture is NA 0.2. In this case, the value of PMMF / Pth estimated from Eq. (47) amounts to 1.8 for the pump wavelength th p = 1 m [206]. A further increase in the SBST accuracy in MMFs can be obtained if one accounts for the inhomogeneous broadening of the Brillouin gain [207]. In a recent work [208], an approximation for the BGS shape was proposed based on numerical evaluation of the overlap between optical and acoustic modes.

5.2. SBS in Microstructured Fibers


Microstructured bersalso called holey bers or PCFsdemonstrate numerous novel effects of light propagation such as endlessly single-mode operation or bandgap guiding [209]. Periodicity of the refractive index variation in the bers cross section and small core of a PCF strongly increase connement of light and sound and may drastically change the dispersion relation of both optical and acoustic modes. This changes the acoustic guiding properties of the ber compared with standard single-mode ber. As a result, a BGS may contain resonances corresponding to not only longitudinal (as in conventional single-mode bers) acoustic modes but also to shear modes or coupled states between longitudinal and shear modes. For small-core PCFs (core diameter is less than the vacuum wavelength), a multi-peaked BGS with Brillouin shifts 10 GHz was observed [210]. One can observe a relatively strong forward scattering Brillouin process, similar in nature to Raman scattering even though acoustic phonons are involved [211]. PCFs typically behave as acoustic antiguides because, unlike conventional bers with Ge-doped cores, the holey cladding has a signicantly lower acoustic index compared with the core. Among other interesting features of Brillouin scattering in PCFs, one can mention both increased (up to 100 relative to fused silica [212] in a PCF based on sulde glass) and reduced [105] Brillouin gain, strong polarization dependence (3 dB SBST difference between two polarizations) of the Brillouin gain [213,214], or reduced guided acoustic wave Brillouin scattering noise [215,216], which can facilitate experiments operating at the quantum noise limit. It was found that with decreasing core size, the SBST increases faster than in a conventional single-mode ber [213]. This is attributed to the diffraction of the acoustic wave out of the small PCF core and the acoustic antiguiding of PCFs. Acoustic resonances in nonlinear PCFs were shown to be controllable coherently. Depending on the time delay between two pulses launched into the PCF, the acoustic resonance in the forward Brillouin spectrum could be either coherently reinforced or suppressed [217]. To observe the effect, the pulse delay should be well within the lifetime of the acoustic
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 37

resonance 11 ns. Small-core PCFs also show promise in enhancing the slowlight delay. For example, up to one-half pulse width delay was achieved in a 50 m PCF by using a single pump [218].

6. Conclusions
We have reviewed stimulated Brillouin scattering (SBS) in optical bers. Brillouin gain of a ber depends on the index prole through the overlap integral between the acoustic mode prole and the squared optical mode prole. We have also introduced the acousto-optic effective area, which is inversely proportional to the Brillouin gain coefcient. Analytical results for the noise-initiated SBS threshold obtained from different approximations were analyzed. We also reviewed several approaches to control Brillouin gain in the ber. An attractive method is an index prole design that can be used to both suppress and enhance Brillouin gain. We also discussed BFAs and distributed sensors. Several recently featured SBS-based or SBS-related applications were reviewed, such as radioover-ber technology, slow light and optical delay lines, and high-power lasers. We have also discussed SBS in large-core multimode and PCFs.

Appendix A: Derivation of Evolutional Equations for Signal and Stokes Waves


We start our analysis with the material density equation (acoustic wave) and Maxwells equations (optical waves) and present intermediate results to make the derivation easy to follow. First, we look for a solution of the equation for the material density [7072,90]

2 t
2

2 vA r 2 =

e
2

2E 2 ,

A.1

where kg/ m3 is the material density uctuation around its mean value 0, 2 = 11 / 0 m2 / s is the damping factor, vA r = Yr / r m2 / s2 is the squared longitudinal sound velocity that depends on the transverse radial coordinate r due to silica doping with GeO2, Yr [Pa] is theYoungs modulus, e = n40p12 [F/m] is the electrostriction constant, n is the glass refraction index, 11 [Pa s] and p12 [dimensionless] are the respective components of the viscosity and electrostriction tensors, and 0 is the vacuum permittivity. The electric eld E on the right-hand side of Eq. (A.1) is represented as a superposition of forward- and backward-propagating electromagnetic waves: 1 Er, z, t = frA1z, tei1t1zu1 + A2z, tei2t+2zu2 + c.c., 2

A.2

where fr is the dimensionless fundamental optical mode prole; Ajz , t, j = 1 , 2, are the slowly varying envelopes of the optical eld; uj are unit vectors in the forward and backward direction of propagation; j and j are, respectively, frequencies and propagation constants of optical waves; and c.c. denotes the complex conjugate. Then the right-hand side of Eq. (A.1) takes the form
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 38

1 2 idtsz 2E 2 f 2 r A 1 z , t A * z, ts e + c.c., 2 2

A.3

where d 1 2 and s 1 + 2 21neff / c 4n / with neff, , and c being the effective refractive index of the optical mode, input signal wavelength, and speed of light, respectively.

A.1. Stationary Solution for the Density Variation


We look for the solution of Eq. (A.1) in the form mz, tmr, eitqz + c.c., 2
m=1

z, t, r, =

A.4

where M is the number of acoustic modes of the optical ber, is the acoustic frequency, and q is the corresponding wave vector of the density variation. The respective derivatives of the material density are

t 2 t
2

2 2 2
1
M

m=1

m meitqz , + i

m 2

m=1

+ 2i

m meitqz , 2

2 z2

m 2

m=1

z2

2iq

m meitqz , q 2

where +c.c. terms are assumed on the right-hand side. With all this, the left-hand 1 M itqz side of Eq. (A.1) can be written as 2 m S, where =1e S = m

m 2

2 vA

t2

+ 2i

m m q 2

m 3

m 2

2iq

z 2 t

i m
2 m 2 + i + vA m ,

A.5
2 2 2 2 2 2 2 2 2 = 2 + / z , and = / r + 1 / r / r + 1 / r / is the transverse Laplacian operator in cylindrical coordinates.

Because of slow variation of the acoustic wave in both the propagation direction (z coordinate) and time, we can discard terms with second- and higher-order de2 rivatives with respect to z and t. Further, the ratio / vA 103 is small (this is just the ratio of the Brillouin gain line width and the Brillouin shift frequency, which 2 will be dened below) in Eq. (A.5), and we can approximate i + vA with its real 2 part vA in terms that contain derivatives, i.e.,
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 39

S = 2im 2 m
2 m vA


2 m + 2
2 vA

2 + 2iqvA m

m + iq2m A.6

q2 m .

Since the attenuation coefcient of hypersonic phonons is very high acoust m / z. If we further = q2 / vA 104 m1 [72], we can also drop the term with assume the steady-state condition, Eqs. (A.1), (A.3), and (A.6) result in 1

2 m=1 =

2 mq2 mv A eitqz im 2 m +


2
2 vA

q2 m

e
4

f 2 r A 1A * 2e idtsz . 2 s

A.7

Similar to a treatment of Kerr nonlinearity in optical bers [79,75], we now assume that the presence of optical elds [i.e., the right-hand side of (A.7)] and attenuation of the acoustic wave do not strongly affect the modal structure of the acoustic wave. In other words, the solution of the modal equation 2 m r , +

2 m 2 vA r

q2 mr, = 0,

A.8

for the unperturbed mode can be used to solve Eq. (A.7). In Eq. (A.8), m is the eigenfrequency of the mth solution of the modal acoustic equation that for a given wave vector q satises Eq. (A.8). In what follows, we consider acoustic modes without axial variation / = 0, since only those modes interact efciently with the axially symmetric optical mode fr. On substitution of Eq. (A.8) into Eq. (A.7), we obtain
2 mq2 m m 2 m = eitqzim 2 m=1

e
4

f 2 r A 1A * 2e idtsz . 2 s A.9

As a nal part of nding a stationary solution to the acousto-optic equation (A.1), we multiply both sides of Eq. (A.9) by mr and integrate over the transverse plane. This procedure is required whenever separation of transverse and longitudinal variables [see Eq. (A.4)] is employed and is also used in treating the ber nonlinearity [75,79]. It was noted in [90] that modes of Eq. (A.8) are not orthogonal. We found, however, that for the rst several modes of Eq. (A.8) 2 8 k. To eliminate the sum on the left-hand 0 mrkrrdr / 0 mrrdr 10 , m side of Eq. (A.9), we can multiply both sides of Eq. (A.9) by k and integrate them over the transverse area. Finally, the wave vector and phase matching conditions for the Brillouin process [72] require that q = s and = d, and we obtain from Eq. (A.9) m z , t =

eq 2A 1 z , t A * z, t mrf 2r 2
2 2m 2 + iq2 2 m r

A.10

where we have denoted integrated quantities by angle brackets, wr = 2 0 wrrdr. Since each acoustic mode interacts independently with the opAdvances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 40

tical eld, from Eqs. (A.4) and (A.10) the full material density variation is

z, t, r =

4 m=1

m r

2 z, t mrf 2r eq A1z, tA* 2 2 m r 2 m 2 + iq2

eitqz + c.c. A.11

The eigenfrequency of the acoustic mode m corresponds to the full resonance occurring both for frequencies (as determined by the energy conservation) and for wave vectors (as determined by the momentum conservation). Strictly speaking, it is a solution of the system of equations q=

1neff1
c 1 r r

1 neff1 c 2

A.12

2 r2

2 r vA

q2 = 0,

A.13

where in Eq. (A.12) we have already used the frequency matching condition 2 = 1 together with a very reasonable approximation neff2 neff1 because the two optical frequencies are very close. The effective index of the optical wave neff1 is found from the optical modal equation, which is given below. Thus, for a xed laser wavelength 1 one can calculate m as the intersection of two curves q obtained from Eqs. (A.12) and (A.13). However, as can be seen from Eq. (A.12), q is a very weak function. Indeed, because 1 q neff1 c 21 21neff1 c 2 1n c ,

and the obtained value of q can be directly substituted into Eq. (A.13) to nd m. The solution of Eq. (A.13) also gives the modal prole r which, strictly speaking, will slightly vary with . However, since changes over a narrow span of frequencies 100 MHz, we can take some generic acoustic mode prole calculated by using q = 21n / c for further calculations.

A.2. Equations for Evolution of Optical Power


The next step is to derive equations for the evolution of the optical power. With several standard approximations similarly used for the derivation of the nonlinear optical propagation equation [72,75,79], Maxwells equations can be reduced to the following equation for the electric eld: E
2

L 2E c2 t2

2PNL t2

=0

A.14

where L and 0 are the linear part of the dielectric constant and the magnetic permeability, respectively, while the nonlinear polarization induced by the acousto-optic interaction is [72,90] PNL = so Eq. (A.14) can be rewritten as
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 41

e
0

E = 0NLE;

A.15

E=
2

tot 2E c2 t2

A.16

where tot = L + NL = n2r inc / + NL with being the loss coefcient of the optical ber. In deriving Eq. (A.16) the following were assumed: the nonlinear polarization PNL can be treated as a small perturbation, i.e., 2NLE / t2 NL2E / t2; the optical eld maintains its polarization so that the scalar approximation is valid; the optical eld is quasi-monochromatic; the nonlinear response is instantaneous; and the weak guiding [219], or the small index gradient [78], condition / 1 is satised. In the linear part of the total dielectric constant tot, we have neglected both differences in the material properties at close frequencies = 1 2 and the spatial dependence of n in the linear attenuation term because IL RL. Assuming independent interaction between the optical mode and each acoustic mode, we can write the nonlinear part of the total dielectric constant as NL = mUA1A* e i12ti1+2z + U *A * A ei12t+i1+2z , A.17 2 1 2 where U=
2 2 eq

mf 2

2 2 m 4 0 0 m 2 + i

A.18

= q2. Using the slowly varying envelope approximation We have designated [75,79], we obtain relations for the derivatives of the total electric eld equation (A.2), E=
2

2 E

2E

i1t1z 2 + 1A 1 e

z2

2 f r 2

A 1e i1t1z + A 2e i2t+2z +

fr 2

fr 2

2i2

A2 z

i2t+2z 2 + c.c, 2A 2 e

2i1

A1 z

2E t2

fr 2

i1t1z i2t+2z A 1 2 + A 2 2 + c.c, 1e 2e

and substitute them into Eq. (A.16) and group the resulting terms by the exponential factors ei1t1z and ei2t+2z to obtain two equations for forward and backward propagating optical waves: A 1 2 f r f r 2 i 1

A1 z

2 + 2 1A 1 = n r i

nc fr

c2

A 1 2 1 A.19

U m r f r c2

2 2 2A 1 A 2 ,

A 2 2 f r + f r 2 i 2

A2 z

2 2 2A 2 = n r i

nc fr

c2

A 2 2 2 A.20

U * m r f r c2

2 2 1 A 1 A 2 .

Neglecting the dependence of the optical modal prole on interaction with acoustic waves, we can substitute the optical modal equation
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 42

2 f r

2 2 j n r

c2

2 j fr = 0

A.21

into Eqs. (A.19) and (A.20), multiply both sides of each equation by fr, and integrate the results over the transverse cross section to nally obtain the coupled ODEs in the form dA1 dz dA2 dz where =

A 1 i A 1 A 2 2 ,

A.22

A 2 + i * A 1 2A 2 ,

A.23

U mf 2 2cn f 2

3 n 9 0p 2 12

mf 2 2

2 2 m f 2 2 0c 3 m 2 + i

A.24

In obtaining Eqs. (A.22) and (A.23) we have used 1 2 n / c, and in Eq. (A.24) we have used Eq. (A.18) together with the denition of e and q 2n / c. Equations (A.22) and (A.23) are written for the evolution of the optical eld, which has a dimension of volts per meter. They can be rewritten for the optical power in watts, which is the optical intensity Ijr = 0cnAj2f 2r / 2 integrated over the transverse cross section, i.e., P j = I j = 0cn 2 Aj2f 2r . A.25

Substitution of Eq. (A.25) into Eqs. (A.22) and (A.23) results in equations for the spatial evolution of the forward (j = 1, upper sign) and backward (j = 2, lower sign) optical power dPj dz = Pj 2 i * 0cnf2 P 1P 2 . A.26

In the text, the guided powers of the pump and Stokes waves are denoted P1 = Pp and P2 = PS, respectively. We can nally write the evolutional equations (A.26) as dPj dz where
ao Am =

= Pj

gm
ao Am

L P 1P 2 ,

A.27

f 2r mrf 2r

2 2 m r

A.28

is the acousto-optic effective area that determines the strength of the acoustooptic interaction in optical bers with different proles of core doping. This denition does not require the normalization of acoustic and optical mode proAdvances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 43

les, since dimensionless functions fr and mr appear to the same power in the numerator and denominator of Eq. (A.28). From Eq. (A.26), the peak Brillouin gain corresponding to the mth acoustic mode gm is given by
3 2 n 8p 2 12

gm =

c4 m 0

A.29

which agrees with the value for the peak Brillouin gain in equations for intensities obtained for bulk media (see, e.g., [72], p. 420 with n4p12 = enote the cgs unit system used thereand m = 2nvA / c). Finally, the Lorentzian prole of the Brillouin gain comes from Eqs. (A.24) and (A.26)

L =

/22 /22 m 2 +

A.30

where

wm =

8 n 2 2 p

A.31

is the FWHM of the BGS corresponding to the mth acoustic mode and m 2

m =

2nvA p

A.32

is the Brillouin frequency shift. The peak Brillouin gain can be introduced in terms of wm, m, and p as 2 n 7p 2 12 c2 p 0v Aw m 4 n 8p 2 12 c3 p 0 mw m

gm =

A.33

The same expression was obtained when the guiding nature of acoustic waves was not accounted for, i.e., by using the plane wave approximation for acoustic waves (see, e.g., [20,30,31,117], [79] p. 207, and [75] p. 357, with an extra factor 2 needed in the numerator). Hence, the peak value of the SBS gain is the same in bulk and waveguide geometries. But the important difference in the effective Brillouin gain coefcient in the equation for the optical power evolution, Eq. (A.27), is due to the modal overlap factor determined by the acousto-optic effective area, Eq. (A.28).
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 44

Appendix B: Approximate Analytical Solution for the SBST in the Short and Long Fiber Approximation
We are looking for an approximate analytical solution to the equation q le q = a , B.1

where short- and long-ber approximations correspond to the cases l = 3 / 2 and l = 5 / 2, respectively [cf. Eqs. (26) and (30)]. To solve Eq. (B.1), we rewrite it as q l ln q = , B.2

where = ln a, and look for a solution in the form q 1 + , where 1. Substitution of this expression into Eq. (B.2), followed by series expansion of the log function ln1 + , yields

l ln l

B.3

which results in a solution of Eq. (B.1) in the form q 1+

l ln l

B.4

Acknowledgments
The authors gratefully acknowledge collaboration with Frank Annunziata, Scott Bickham, Aleksandra Boskovic, Sergey Darmanyan, Allen Dixon, John Downie, Keith Emig, Alan Evans, Tom Hanson, Jason Hurley, Shiva Kumar, Ming-Jun Li, Claudio Mazzali, Manjusha Mehendale, Raj Mishra, Steven Rosenblum, Sergey Ten, Sergio Tsuda, Michael Vasilyev, Rich Vodhanel, Bill Wood, and Andy Woodn.

References and Notes


1. L. Brillouin, Diffusion de la lumire par un corps transparent homogne, Ann. Phys. 17, 88 (1922). 2. I. L. Fabelinskii, The discovery of combination scattering of light in Russia and India, Phys. Usp. 46, 11051112 (2003). 3. B. R. Masters, C. V. Raman and the Raman effect, Opt. Photonics News 20(3), 4145 (2009). 4. V. Sundar and R. E. Newnham, Electrostriction, in The Electrical Engineering Handbook, 2nd ed., R. C. Dorf, ed. (CRC Press, 1997), pp. 1193 1200. 5. R. Y. Chiao, C. H. Townes, and B. P. Stoicheff, Stimulated Brillouin scattering and coherent generation of intense supersonic waves, Phys. Rev. Lett. 12, 592595 (1964). 6. E. L. Buckland and R. W. Boyd, Electrostrictive contribution to the
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 45

7. 8.

9.

10.

11.

12.

13.

14.

15.

16.

17.

18.

19.

20.

21.

22.

23.

intensity-dependent refractive index of optical bers, Opt. Lett. 21, 1117 1119 (1996). E. L. Buckland, Mode-prole dependence of the electrostrictive response in bers, Opt. Lett. 24, 872874 (1999). E. M. Dianov, M. E. Sukharev, and A. S. Biryukov, Electrostrictive response in single-mode ring-index-prole bers, Opt. Lett. 25, 390392 (2000). E. M. Dianov, M. E. Sukharev, and A. S. Biryukov, Electrostrictive response in single-mode ring-index prole bers: errata, Opt. Lett. 25, 987 (2000). A. S. Biryukov, M. E. Sukharev, and E. M. Dianov, Excitation of sound waves upon propagation of laser pulses in optical bres, Quantum Electron. 32, 765775 (2002). P. D. Townsend, A. J. Poustie, P. J. Hardman, and K. J. Blow, Measurement of the refractive-index modulation generated by electrostriction-induced acoustic waves in optical bers, Opt. Lett. 21, 333335 (1996). A. Fellegara, A. Melloni, and M. Martinelli, Measurement of the frequency response induced by electrostriction in optical bers, Opt. Lett. 22, 16151617 (1997). E. L. Buckland and R. W. Boyd, Measurement of the frequency response of the electrostrictive nonlinearity in optical bers, Opt. Lett. 22, 676678 (1997). A. Melloni, M. Frasca, A. Garavaglia, A. Tonini, and M. Martinelli, Direct measurement of electrostriction in optical bers, Opt. Lett. 23, 691693 (1998). A. S. Biryukov, S. V. Erokhin, S. V. Kushchenko, and E. M. Dianov, Electrostriction temporal shift of laser pulses in optical bres, Quantum Electron. 34, 10471053 (2004). N. Shibata, Y. Azuma, T. Horiguchi, and M. Tateda, Identication of longitudinal acoustic modes guided in the core region of a single-mode optical ber by Brillouin gain spectra measurements, Opt. Lett. 13, 595597 (1988). N. Shibata, K. Okamoto, and Y. Azuma, Longitudinal acoustic modes and Brillouin-gain spectra for GeO2-doped-core single-mode bers, J. Opt. Soc. Am. B 6, 11671174 (1989). A. Yeniay, J. M. Delavaux, and J. Toulouse, Spontaneous and stimulated Brillouin scattering gain spectra in optical bers, J. Lightwave Technol. 20, 14251432 (2002). Y. Koyamada, S. Sato, S. Nakamura, H. Sotobayashi, and W. Chujo, Simulating and designing Brillouin gain spectrum in single-mode bers, J. Lightwave Technol. 22, 631639 (2004). M. Nikls, L. Thvenaz, and P. A. Robert, Brillouin gain spectrum characterization in single-mode optical bers, J. Lightwave Technol. 15, 1842 1851 (1997). J. Yu, Y. Park, K. Oh, and I. Kwon, Brillouin frequency shifts in silica optical ber with the double cladding structure, Opt. Express 10, 9961002 (2002). A. Kobyakov, S. Kumar, D. Chowdhury, A. B. Rufn, M. Sauer, S. R. Bickham, and R. Mishra, Design concept for optical bers with enhanced SBS threshold, Opt. Express 13, 53385346 (2005). A. H. McCurdy, Modeling of stimulated Brillouin scattering in optical -

Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001

46

24.

25.

26.

27.

28. 29.

30. 31.

32.

33. 34. 35.

36. 37.

38.

39.

40.

41.

42.

bers with arbitrary radial index prole, J. Lightwave Technol. 23, 3509 3516 (2005). S. Afshar, V. P. Kalosha, X. Bao, and L. Chen, Enhancement of stimulated Brillouin scattering of higher-order acoustic modes in single-mode optical ber, Opt. Lett. 30, 26852687 (2005). A. B. Rufn, M.-J. Li, X. Chen, A. Kobyakov, and F. Annunziata, Brillouin gain analysis for bers with different refractive indices, Opt. Lett. 30, 31233125 (2005). V. Lanticq, S. Jiang, R. Gabet, Y. Jaoun, F. Taillade, G. Moreau, and G. P. Agrawal, Self-referenced and single-ended method to measure Brillouin gain in monomode optical bers, Opt. Lett. 34, 10181020 (2009). L. Tartara, C. Codemard, J.-N. Maran, R. Cherif, and M. Zghal, Full modal analysis of the Brillouin gain spectrum of an optical ber, Opt. Commun. 282, 24312436 (2009). B. G. Ward and J. B. Spring, Brillouin gain in optical bers with inhomogeneous acoustic velocity, Proc. SPIE 7195, 71951J (2009). P. D. Dragic, Estimating the effect of Ge doping on the acoustic damping coefcient via a highly Ge-doped MCVD silica ber, J. Opt. Soc. Am. B 26, 16141620 (2009). K. Ogusu, H. Li, and M. Kitao, Brillouin-gain coefcients of chalcogenide glasses, J. Opt. Soc. Am. B 21, 13021304 (2004). K. S. Abedin, Observation of strong stimulated Brillouin scattering in single-mode As2Se3 chalcogenide ber, Opt. Express 13, 1026610271 (2005). M. O. van Deventer and A. J. Boot, Polarization properties of stimulated Brillouin scattering in single-mode bers, J. Lightwave Technol. 12, 585 590 (1994). H. E. Engan, Analysis of polarization-mode coupling by acoustic torsional waves in optical bers, J. Opt. Soc. Am. A 13, 112118 (1996). Y. Imai and M. Yoshida, Polarization characteristics of ber-optic SBS phase conjugation, Opt. Fiber Technol. 6, 4248 (2000). P. Narum and R. W. Boyd, Nonfrequency-shifted phase conjugation by Brillouin-enhanced four-wave mixing, IEEE J. Quantum Electron. 23, 12111216 (1987). A. M. Scott and K. D. Ridley, A review of Brillouin-enhanced four-wave mixing, IEEE J. Quantum Electron. 25, 438459 (1989). K. Inoue, T. Hasegawa, and H. Toba, Inuence of stimulated Brillouin scattering and optimum length in ber four-wave mixing wavelength conversion, IEEE Photon. Technol. Lett. 7, 327329 (1995). K. Ogusu, Interplay between cascaded stimulated Brillouin scattering and four-wave mixing in a ber FabryPerot resonator, J. Opt. Soc. Am. B 20, 685694 (2003). J. D. Downie and J. Hurley, Experimental study of SBS mitigation and transmission improvement from cross-phase modulation in 10.7 Gb/ s unrepeatered systems, Opt. Express 15, 95279534 (2007). P. Narum, A. L. Gaeta, M. D. Skeldon, and R. W. Boyd, Instabilities of laser beams counterpropagating through a Brillouin-active medium, J. Opt. Soc. Am. B 5, 623628 (1988). D. E. Watkins, A. M. Scott, and K. D. Ridley, Determination of the threshold for instability in four-wave mixing mediated by Brillouin scattering, IEEE J. Quantum Electron. 26, 21302137 (1990). A. A. Fotiadi, G. Ravet, P. Mgret, and M. Blondel, Multi-cascaded SBS in
47

Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001

43.

44.

45.

46.

47.

48.

49.

50.

51. 52.

53.

54.

55. 56. 57.

58.

59.

an optical ber supported by Rayleigh backscattering, Proc. SPIE 5480, 7181 (2003). C. N. Pannell, P. St. J. Russell, and T. P. Newson, Stimulated Brillouin scattering in optical bers: the effect of optical amplication, J. Opt. Soc. Am. B 10, 684690 (1993). S. L. Zhang and J. J. OReilly, Effect of stimulated Brillouin scattering on distributed erbium-doped ber amplier, IEEE Photon. Technol. Lett. 5, 537539 (1993). M. F. dos Santos Ferreira, Impact of stimulated Brillouin scattering in optical bers with distributed gain, J. Lightwave Technol. 13, 16921697 (1995). B. Foley, M. L. Dakss, R. W. Davies, and P. Melman, Gain saturation in ber Raman ampliers due to stimulated Brillouin scattering, J. Lightwave Technol. 7, 20242032 (1989). M. F. dos Santos Ferreira, J. F. Rocha, and J. L. Pinto, Impact of stimulated Brillouin scattering on bre Raman ampliers, Electron. Lett. 27, 1576 1577 (1991). S. Hamidi, D. Simeonidou, A. S. Siddiqui, and T. Chaleon, Effect of pump laser mode structure on the gain of forward pumped Raman bre amplier in the presence of stimulated Brillouin scattering, Electron. Lett. 28, 1768 1770 (1992). A. Kobyakov, M. Mehendale, M. Vasilyev, S. Tsuda, and A. F. Evans, Stimulated Brillouin scattering in Raman-pumped bers: a theoretical approach, J. Lightwave Technol. 20, 16351643 (2002). A. P. Kng, A. Agarwal, D. F. Grosz, S. Banerjee, and D. N. Maywar, Analytical solution of transmission performance improvement in ber spans with forward Raman gain and its application to repeaterless systems, J. Lightwave Technol. 23, 11821188 (2005). G. Valley, A review of stimulated Brillouin scattering excited with a broad-band pump laser, J. Lightwave Technol. 22, 704712 (1986). P. Narum, M. Skeldon, and R. W. Boyd, Effect of laser mode structure on stimulated Brillouin scattering, J. Lightwave Technol. 22, 21612167 (1986). K. Ogusu, Effect of stimulated Brillouin scattering on nonlinear pulse propagation in ber Bragg gratings, J. Opt. Soc. Am. B 17, 769774 (2000). H. Lee and G. P. Agrawal, Suppression of stimulated Brillouin scattering in optical bers using ber Bragg gratings, Opt. Express 11, 34673472 (2003). H. Li and K. Ogusu, Dynamic behavior of stimulated Brillouin scattering in a single-mode optical ber, Jpn. J. Appl. Phys. 38, 63096315 (1999). A. Djupsjbacka, C. Jacobsen, and B. Tromborg, Dynamic stimulated Brillouin scattering analysis, J. Lightwave Technol. 18, 416424 (2000). V. Grimalsky, S. Koshevaya, G. Burlak, and B. Salazar-H, Dynamic effects of the stimulated Brillouin scattering in bers due to acoustic diffraction, J. Opt. Soc. Am. B 19, 689694 (2002). E. M. Dianov, A. V. Luchnikov, A. N. Pilipetskii, and A. N. Starodumov, Electrostriction mechanism of soliton interaction in optical bers, Opt. Lett. 15, 314316 (1990). E. M. Dianov, A. V. Luchnikov, A. N. Pilipetskii, and A. N. Starodumov, Long-range interaction of soliton pulse trains in a single-mode bre, Sov. Lightwave Commun. 1, 3743 (1991).
48

Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001

60. C. Montes and A. M. Rubenchik, Stimulated Brillouin scattering from trains of solitons in optical bers: information degradation, J. Opt. Soc. Am. B 9, 18571875 (1992). 61. A. N. Pilipetskii, A. V. Luchnikov, and A. M. Prokhorov, Soliton pulse long-range interaction in optical bres: the role of light polarization and bre geometry, Sov. Lightwave Commun. 3, 2939 (1993). 62. E. A. Golovchenko and A. N. Pilipetskii, Acoustic effect and the polarization of adjacent bits in soliton communication lines, J. Lightwave Technol. 12, 10521056 (1994). 63. A. Fellegara and S. Wabnitz, Electrostrictive cross-phase modulation of periodic pulse trains in optical bers, Opt. Lett. 23, 13571359 (1998). 64. Y. Jaoun, L. du Mouza, and G. Debarge, Electrostriction-induced acoustic effect in ultralong-distance soliton transmission systems, Opt. Lett. 23, 11851187 (1998). 65. D. A. Fishman and J. A. Nagel, Degradations due to stimulated Brillouin scattering in multigigabit intensity-modulated ber-optic systems, J. Lightwave Technol. 11, 17211728 (1993). 66. F. Forghieri, R. W. Tkach, and A. R. Chraplyvy, Fiber nonlinearities and their impact on transmission systems, in Optical Fiber Telecommunications III, I. P. Kaminov and T. L. Koch, eds. (Academic, 1997), vol. A, pp. 196264. 67. X. P. Mao, G. E. Bodeep, R. W. Tkach, A. R. Chraplyvy, T. E. Darcie, and R. M. Derosier, Brillouin scattering in externally modulated lightwave AMVSB CATV transmission systems, IEEE Photon. Technol. Lett. 4, 287 289 (1992). 68. F. W. Willems, W. Muys, and J. S. Leong, Simultaneous suppression of stimulated Brillouin scattering and interferometric noise in externally modulated lightwave AM-SCM systems, IEEE Photon. Technol. Lett. 6, 14761478 (1994). 69. F. W. Willems, J. C. van der Plaats, and W. Muys, Harmonic distortion caused by stimulated Brillouin scattering suppression in externally modulated lightwave AM-CATV systems, Electron. Lett. 30, 343345 (1994). 70. I. L. Fabelinskii, Molecular Scattering of Light (Plenum, 1968). 71. B. Ya. Zeldovich, N. F. Pilipetsky, and V. V. Shkunov, Principles of Phase Conjugation (Springer-Verlag, 1985), chap. 2. 72. R. W. Boyd, Nonlinear Optics, 2nd ed. (Academic, 2003), chap. 9. 73. R. H. Pantell and H. E. Puthoff, Fundamentals of Quantum Electronics (Wiley, 1969). 74. A. Yariv, Quantum Electronics, 3d ed. (Wiley, 1989). 75. G. P. Agrawal, Nonlinear Fiber Optics, 3rd ed. (Academic, 2001), chap. 9. 76. C. C. Lee and S. Chi, Measurement of stimulated Brillouin scattering threshold for various types of bers using Brillouin optical time-domain reectometer, IEEE Photon. Technol. Lett. 12, 672674 (2000). 77. R. M. Shelby, M. D. Levenson, and P. W. Bayer, Resolved forward Brillouin scattering in optical bers, Phys. Rev. Lett. 54, 939942 (1985). 78. J. A. Buck, Fundamentals of Optical Fibers (Wiley Interscience, 1995). 79. K. Okamoto, Fundamentals of Optical Waveguides (Academic, 2000). 80. R. A. Waldron, Some problems in the theory of guided microsonic waves, IEEE Trans. Microwave Theory Tech. MTT-17, 893904 (1969). 81. R. N. Thurston, Elastic waves in rods and clad rods, J. Acoust. Soc. Am. 64, 137 (1978). 82. P. J. Thomas, N. L. Rowell, H. M. van Driel, and G. I. Stegeman, Normal
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 49

acoustic modes and Brillouin scattering in single-mode optical bers, Phys. Rev. B 19, 49864998 (1979). 83. J. D. Achenbach, Wave Propagation in Elastic Solids (North Holland, 1973). 84. B. A. Auld, Acoustic Fields and Waves in Solids, 2nd ed. (Krieger, 1990). 85. A. Safaai-Jazi, C.-K. Jen, and G. W. Farnell, Analysis of weakly guiding ber acoustic waveguide, IEEE Trans. Ultrason. Ferroelectr. Freq. Control UFFC-33, 5968 (1986). 86. C.-K. Jen, A. Safaai-Jazi, and G. W. Farnell, Leaky modes in weakly guided ber acoustic waveguides, IEEE Trans. Ultrason. Ferroelectr. Freq. Control UFFC-33, 634643 (1986). 87. A. Safaai-Jazi and R. O. Claus, Acoustic modes in optical berlike waveguides, IEEE Trans. Ultrason. Ferroelectr. Freq. Control UFFC-35, 619627 (1988). 88. K. Tajima, Exact acoustic leaky wave solutions for single-mode bres, Electron. Lett. 27, 251253 (1991). 89. J. Qu and L. Jacobs, Cylindrical waveguides and their applications in ultrasonic evaluation, in Ultrasonic Nondestructive Evaluation, T. Kundu, ed. (CRC Press, 2004). 90. E. Peral and A. Yariv, Degradation of modulation and noise characteristics of semiconductor lasers after propagation in optical ber due to shift induced by stimulated Brillouin scattering, IEEE J. Quantum Electron. 35, 11851195 (1999). 91. G. Canat, A. Durcu, G. Lesueur, L. Lombard, P. Bourdon, V. Jolivet, and Y. Jaoun, Characteristics of the Brillouin spectra in erbiumytterbium bers, Opt. Express 16, 32123222 (2008). 92. W. Zou, Z. He, and K. Hotate, Two-dimensional nite-element modal analysis of Brillouin gain spectra in optical bers, IEEE Photon. Technol. Lett. 18, 24872489 (2006). 93. W. Zou, Z. He, and K. Hotate, Analysis on the inuence of intrinsic thermal stress on Brillouin gain spectra in optical bers, Proc. SPIE 6371, 637104 (2006). 94. W. Zou, Z. He, and K. Hotate, Acoustic modal analysis and control in w-shaped triple-layer optical bers with highly-germanium-doped core and F-doped innner cladding, Opt. Express 16, 1000610017 (2008). 95. R. G. Smith, Optical power handling capacity of low loss optical bers as determined by stimulated Raman and Brillouin scattering, Appl. Opt. 11, 24892494 (1972). ewski, M. Levenstein, and M. G. Raymer, Statistics of stimulated 96. K. Rzz Stokes pulse energies in the steady-state regime, Opt. Commun. 43, 451 454 (1982). ewski, and P. Narum, Noise initiation of stimulated 97. R. W. Boyd, K. Rzz Brillouin scattering, Phys. Rev. A 42, 55145521 (1990). 98. A. L. Gaeta and R. W. Boyd, Stochastic dynamics of stimulated Brillouin scattering in an optical ber, Phys. Rev. A 44, 32053209 (1991). 99. W. Jinsong, T. Weizhong, and Z. Wen, Stimulated Brillouin scattering initiated by thermally excited acoustic waves in absorption media, Opt. Commun. 123, 574576 (1996). 100. A. A. Fotiadi, R. Kiyan, O. Deparis, P. Megret, and M. Blondel, Statistical properties of stimulated Brillouin scattering in single-mode optical bers above threshold, Opt. Lett. 27, 8385 (2002). 101. S. Le Floch and P. Cambon, Theoretical evaluation of the Brillouin
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 50

102.

103.

104.

105.

106. 107. 108. 109. 110. 111.

112.

113.

114.

115. 116. 117.

118.

119.

120.

threshold and the steady-state Brillouin equations in standard single-mode optical bers, J. Opt. Soc. Am. A 20, 11321137 (2003). A. Kobyakov, S. A. Darmanyan, and D. Chowdhury, Exact analytical treatment of noise initiation of SBS in the presence of loss, Opt. Commun. 260, 4649 (2006). P. Bayvel and P. M. Radmore, Solutions of the SBS equations in single mode optical bres and implications for bre transmission systems, Electron. Lett. 26, 434436 (1990). R. D. Esman and K. J. Williams, Brillouin scattering: beyond threshold, in Optical Fiber Communication Conference, vol. 2 of 1996 OSA Technical Digest Series (Optical Society of America, 1996), paper ThF5. J. C. Beugnot, T. Sylvestre, D. Alasia, H. Maillotte, V. Laude, A. Monteville, L. Provino, N. Traynor, S. F. Mafang, and L. Thvenaz, Complete experimental characterization of stimulated Brillouin scattering in photonic crystal ber, Opt. Express 15, 1551715522 (2007). V. I. Kovalev and R. G. Harrison, Threshold for stimulated Brillouin scattering in optical ber, Opt. Express 15, 1762517630 (2007). T. H. Russell and W. B. Roh, Threshold of second-order stimulated Brillouin scattering in optical ber, J. Opt. Soc. Am. B 19, 23412345 (2002). M. Abramowitz and I. Stegun, Handbook of Mathematical Functions (Dover, 1965). G. B. Arfken and H. J. Weber, Mathematical Methods for Physicists, 5th ed. (Academic, 2001). For this ber, there is another strong peak in the BGS at 10.9 GHz. K. Shiraki, M. Ohashi, and M. Tateda, Performance of strain-free stimulated Brillouin scattering suppression ber, J. Lightwave Technol. 14, 549554 (1996). C. McIntosh, A. Yeniay, J. Toulouse, and J. M. P. Delavaux, Stimulated Brillouin scattering in dispersion-compensating bers, Opt. Fiber Technol. 3, 173176 (1997). J. H. Lee, Z. Yusoff, W. Belardi, M. Ibsen, T. M. Monro, and D. J. Richardson, Investigation of Brillouin effects in small-core holey optical ber: lasing and scattering, Opt. Lett. 27, 927929 (2002). F. Poletti, K. Furusawa, Z. Yusoff, N. G. R. Broderick, and D. J. Richardson, Nonlinear tapered holey bers with high stimulated Brillouin scattering threshold and controlled dispersion, J. Opt. Soc. Am. B 24, 2185 2194 (2007). D. Cotter, Stimulated Brillouin scattering in monomode optical ber, J. Opt. Commun. 4, 1019 (1983). X. S. Yao, Brillouin selective sideband amplication of microwave photonic signals, IEEE Photon. Technol. Lett. 10, 138140 (1998). A. Loayssa, D. Benito, and M. J. Garde, Applications of optical carrier Brillouin processing to microwave photonics, Opt. Fiber Technol. 8, 2442 (2002). T. Tanemura, Y. Takushima, and K. Kikuchi, Narrowband optical lter, with a variable transmission spectrum, using stimulated Brillouin scattering in optical ber, Opt. Lett. 27, 15521554 (2002). Y. Shen, X. Zhang, and K. Chen, A simple lter based on stimulated Brillouin scattering for carrier-suppression of microwave photonic signals, Proc. SPIE 5625, 109116 (2005). S. Tonda-Goldstein, D. Dol, J.-P. Huignard, G. Charlet, and J. Chazelas,

Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001

51

121.

122.

123.

124.

125.

126.

127. 128.

129. 130.

131.

132.

133.

134. 135.

136.

137.

Stimulated Brillouin scattering for microwave signal modulation depth increase in optical links, Electron. Lett. 36, 944946 (2000). M. J. LaGasse, W. Charczenko, M. C. Hamilton, and S. Thaniyavarn, Optical carrier ltering for high dynamic range bre optic links, Electron. Lett. 30, 21572158 (1994). K. J. Williams and R. D. Esman, Stimulated Brillouin scattering for improvement of microwave bre-optic link efciency, Electron. Lett. 30, 19651966 (1994). A. Wiberg and P. O. Hedekvist, Photonic microwave generator utilizing narrowband Brillouin amplication and ber-based oscillator, Proc. SPIE 5466, 148156 (2004). T. Schneider, M. Junker, and D. Hannover, Generation of millimetrewave signals by stimulated Brillouin scattering for radio over bre systems, Electron. Lett. 40, 15001501 (2004). L. Xing, L. Zhan, S. Luo, and Y. Xia, High-power low-noise ber Brillouin amplier for tunable slow-light delay buffer, IEEE J. Quantum Electron. 44, 11331138 (2008). W. H. Press, S. A. Teukolsky, W. T. Vetterling, and B. P. Flannery, Numerical Recipies in C. The Art of Scientic Computing, 2nd ed., (Cambridge Univ. Press, 1995), chap. 17. C. L. Tang, Saturation and spectral characteristics of the Stokes emission in the stimulated Brillouin process, J. Appl. Phys. 37, 29452955 (1966). L. Chen and X. Bao, Analytical and numerical solution for steady state stimulated Brillouin scattering in a single-mode ber, Opt. Commun. 152, 6570 (1998). R. H. Enns and L. P. Batra, Saturation and depletion in stimulated light scattering, Phys. Lett. 28A, 591592 (1969). M. Vasilyev and A. Kobyakov, Effect of pump depletion on the noise gure of distributed Raman ampliers, in Conference on Lasers and ElectroOptics/Quantum Electronics and Laser Science Conference, Technical Digest (Optical Society of America, 2003), paper CWL3. A. Kobyakov, S. A. Darmanyan, M. Sauer, and D. Chowdhury, High-gain Brillouin amplication: an analytical approach, Opt. Lett. 31, 19601962 (2006). Y. Y. Huang, A. Sarkar, and P. C. Schultz, Relationship between composition, density and refractve index for germania silica glasses, J. NonCryst. Solids 27, 2937 (1978). N. Lagakos, J. A. Bucaro, and R. Hughes, Acoustic sensitivity predictions of single-mode optical bers using Brillouin scattering, Appl. Opt. 19, 36683670 (1980). S. T. Gulati and J. D. Helnstine, Fatigue behavior of GeO2 SiO2 glasses, Mater. Res. Soc. Symp. Proc. 531, 133137 (1998). S. R. Bickham, A. Kobyakov, and S. Li, Nonlinear optical bers with increased SBS thresholds, in Optical Fiber Communication Conference and Exposition and The National Fiber Optic Engineers Conference, Technical Digest (CD) (Optical Society of America, 2006), paper OTuA3. W. Zou, Z. He, and K. Hotate, Experimental study of Brillouin scattering in uorine-doped single-mode optical bers, Opt. Express 16, 18804 18812 (2008). X. P. Mao, R. W. Tkach, A. R. Chraplyvy, R. M. Jopson, and R. M. Derosier, Stimulated Brillouin threshold dependence on ber type and uniformity, IEEE Photon. Technol. Lett. 4, 6669 (1992).
52

Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001

138. C. A. S. de Oliveira, C. K. Jen, A. Shang, and C. Saravanos, Stimulated Brillouin scattering in cascaded bers of different Brillouin frequency shift, J. Opt. Soc. Am. B 10, 969972 (1993). 139. K. Shiraki, M. Ohashi, and M. Tateda, SBS threshold of a ber with a Brillouin frequency shift distribution, J. Lightwave Technol. 14, 5057 (1996). 140. A. Kobyakov, M. Sauer, and J. E. Hurley, SBS threshold of segmented bers, in Optical Fiber Communication Conference and Exposition and The National Fiber Optic Engineers Conference, Technical Digest (CD) (Optical Society of America, 2005), paper OME5. 141. S. Rae, I. Bennion, and M. J. Cardwell, New numerical model for stimulated Brillouin scattering in optical bers with nonuniformity, Opt. Commun. 123, 611616 (1996). 142. Y. Yamamoto, T. Miyamoto, M. Onishi, and E. Sasaoka, Zero-water-peak pure-silica-core ber compatible with ITU-T G.652 single-mode ber and its applicability to access networks, in Optical Fiber Communication Conference and Exposition and The National Fiber Optic Engineers Conference, Technical Digest (CD) (Optical Society of America, 2005), paper JWA63. 143. P. S. Devgan, V. J. Urick, K. J. Williams, and J. F. Diehl, Long-haul microwave analog link with shot-noise-limited performance above the stimulated Brillouin scattering threshold, in 2008 International Topical Meeting on Microwave Photonics and 2008 Asia-pacic Microwave Photonics Conference (IEEE, 2009), pp. 326329. 144. F. W. Willems and W. Muys, Suppression of interferometric noise in externally modulated lightwave AM-CATV systems by phase modulation, Electron. Lett. 29, 20622063 (1993). 145. N. Yoshizawa and T. Imai, Stimulated Brillouin scattering suppression by means of applying strain distribution to ber with cabling, J. Lightwave Technol. 11, 15181522 (1993). 146. J. M. C. Boggio, J. D. Marconi, and H. L. Fragnito, Experimental and numerical investigation of the SBS-threshold increase in an optical ber by applying strain distributions, J. Lightwave Technol. 23, 38083814 (2005). 147. K. Shiraki, M. Ohashi, and M. Tateda, Suppression of stimulated Brillouin scattering in a bre by changing the core radius, Electron. Lett. 31, 668669 (1995). 148. J. Hansryd, F. Dross, M. Westlund, P. A. Andrekson, and S. N. Knudsen, Increase in the SBS threshold in a short highly nonlinear ber by applying a temperature distribution, J. Lightwave Technol. 19, 16911697 (2001). 149. M. Ohashi and M. Tateda, Design of strain-free-ber with nonuniform dopant concentration for stimulated Brillouin scattering suppression, J. Lightwave Technol. 11, 19411945 (1993). 150. H. Al-Raweshidy and S. Komaki, eds., Radio over Fiber Technologies for Mobile Communications Networks (Artech House, 2002). 151. A. Kobyakov, M. Sauer, N. Nishiyama, A. Chamarti, F. Annunziata, J. Hurley, C. Caneau, J. George, and C.-E. Zah, 802.11a/g WLAN radio transmission at 1.3 m over 1.1 km multimode and 30 km standard singlemode ber using InP VCSEL, in European Conference on Optical Communications, 2006. ECOC 2006 (2006), paper Tu1.6.1. 152. X. Qian, A. Wonfor, R. V. Penty, and I. H. White, Overcoming transmission impairments in wide frequency range radio-over-bre distribution
Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 53

153.

154.

155.

156.

157. 158. 159. 160. 161.

162.

163.

164.

165.

166.

167. 168.

169.

170.

systems, in European Conference on Optical Communications, 2006. ECOC 2006 (2006), paper We3. H. Le Bras, M. Moignard, and B. Charbonnier, Brillouin scattering in radio over ber transmission, in National Fiber Optic Engineers Conference, OSA Technical Digest Series (CD) (Optical Society of America, 2007), paper JWA86. M. Sauer, A. Kobyakov, and A. B. Rufn, Radio-over-ber transmission with mitigated stimulated Brillouin scattering, IEEE Photon. Technol. Lett. 19, 14871489 (2007). R. B. Ellis, F. Weiss, and O. M. Anton, HFC and PON-FTTH networks using higher SBS threshold singlemode optical bre, Electron. Lett. 43, 405407 (2007). M. D. Vaughn, A. B. Rufn, A. Kobyakov, A. Woodn, C. Mazzali, R. Whitman, A. Boskovic, R. E. Wagner, D. Kozischek, and D. Meis, Techno-economic study of the value of high stimulated Brillouin scattering threshold single-mode ber utilization in ber-to-the-home access networks, J. Opt. Netw. 5, 4057 (2006). M. Sauer, A. Kobyakov, and J. George, Radio over ber for picocellular network architectures, J. Lightwave Technol. 25, 33013320 (2007). M. Islam, ed., Raman Ampliers for Telecommunications (Springer, 2004). C. Headley and G. P. Agrawal, Raman Amplication in Fiber-Optical Communication Systems (Elsevier, 2004). A. Kobyakov, Prospects of Raman-assisted transmission systems, Proc. SPIE 5246, 174188 (2003). R. Chi, K. Lu, X. Dong, W. Chen, G. Yang, and Z. Liu, Gain saturation and nonlinear effect of erbium-doped ber amplier/discrete compensating Raman amplier hybrid ber ampliers in the C-band, Opt. Eng. 43, 346349 (2004). L. Grner-Nielsen, S. N. Knudsen, B. Edvold, T. Veng, D. Magnussen, C. C. Larsen, and H. Damsgaard, Dispersion compensating bers, Opt. Fiber Technol. 6, 164180 (2000). M. Mehendale, A. Kobyakov, M. Vasilyev, S. Tsuda, and A. F. Evans, Effect of Raman amplication on stimulated Brillouin scattering in dispersion compensating bres, Electron. Lett. 38, 268269 (2002). T. Okuno and M. Nishimura, Effects of stimulated Raman amplication in optical bre on stimulated Brillouin scattering threshold power, Electron. Lett. 38, 1416 (2002). L. Thvenaz, Slow and fast light using stimulated Brillouin scattering: a highly exible approach, in Slow LightScience and Applications, J. B. Khurgin and R. S. Tucker, eds. (CRC Press, 2009), chap. 9. G. Qin, H. Sotobayashi, M. Tsuchiya, A. Mori, and Y. Ohishi, Stimulated Brillouin amplication in a tellurite ber as a potential system for slow light generation, Jpn. J. Appl. Phys. 46, L810L812 (2007). L. Thvenaz, Slow and fast light in optical bres, Nat. Photonics 2, 474 481 (2008). Y. Okawachi, M. S. Bigelow, J. E. Sharping, Z. Zhu, A. Schweinsberg, D. J. Gauthier, R. W. Boyd, and A. L. Gaeta, Tunable all-optical delays via Brillouin slow light in an optical ber, Phys. Rev. Lett. 94, 153902 (2005). L. Ren and Y. Tomita, SBS-based slow light in optical bers: optimum design considerations for undistorted slow-light signal propagation in steady-state and transient regimes, Proc. SPIE 7226, 722605 (2009). K.-Y. Song, M. G. Herrez, and L. Thvenaz, Observation of pulse delay54

Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001

171. 172.

173.

174.

175.

176. 177.

178. 179.

180.

181.

182.

183.

184.

185.

186.

187.

ing and advancement in optical bers using stimulated Brillouin scattering, Opt. Express 13, 8288 (2005). Z. Zhu, D. J. Gauthier, and R. W. Boyd, Stored light in an optical ber via stimulated Brillouin scattering, Science 318, 17481750 (2007). M. D. Stenner, M. A. Neifeld, Z. Zhu, A. M. C. Dawes, and D. J. Gauthier, Distortion management in slow-light pulse delay, Opt. Express 13, 999510002 (2005). K.-Y. Song, M. G. Herrez, and L. Thvenaz, Gain assisted pulse advancement using single and double Brillouin gain peaks in optical bers, Opt. Express 13, 97589765 (2005). Z. Shi, R. Pant, Z. Zhu, M. D. Stenner, M. A. Neifeld, D. J. Gauthier, and R. W. Boyd, Design of a tunable time-delay element using multiple gain lines for increased fractional delay with high data delity, Opt. Lett. 32, 19861988 (2007). T. Sakamoto, T. Yamamoto, K. Shiraki, and T. Kurashima, Low distortion slow light in at Brillouin gain spectrum by using optical frequency comb, Opt. Express 16, 80268032 (2008). M. G. Herrez, K.-Y. Song, and L. Thvenaz, Arbitrary-bandwidth Brillouin slow light in optical bers, Opt. Express 14, 13951400 (2006). Z. Zhu, A. M. C. Dawes, D. J. Gauthier, L. Zhang, and A. E. Willner, Broadband SBS slow light in an optical ber, J. Lightwave Technol. 25, 201206 (2007). K. Y. Song and K. Hotate, 25 GHz Brillouin slow light in optical bers, Opt. Lett. 32, 217219 (2007). R. Pant, M. D. Stenner, M. A. Neifeld, Z. Shi, R. W. Boyd, and D. J. Gauthier, Maximizing the opening of eye diagrams for slow-light systems, Appl. Opt. 46, 65136519 (2007). B. Zhang, L. Yan, I. Fazal, L. Zhang, A. E. Willner, Z. Zhu, and D. J. Gauthier, Slow light on Gbit/s differential-phase-shift-keying signals, Opt. Express 15, 18781883 (2007). T. Horiguchi, K. Shimizu, T. Kurashima, M. Tateda, and Y. Koyamada, Development of a distributed sensing technique using Brillouin scattering, J. Lightwave Technol. 13, 12961302 (1995). X. Bao, J. Dhliwayo, N. Heron, D. J. Webb, and D. A. Jackson, Experimental and theoretical studies on a distributed temperature sensor based on Brillouin scattering, J. Lightwave Technol. 13, 13401346 (1995). L. Thvenaz, M. Facchini, A. Fellay, P. Robert, D. Inaudi, and B. Dardel, Monitoring of large structures using distributed Brillouin ber sensing, Proc. SPIE 3746, 345348 (1999). K. T. V. Graffan and B. T. Meggitt, eds., Optical Fiber Sensor Technology : Volume 4: Chemical and Environmental Sensing (Kluwer Academic, 1999). Y. Li, F. Zhang, and T. Yoshino, Wide-range temperature dependence of Brillouin shift in a dispersion-shifted ber and its annealing effect, J. Lightwave Technol. 21, 16631667 (2003). S. Le Floch and P. Cambon, Study of Brillouin gain spectrum in standard single-mode optical ber at low temperatures 1.4 370 K and high hydrostatic pressures 1 250 bars, Opt. Commun. 219, 395410 (2003). W. Zou, Z. He, and K. Hotate, Investigation of strain- and temperaturedependences of Brillouin frequency shifts in GeO2-doped optical bers, J. Lightwave Technol. 26, 18541861 (2008).

Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001

55

188. S. Yin, P. B. Rufn, and F. T. S. Yu, eds., Fiber Optic Sensors, 2nd ed. (CRC Press, 2008). 189. M. Tateda, First measurement of strain distribution along eld installed optical bers using Brillouin spectroscopy, J. Lightwave Technol. 8, 12691272 (1990). 190. F. Ravet, L. Zou, X. Bao, L. Chen, R. F. Huang, and H. A. Khoo, Pipeline buckling detection by the distributed Brillouin sensor, in Sensing Issues in Civil Structural Health Monitoring, F. Ansari, ed. (Springer, 2005), pp. 515524. 191. L. Zou, G. A. Ferrier, S. Afshar, Q. Yu, L. Chen, and X. Bao, Distributed Brillouin scattering sensor for discrimination of wall-thinning defects in steel pipe under internal pressure, Appl. Opt. 43, 15831588 (2004). 192. X. Bao, Optical ber sensors based on Brillouin scattering, Opt. Photonics News 20(9), 4145 (2009). 193. J. Limpert, F. Roser, S. Klingebiel, T. Schreiber, C. Wirth, T. Peschel, R. Eberhardt, and A. Tnnermann, The rising power of ber lasers and ampliers, IEEE J. Sel. Top. Quantum Electron. 13, 537545 (2007). 194. S. Gray, D. T. Walton, X. Chen, J. Wang, M.-J. Li, A. Liu, A. B. Rufn, J. A. Demeritt, and L. A. Zenteno, Optical bers with tailored acoustic speed proles for suppressing stimulated Brillouin scattering in highpower, single-frequency sources, J. Lightwave Technol. 15, 3746 (2009). 195. M. D. Mermelstein, S. Ramachandran, J. M. Fini, and S. Ghalmi, SBS gain efciency measurements and modeling in a 1714 m2 effective area LP08 higher order mode optical ber, Opt. Express 15, 1595215963 (2007). 196. A. Liem, J. Limpert, H. Zellmer, and A. Tnnermann, 100-W single frequency master-oscillator ber power amplier, Opt. Lett. 28, 15371539 (2003). 197. D. N. Payne, Y. Jeong, J. Nilsson, J. K. Sahu, D. B. S. Soh, C. Alegria, P. Dupriez, C. A. Codemard, V. N. Philippov, V. Hernandez, R. Horley, L. Hickey, L. Wanzcyk, C. E. Chryssou, J. A. Alvarez-Chavez, and P. Turner, Kilowatt-class single-frequency ber sources, Proc. SPIE 5709, 133 141 (2005). 198. V. I. Kovalev and R. G. Harrison, Suppression of stimulated Brillouin scattering in high-power single-frequency ber ampliers, Opt. Lett. 31, 161163 (2006). 199. D. P. Machewirth, Q. Wang, B. Samson, K. Tankala, M. OConnor, and M. Alam, Current developments in high-power monolithic polarization maintaining ber ampliers for coherent beam combining applications, Proc. SPIE 6453, 64531 (2007). 200. M.-J. Li, X. Chen, J. Wang, S. Gray, A. Liu, J. A. Demeritt, A. B. Rufn, A. M. Crowley, D. T. Walton, and L. A. Zenteno, Al/ Ge co-doped large mode area ber with high SBS threshold, Opt. Express 15, 82908299 (2007). 201. M. D. Mermelstein, M. J. Andrejco, J. Fini, A. Yablon, C. Headley, D. J. DiGiovanni, and A. H. McCurdy, 11.2 dB gain suppression in a large mode area Yb-doped optical ber, Proc. SPIE 6873, 68730 (2008). 202. S. Gray, A. Liu, D. T. Walton, J. Wang, M.-J. Li, X. Chen, A. B. Rufn, J. A. Demeritt, and L. A. Zenteno, 502 Watt, single transverse mode, narrow linewidth, bidirectionally pumped Yb-doped ber amplier, Opt. Express 15, 1704417050 (2007). 203. A. Mocofanescu, L. Wang, R. Jain, K. D. Shaw, P. R. Peterson, and A. Gavrielides, Experimental and theoretical investigations on stimulated BrilAdvances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001 56

204.

205.

206.

207.

208.

209. 210.

211.

212.

213.

214.

215.

216.

217.

218.

219.

louin scattering (SBS) in multimode bers at 1550 nm wavelength, Proc. SPIE, 5581, 654661 (2004). A. Mocofanescu, L. Wang, R. Jain, K. Shaw, A. Gavrielides, P. Peterson, and M. Sharma, SBS threshold for single mode and multimode GRIN bers in an all ber conguration, Opt. Express 13, 20192024 (2005). A. Fotiadi and E. A. Kuzin, Stimulated Brillouin scattering associated with hypersound diffraction in multimode optical bers, presented at Quantum Electronics and Laser Science Conference, Anaheim, Calif, June 27 1996, paper QFC4. K. Tei, Y. Tsuruoka, T. Uchiyama, and T. Fujioka, Critical power of stimulated Brillouin scattering in multimode optical bers, Jpn. J. Appl. Phys. 40, 31913194 (2001). V. I. Kovalev and R. G. Harrison, Waveguide-induced inhomogeneous spectral broadening of stimulated Brillouin scattering in optical ber, Opt. Lett. 27, 20222024 (2002). S. Yoo, J. K. Sahu, and J. Nilsson, Optimized acoustic refractive index proles for suppression of stimulated Brillouin scattering in large core bers, in Optical Fiber Communication Conference, OSA Technical Digest (CD) (Optical Society of America, 2009), paper JWA5. P. St. J. Russell, Photonic-crystal bers, J. Lightwave Technol. 24, 4729 4749 (2006). P. Dainese, P. St. J. Russell, N. Joly, J. C. Knight, G. S. Wiederhecker, H. L. Fragnito, V. Laude, and A. Khelif, Stimulated Brillouin scattering from multi-GHz-guided acoustic phonons in nanostructured photonic crystal bres, Nat. Phys. 2, 388392 (2006). P. Dainese, P. St. J. Russell, G. S. Wiederhecker, N. Joly, H. L. Fragnito, V. Laude, and A. Khelif, Raman-like light scattering from acoustic phonons in photonic crystal ber, Opt. Express 14, 41414150 (2006). C. Fortier, J. Fatome, S. Pitois, F. Smektala, G. Millot, J. Troles, F. Desevedavy, P. Houizot, L. Brilland, and N. Traynor, Experimental investigation of Brillouin and Raman scattering in a 2SG sulde glass microstructured chalcogenide ber, Opt. Express 16, 93989404 (2008). J. E. McElhenny, R. K. Pattnaik, J. Toulouse, K. Saitoh, and M. Koshiba, Unique characteristic features of stimulated Brillouin scattering in smallcore photonic crystal bers, J. Opt. Soc. Am. B 25, 582593 (2008). J. E. McElhenny, R. Pattnaik, and J. Toulouse, Polarization dependence of stimulated Brillouin scattering in small-core photonic crystal bers, J. Opt. Soc. Am. B 25, 21072115 (2008). D. Elser, U. L. Andersen, A. Korn, O. Glckl, S. Lorenz, C. Marquardt, and G. Leuchs, Reduction of guided acoustic wave Brillouin scattering in photonic crystal bers, Phys. Rev. Lett. 97, 133901 (2006). J.-C. Beugnot, T. Sylvestre, H. Maillotte, G. Mlin, and V. Laude, Guided acoustic wave Brillouin scattering in photonic crystal bers, Opt. Lett. 32, 1719 (2007). G. S. Wiederhecker, A. Brenn, H. L. Fragnito, and P. St. J. Russell, Coherent control of ultrahigh-frequency acoustic resonances in photonic crystal bers, Phys. Rev. Lett. 100, 203903 (2008). S. Yang, H. Chen, C. Qiu, M. Chen, M. Chen, S. Xie, J. Li, and W. Chen, Slow-light delay enhancement in small-core pure silica photonic crystal ber based on Brillouin scattering, Opt. Lett. 33, 9597 (2008). C. Vassalo, Optical Waveguide Concepts (Elsevier, 1991).

Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001

57

Andrey Kobyakov graduated with a Master of Science degree with distinctions in Electrical and Computer Engineering from the Moscow Institute of Physics and Technology, Russian in 1992. In 1998, he received his Dr. rer.nat. (Ph.D.) in Optics, graduating magna cum laude from Friedrich-Schiller University in Jena, Germany. Upon completion of his degree, Dr. Kobyakov worked at the Photonics Group of Friedrich-Schiller University in Jena, studying discrete solitons in optical waveguide arrays and all-optical effects in media with quadratic nonlinearity. From 1999 to 2001 he worked at the School of Optics/CREOL, University of Central Florida, USA, as a post-doctoral Fellow, where he did research in nonlinear absorption and optical limiting. In 2001, Andrey joined Corning Incorporated at the Photonics Research and Test Center in Somerset, New Jersey, as a Senior Research Scientist. In 2002 he came to Corning, New York, to work with the Science and Technology Division. He was promoted to Research Associate in 2006. At Corning, Inc., Dr. Kobyakovs research areas include nonlinear effects in optical bers, in particular, stimulated Brillouin scattering, long-haul optical transmission systems, Raman ampliers, photonic bandgap bers, photonic metamaterials, nanoplasmonic structures, and photovoltaics. He has also been involved in the study of radio-over-ber transmission and wireless networks. Dr. Kobyakov has authored and coauthored more than a hundred technical publications in peer-reviewed journals and conference proceedings. He was been awarded two patents; six other patent applications are pending. Dr. Kobyakov is a member of the Optical Society of America (OSA).

Michael Sauer is a Research Associate at the Science and Technology division of Corning Incorporated in Corning, New York, where he is responsible for high-speed optical networks and communication research. His interests include ber-wireless system design, high-speed ber-optic transmission systems, digital signal processing techniques, modulation formats for high data rate systems, signal conditioning with ber-based components, optical network architectures, and optical packet switching. Prior to joining Corning in 2001, he was a Research Scientist at the Communications Laboratory of Dresden University of Technology. His research areas include ber Bragg gratings, generation and transmission of millimeter-wave signals, and architectures of millimeterwave communications systems. He received a Dr.-Ing. (Ph.D.) degree in electrical engineering from Dresden University of Technology, Germany, in 2000. Dr. Sauer is a member of the IEEE Photonics Society and the IEEE Communications Society.

Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001

58

Dipak Chowdhury received his B.Sc. degree in electrical engineering from Bangladesh University of Engineering and Technology, Dhaka, Bangladesh, in 1986 and the M.S. and Ph.D. degrees in electrical engineering from Clarkson University, Potsdam, New York, in 1989 and 1991, respectively. From 1991 to 1993, he had a joint appointment as a Research Associate in the applied physics department at Yale University, New Haven, Connecticut, and in the electrical engineering department at New Mexico State University, Las Cruces, New Mexico. For his graduate and postgraduate work, he performed numerical modeling of, and experimental on, optical scattering from nonlinear microcavities. He joined the research and development facility of Corning, Incorporated, Corning, New York, in 1993. Since joining Corning, he focused his research effort on various aspects of optical communication systems. From 2001 till 2007 he was Research Director of Modeling and Simulation, managing optical system and device modeling and general process modeling for Corning Research. Currently he is the Director of Corning European Technology Center, Avon, France, and President, Corning, S.A.S. His research interests include linear and nonlinear devices, impairments in ber-optic systems and networks, e.g., nonlinearities, crosstalk, chromatic dispersion, and polarization-mode dispersion, and efcient algorithms for stimulating ber-optic systems.

Advances in Optics and Photonics 2, 159 (2010) doi:10.1364/AOP.2.000001

59

You might also like