Sunteți pe pagina 1din 211

OPTICAL SOLITONS:

GENERATION, AMPLIFICATION
AND NONLINEAR EFFECTS

Peter Gerard John Wigley, B.Sc.

Thesis submitted for the degree of Doctor of Philosophy of the University of London

and for the Diploma of Membership of Imperial College

UNIVERSITY OF LONDON

IMPERIAL COLLEGE OF SCIENCE, TECHNOLOGY AND MEDICINE

Department of Physics

Femtosecond Optics Group

November 1990

1
2
ABSTRACT

Nonlinear processes have been applied to signals propagating in the anomalous

dispersion regime of silica optical fibres, to investigate the generation and

amplification of solitons. Amplification is of vital importance to the transmission of

solitons in an all optical communications system, and soliton evolution in the presence

of relatively high gain in the adiabatic regime has been investigated. In particular, the

amplification of noise, modulational instability and CW radiation has been shown to

produce femtosecond soliton structures, which in some cases may be useful. In others,

it may lead to the evolution of error pulses.

A number of sources have been constructed based on fibre oscillators, with

gain provided by Raman or Erbium doped fibre amplification. The sources have

various applications for soliton investigations, operating from 1.0 to 1.6μm. Using

these sources, amplification in doped fibres has been investigated in the picosecond

and femtosecond regime.

The process of modulational instability has been exploited to produce sources

of fixed or tunable repetition rate (THz) soliton-like pulses. These ultra-high repetition
rate trains may be of importance for use as control signals in multiplexing circuits or

routing switches. The process has been induced through cross phase modulation and

through a four wave mixing interaction. Cross phase modulation has been used to

produce soliton-like pulse shaping in a laser diode signal of sub-fundamental soliton

power.

Seeding of the Raman soliton continuum process by a CW signal has been

shown to fix the wavelength of the scattered radiation, producing a source of

fundamental solitons tunable across the Raman gain band.

3
Since noise can evolve into solitons in the presence of gain, a method of

suppressing pedestals, inter-pulse radiation or non-soliton structures has been

considered, based on a nonlinear optical loop mirror. The role of gain or loss in such a

device has been shown to improve the theoretical switching characteristic

considerably.

4
TABLE OF CONTENTS Page No.

ABSTRACT 3

1 GENERAL INTRODUCTION
1.1 Introduction. 9

1.2 Historical Perspective - Solitons In Fibres. 11

1.3 Dispersion in Optical Fibres. 15

1.4 Optical Fibre Nonlinearity. 26

1.5 Stimulated Raman Scattering and Raman Amplification. 42

1.6 References. 48

2 FIBRE RAMAN LASER SOURCES


2.1 Fibre Raman Oscillators. 53

2.2 A CW fibre Raman Laser at 1.41μm. 55

2.3 A Synchronously Pumped Dispersion Compensated

Raman Source at 1.41μm. 59

2.4 References. 66

3 MODE LOCKED DOPED FIBRE LASER SOURCES


3.1 Introduction. 67

3.2 Active Mode Locking of an Erbium Doped Fibre Laser

through Amplitude Modulation. 69

3.3 Active Mode Locking of an Erbium Fibre Laser using an

Intra-Cavity Laser Diode Device. 74

3.4 Mode Locking of a CW Neodymium Doped Fibre Laser with

Linear External Cavity Feedback. 82

3.5 References. 89

5
4 PULSE AMPLIFICATION IN DOPED FIBRES Page No.

4.1 Doped Fibre Amplification. 91

4.2 Picosecond Amplification in Neodymium Doped Fibre. 92

4.3 Femtosecond Amplification in Erbium Doped Fibre. 98

4.4 References. 106

5 MODULATIONAL INSTABILITY AND

CROSS PHASE MODULATION


5.1 Modulational Instability. 107

5.2 Modulational Instability from a Mode Locked

Erbium Fibre Laser Source. 113

5.3 Cross Phase Modulation. 119

5.4 Induced Modulational Instability from Pulses in the

Normal Dispersion Regime. 122

5.5 Picosecond Pulse Generation through

Cross Phase Modulation. 131

5.6 2THz Pulse Train Generation through

Induced Modulational Instability. 140

5.7 References. 146

6 SOLITON RAMAN GENERATION AND

AMPLIFICATION
6.1 Soliton Raman Generation. 147

6.2 Modulational Instability as a Precursor. 148

6.3 Semiconductor Laser Diode Signals as a Precursor. 152

6.4 Stimulated Raman Amplification of Noise Bursts

into Solitons. 161

6.5 Stimulated Raman Amplification of Dispersive Pulses

into Solitons. 167

6.6 References. 171

6
7 NONLINEAR OPTICAL LOOP MIRROR Page No.

PULSE PROCESSING
7.1 Introduction. 173

7.2 Intensity Dependent Transmission through a

Nonlinear Optical Loop Mirror. 174

7.3 Pedestal Suppression and Nonlinear Pulse Shaping

employing a Nonlinear Optical Loop Mirror 180

7.4 References. 191

8 CONCLUSIONS 193

References 195

APPENDIX ULTRASHORT PULSE MEASUREMENT:


AUTOCORRELATION 197

References 204

JOURNAL AND CONFERENCE PUBLICATIONS 205

ACKNOWLEDGEMENTS 208

7
8
1. GENERAL INTRODUCTION

1.1 Introduction

The information capacity of optical fibre is known to be many orders of

magnitude greater than that obtainable with present day transmission media. However,

severe limitations exist in the optical power and speed with which information can be

relayed down a fibre. These limitations have only come to light in recent years, since

absorption losses have been reduced. The advent of low loss high qualtiy silica optical

fibres now enables transmission over more than ~20km with only half power loss.

The main physical limitation imposed is nonlinearity, which for long

transmission lengths results in signal distortion. Amplification within fibres, removing


U U

the necessity for electronic repeaters which limit the system bandwidth, can extend the

system length to over 1000km. Over long lengths, cumulative nonlinearity causes

severe spectral broadening, leading to pulse distortion through dispersion, which is

always present.

Silica has a very small coefficient of nonlinearity, but long propagation

distances multiply the effect, which scales with length. Nonlinearity places a power

limit on the pulses used in transmission and exaggerates the effect of dispersion in

many cases where spectral broadening occurs. The peak power of a pulse induces a

nonlinear polarisation of the transmission medium, which lead to a number of

interesting effects such as self phase modulation and Raman scattering.

If propagation occurs in the anomalous dispersion regime, negative dispersion

can exactly balance the effect of nonlinear spectral broadening, resulting in the

generation of an envelope soliton, which propagates without change in shape. In this

thesis, the generation, amplification and nonlinear propagation of solitons is

considered, with relevance to long distance transmission and ultra high bit rate

9
communications.

A nonlinear transmission system based on solitons must be carefully tailored to

exact requirements, as certain changes in system parameters can lead to large changes

in response. For example, in a high repetition rate system, small changes in inter pulse

spacing can lead to detection errors through pulse-pulse interactions. Optical fibre

provides an ideal environment in which experiments on soliton propagation can be

studied.

In this thesis a variety of techniques are investigated for generating solitons

and soliton-like structures, which are used to illustrate nonlinear coupling. The

application of all fibre amplification to long distance and to ultra-high bit rate

transmission is considered. Two methods are considered: active (doped) fibre and

stimulated Raman amplification. Raman Amplification of noise and sub-fundamental

power pulses are shown to results in soliton generation, and the influence of seed

signals on Raman soliton continuum generation is considered. The pulsed response of

doped fibre amplifiers are considered, in the subpicosecond and femtosecond regimes.

Nonlinear coupling mechanisms are investigated, producing ultrashort pulses

at Terahertz repetition rates, for use as signal or clock pulses in a soliton based

transmission system. The potential for producing Terahertz repetition rate pulse trains

from fibre laser sources is considered. Nonlinear coupling is shown to result in

ultrashort pulse generation through soliton shaping of weak signals in the anomalous

dispersion regime.

Finally, a method for incorporating intensity discrimination in a transmission

line is presented based on an all-fibre interferometer. The influence of gain on such a

device is also considered.

10
1.2 Historical Perspective - Solitons in Fibres

The basis of soliton propagation in optical fibres has been the subject of

scientific interest for more than one hundred years. Numerical and analytical work on

the solitary wave solutions of the nonlinear Schroedinger equation (NLS) have

produced exact solutions using the inverse scattering method1. The NLS describes the

propagation of an envelope wave in a weakly nonlinear, strongly dispersive medium.

The nonlinearity in this case effects the propagation constant so as to make it

dependent on the intensity of the wave envelope.

The NLS was solved numerically in 19732,3 to describe the propagation of

coherent pulses in optical fibres. The two physical properties which make soliton

propagation possible are dispersion and nonlinearity, and hence the NLS is directly

applicable. Dispersion in fibres arises from three sources: modal, waveguide and

material, and always results in pulse broadening, irrespective of its sign, in the absence

of nonlinearity. A pulse of finite duration requires a finite bandwidth to support it, and

hence shorter pulses suffer more relative dispersion than longer ones. However it was

demonstrated numerically that nonlinearity, which relates the wave velocity to the

intensity, can exactly compensate for dispersion, resulting in the possibility of the

preservation of pulse shape in a lossless transmission line.

Solitons exhibit the following attractive properties for use in optical

communications:

i) shape, width and speed of soliton pulses are preserved in the abscence of loss,

i) solitons are stable against small perturbations

iii) solitons survive collisions whilst maintaining shape and speed (elastic

scattering)

iv) solitons represent the common asymptotic state for a variety of completely

different initial pulse states.

11
This latter property is of special significance in chapter 6 of this thesis, where

additional results are presented which verify this statement.

Since solitons in fibres were theoretically proposed2,3, the most important


advance has been reduction of the intrinsic loss present in fibres, enabling longer

propagation distances. Since dispersion occurs over a length scale characteristic of the

initial pulse duration and fibre parameters, ultrashort pulses must propagate over a

sufficient length for dispersive effects to become noticeable. For high power pulse

propagation, nonlinearity can counteract the dispersive tendency, resulting in a marked

decrease or cancellation of the overall pulse broadening.

The first soliton propagation experiments took place in 19805. The

experiments awaited manufacture of optical fibres with low loss, wich can now be as

low as ~0.2dBkm-1 in the and 1.55μm region4, and high power lasers producing

controllable and reproducable, picosecond ultrashort pulses tunable in the anomalous

dispersion regime. ~10ps pulses at ~1.5μm from an F-centre NaCl colour centre

source were used. The results were explained5,6,7 as an exact or over balence between

Self Phase Modulation (SPM)8 and negative Group Velocity Dispersion9. Group

Velocity Dispersion (GVD) arising from the second order derivative of the

propagation constant in the fibre, vanishes to zero around 1.3μm, becomes negative

above and positive below for standard single-mode fibre (see section 1.3). Propagation

of low intensity (linear regime) pulses in a medium with GVD, regardless of sign,

results in temporal broadening. Whilst broadening, the wavelengths contained in the

pulse redistribute themselves in such a way as to produce an approximately linear

change in frequency across the pulse, or a 'chirp'.

High power pulses (nonlinear regime) in optical fibres induce an intensity

dependent change in the local refractive index across the pulse, through a process

know as Self Phase Modulation (SPM) This process, independent of dispersion,

12
induces an approximately linear chirp across the centre of a pulse. In the positive GVD

regime, the two chirping mechanisms (positive GVD and SPM) work together and the

result is enhanced temporal broadening. In the negative GVD regime however, the

processes oppose one another and may balance. Propagation can then occur without

noticeable dispersion, if the medium loss is ignored.

The SPM phenomena described above is power dependent. For a particular

range of powers around a critical value, an exact balance can be struck between

dispersion (if negative) and nonlinearity. A pulse of this nature propagates without

change in shape and is known as the fundamental (or N=1) soliton. At higher powers,

an overbalance occurs in favour of the nonlinearity, and the pulse shape undergoes a

series of continuous temporal and spectral reshapings over the soliton period.

Traditionally pulses whose temporal duration was limited only by their spectral

bandwidth, known as transform-limited pulses, have been used to excite solitons, since

this results in the lowest required energy and most predictable behaviour. In this

thesis, a variety of sources providing non transform-limited (chirped) pulses and

pulses consisting of noise are used to illustrate the universality of the steady state

soliton solution.

Fundamental solitons propagate without change in shape only in a lossless

medium. In the presence of loss, the energy of a soliton is dissipated. It has been

shown 10,10a,10b,10c,10d that in the presence of a small loss, treated as a perturbation, the

duration of the pulse broadens adiabatically to preserve its soliton nature.

In the presence of gain as opposed to loss, the reverse applies, hence optical

pulse reshaping can be achieved through amplification. Two techniques for optical

amplification are discussed in this thesis: doped fibre and Raman amplification. Over

many thousands of kilometres, using Raman gain to compensate for loss, distortionless

transmission has been demonstrated over 6000km using 50ps pulses11 (more recently

13
100km transmission has been achieved11a). By exceeding optical loss with gain,

soliton compression has been achieved, in which the pulse duration shortens in order

to maintain fundamental soliton propagation as the total pulse energy increases7.

All of the properties desribed above refer to the propagation of bright envelope

solitons derived from pulses which would normally disperse under linear conditions.

The durations are less than ~100ps, in order for dispersive broadening to become

noticeable over the typical fibre lengths used. An interesting mode of propagation

exists however if high power continuous signals or very long (>50ps) quasi-

continuous pulses are used. It has been shown that these waves are unstable, and

exhibit Modulational Instability (MI)12,13,14. This refers to the growth of any

modulation in amplitude or phase on a wave, leading to a breakup of the CW state into

a train of solitary waves. The process results in the growth of an amplitude modulation

of a short period (~1ps), accompanied by the characteristic development of spectral

sidebands. The modulation grows at a period characteristic of the interaction between

nonlinearity and dispersion, and is ultimately limited by higher order dispersive terms.

In this thesis, techniques for generating pulses at up to 2THz repetition rate are

demonstrated using the process of MI.

Self Phase Modulation (SPM) of an intense optical pulse results in a phase

change which is proportional to the local pulse intensity. The central portion therefore

undergoes a larger phase change than the wings. Hence by incorporating nonlinearity

into a fibre-based interferometer it is possible to produce intensity discrimination,

reshaping and switching of optical pulses15. A device called a Nonlinear Optical Loop

Mirror (NOLM) has been proposed as a means of switching optical pulses, since it

exhibits a transmissivity which is intensity dependent. In this thesis, such a device is

used to 'clean up' the pedestals on noisy pulses through its intensity dependent

transmission.

Optical fibres confine intense optical radiation within a core which is typically

14
less than 100μm in area, over extremely long lengths, they are ideal media for

investigating and exploiting many nonlinear effects occurring at relatively low average

powers. Many of those observed so far include: Self Phase Modulation (SPM)8,

Stimulated Raman Scattering (SRS)16, Stimulated Brillouin Scattering (SBS)17, Four

Wave Mixing (FWM)18, phase conjugation19, induced gratings (photosensitivity)20,

intensity dependent polarisation rotation (Kerr effect)21, Second Harmonic Generation

(SHG)22, and third harmonic generation23.

All of these effects need to be carefully controlled in any soliton transmission

system, to avoid effects such as cross-talk between channels in a multiplexed system

for example. Some of the consequences are exploited in this thesis, to provide tunable

sources of radiation and to study coupling between signals. All of the relevant physical

processes involved are introduced in more detail in the next section.

1.3 Dispersion in Optical Fibres

Standard step index optical fibres consist of a core of transparent optical

material such as silica of very high purity, surrounded by a cladding layer of similar

material but with a very slightly lower refractive index. Fig 1.3.1. shows a schematic

illustration of a standard, step index fibre, showing the step-like refractive index

profile below as a function of radial distance from the core.

An optical fibre is generally characterised by two parameters, the difference in


refractive index from core to cladding Δn=n1-n2, and the core radius, a. The overall

fibre radius, b, is much larger (typically 60μm) than the core, and has little effect on

propagation within the limits considered here. The core diameter 2a (typically 4-

10μm) determines the number of transverse modes which can propagate. Larger core
diameter fibres can support more than one transverse mode (multimode) but are not

considered here.

15
Fig.1.3.1 Schematic of the cross-section and refractive index profile of a step-index
fibre.

The number of transverse modes allowed to propagate is characterised by a

normalised mode frequency parameter V defined as:

V = 2π a (n21 -n22 ) ≈ 2π a (2Δn)


λ λ
(1.3.1)

where λ is the guided light wavelength, and Δn=n1-n2. It can be shown that for

V<2.405, only a single transverse mode is allowed to propagate. Higher order modes
all have a different propagation constant, resulting in substantial differences in flight
times if a pulse is transmitted in a multimode fibre. This results in a large intermodal
dispersion occurring in multimode fibres, however only single mode fibres are
considered here.

The transverse area occupied by the confined fundamental mode in a single


mode fibre is a very important parameter, as it determines the intensity of the guided
wave, and hence the strength of any nonlinear effects. This mode field area or
effective core area is different to the core area through the dependence of V on λ in
Eq(1.3.1). For a given core radius, a, and refractive index difference, Δn, a wavelength

16
exists below which the fibre will become multimode, known as the cut-off
wavelength, λc. Above this wavelength, the beam waist becomes progressively larger,

as more of the field is confined as an evanescent wave at the core-cladding interface.


This results in a lower intensity in the core, and at very long wavelengths (>1.6μm)
eventually results in large optical losses. This dependence of the ratio of the beam
waist w to core radius a, on the V-number and hence wavelength λ is plotted
approximately in Fig 1.3.2.

Fig.1.3.2 Variation of mode-field width parameter w with V-number for a fibre of


core radius a

Typically a fibre with Δn=0.017, a=2.5μm, and a λc=1.2μm single mode cutoff

wavelength will from Eq(1.3.1) have a V-number of V=1.9. From Fig.1.3.2., the ratio
of the beam waist to the core diameter a for V=1.9 is w/a=1.3. Hence the mode-field
area is 33μm2, which is substantially larger than the core area, which is 20μm2.
Throughout this thesis, the mode-field area is quoted where relevant, as many of the
calculations are intensity dependent.

17
The major cause of pulse spreading from dispersion in single mode fibres
arises from the finite bandwidth of the signal source. The group delay for each
component in the wave packet is dependent on the wavelength through both material
U

dispersion, which dominates over the slowly varying waveguide dispersion.


U U U

Waveguide dispersion can be visualised as the small change in the group velocity
which occurs as more of the mode field lies in the cladding rather than the core of the
fibre, as the signal wavelength deviates from the single-mode cutoff value.

By neglecting the transverse light-guiding influence of the fibre, the refractive


index of the fibre can be considered spatially uniform, and the mode considered a z-
directed plane wave. This simplifies most of the nonlinear effects considered here to
simple time and distance dimensions. In a structure such as silica, an electromagnetic
wave is influenced by the molecular structure in such a way as to induce a dependence
of the group velocity on the wavelength through a dispersion relation. In this way
material chromatic dispersion in bulk silica is the dominant pulse spreading
mechanism in an optical fibre. Fundamentally the origin of chromatic dispersion is the
presence of optical resonances at which the material absorbs light through bound
electronic resonant transitions. Of course for transmission, only transparent
wavelengths are used, and for silica fibres this region is from about 600nm to 1.6μm.
The chromatic dispersion curve for bulk silica34 is shown in Fig.1.3.3.

The propagation constant β at optical frequency ω is given by:

βω =nω ω
c (1.3.2)

where n(ω) is the refractive index at frequency ω, and c is the vacuum speed of light.
The dependence of the propagation β on the wavelength λ is characterised by a series
of differentials:

18
m
β m= d β where m=0,1,2...
dωm
(1.3.3)

Fig.1.3.3. Variation of chromatic dispersion parameter β2 with wavelength for fused


silica. The value of β2 vanishes around 1.27μm, and becomes negative above this
wavelength.

It is β2 which is plotted in Fig.1.3.3., in units of ps2km-1, over the region of


lowest loss from ~600nm to 1.6μm. β1 is related to the group velocity Vg of a pulse by

the relation:

β1= 1
Vg (1.3.4)

β2 is given by:

β 2 = d 1 = -1 dVg
dω Vg Vg dω (1.3.5)

19
and is therefore a measure of the dependence of the group velocity on frequency or
wavelength. From Fig.1.3.3. it can be seen that GVD is positive below and negative
above 1.27μm, at which wavelength it vanishes to zero. This is know as the zero GVD
wavelength λ0. However total dispersion does not vanish altogether. The total can be
expanded as a Taylor series around a central frequency ω0 as follows:

β ω = β 0 +β 1 ω-ω0 +1 β 2 ω-ω0 2 +....


2 (1.3.6)

Pulse broadening calculations often involve expansion of the (slowly varying)


propagation constant β(ω). The lowest order term to contribute to pulse broadening is
the second derivative, given by β2. The expansion needs to include within β(ω) the

contributions from both material and waveguide dispersion in order to characterise


temporal dispersive broadening in a fibre. Material dispersion can be measured in bulk
silica, and waveguide dispersion can be included by adding a weighting factor which
depends on the material properties of the cladding, and characteristics such as the the
mode field diameter. However, the waveguide dispersion is generally negligible, and
need only be included in propagation calculations near λo, where β2 vanishes to zero.
Its effect is to shift λo toward longer wavelengths. For standard telecommunications
grade fibre, λo=1.31μm, but Dispersion-Shifted Fibre (DSF)24 can be manufactured
with λo>1.5μm by adjusting waveguide parameters such as the core diameter, Δn, or

the exact shape of the refractive index profile function with radial distance given in
Fig.1.3.1.

Fig.1.3.4. shows the total calculated chromatic dispersion of a typical standard


U U

single mode fibre with λo=1.31μm, on the same axes as the measured group delay

function.

20
Fig.1.3.4. Measured group delay and computed total dispersion parameter D(ps nm-1
km-1) vs wavelength for a standard telecommunications grade fibre.(Courtesy BTRL).

The group delay function is measurable, from which the total dispersion can be
approximately deduced using Eq(1.3.5). Note that although β2 is expressed in units of

(ps2km-2), a more commonly used expression is the dispersion parameter D, which is


related to β2 by the relation:

D = -2πc β 2 (ps nm-1 km-1 )


2
λ (1.3.7)

It is this parameter (which has an opposite sign to β2),which is plotted as total

dispersion in Fig.1.3.4. The dispersion parameters are important since they quantify
the amount of broadening experienced by a pulse with a known bandwidth or duration,
and also determine how long signals at different wavelengths will overlap and
possibly interact within a fibre length, the so-called 'walk-off length'.

A length scale can be defined34 over which dispersion becomes noticeable in a


pulse. The dispersion length LD for a pulse of duration T is given by:

21
2
LD = T
β2 (1.3.8)

The envelope of a modulated light wave such as a pulse is distorted through


dispersion. If the envelope function U(x,t) is a slowly varying function of propagation
distance x and time t, (which is true for all but the very shortest pulses of only a few
femtoseconds duration), and the following coordinate system is incorporated by a
simple Galilei transformation to move in a reference frame at the average group
velocity:

T = t - β1x, Z=x (1.3.9)

then it can be shown26,27 that the electric field within the fibre can be reduced to a
radial part and a longitudinal part. The radial part is neglected25 by assuming it does
not change with propagation in Z, and the following equation for the evolution in Z is
obtained:

β 2 Ž2 U
- i ŽU + =0
ŽZ 2 ŽT2
(1.3.10)

This equation shows how the pulse shape U(Z,T) distorts with propagation distance Z.
Since ∂2U/∂T2 becomes arbitrarily large for shorter pulses, the envelope distortion is
exaggerated for shorter pulses.

For an initially Gaussian pulse shape with (1/e intensity point) duration τ0 the

initial envelope follows:

U Z=0,T = exp - T22


2τ0 (1.3.11)

it can be shown25 that the solution of Eq(1.3.10) yields an output pulse of the form:

22
-1 - T2
iβ 2 Z 2
U Z,T = 1+ exp iβ 2 Z
τ20 2τ20 1+
τ20 (1.3.12)

From this result it can be seen that the pulse envelope broadens according to
the following well know formula28 for the output pulse duration τ1:

2
iβ 2 Z 2
τ1 = τ0 1+ = τ0 1+ LZ
τ20 D
(1.3.13)

The dispersion length parameter LD therefore gives a measure of the effect of


dispersion. Note that the amount of broadening is independent of the sign of β2, and
that if β2=0, the pulse does not appear to broaden at all. However around λ0, higher

order dispersion terms such as ∂3β/∂ω3 will begin to dominate in Eq(1.3.6) and would
need to be considered in propagation as they too result in pulse distortion. Pulse
propagation around λo is very important in linear communications, and provides the

highest bit rate transmission provided no nonlinear effects are allowed to manifest
themselves. In fact low distortion picosecond pulse propagation has been shown in
lengths of fibre at λ0=1.32μm, with pulses of 5ps over 1km28a. At the zero dispersion

wavelength the pulsewidth limitation is given approximately by10:

1
τmax (ps) = 1.4 L 3 (km) (1.3.13a)

and hence for a 20km fibre, τmax is 3.8ps, which corresponds to a maximum

transmission rate of around 0.1Tbits/second. However due to fluctuations in the value


of λo along the fibre owing to manufacturing imperfections, only one tenth of this

capacity is experimentally obtainable, resulting in realizable transmission rates of


~10Gbits/second over 20km.

23
Eq(1.3.13) can be rewritten:

U(Z,T) = u (Z,T) ei∅(Z,T) (1.3.14)

which indicates that the phase of the output pulse is evolving with a quadratic
dependence on time, and with a sign of curvature given by that of β2. For λ<λ0, i.e. a
pulse in the normal dispersion regime, β2>0 and the long wavelength (red shifted)

components travel faster along the fibre than the short (blue) components. The reverse
occurs for β2<0, in the region λ>λ0 known as the anomalous dispersion regime. The

time derivative of the instantaneous pulse phase defines a local frequency within the
pulse, which acquires a linear characteristic in the central region of the pulse known as
a frequency chirp. A chirped pulse has a longer duration than that determined by the
Fourier-transform limit of its spectral bandwidth. The chirp induced through quadratic
dispersion can be either positive or negative, depending on the sign of the dispersion
parameter D. Dispersive broadening always occurs for a finite β2, unless the intensity

of the signal pulse becomes sufficiently high that nonlinear effects start to dominate.
This is discussed in the next section.

Fig 1.3.5. Dispersive broadening of a Gaussian input pulse (shown dashed) at Z=2LD
and Z=4LD (Arbitrary scales).

24
Fig.1.3.5. shows the temporal dispersive broadening induced on an input pulse
launched at Z=0 (shown dashed), after propagating over lengths Z=2LD and Z=4LD,
where LD is the dispersion length given in Eq.(1.3.8). Clearly, dispersion places a

definite upper limit to the transmission bandwidth of an optical fibre transmission line
enabling only a small fraction of the available bandwidth around the low disperion,
low loss wavelength region in the fibre to be used. The lowest loss wavelength region
is around 1.5μm , as shown in Fig.1.3.6.

Fig.1.3.6. Recorded loss profile for a standard single mode fibre. The dashed curve
represents intrinsic loss present through Rayleigh scattering and absorption in silica.
(Courtesy BTRL).

The zero dispersion wavelength can be tailored to coincide with the lowest loss
region around 1.5μm. This then offers an available low dispersion bandwidth of
around 200nm, as shown in Fig.1.3.6. The peak around 1.4μm is due to an overtone of
the fundamental water absorption occurring at 2.73μm, and can be reduced by careful
exclusion of water in the manufacturing process. The dotted curve represents the
fundamental intrinsic loss represesented through Rayleigh scattering by random
density fluctuations fused into the silica. This loss is inversely proportional to λ4, and
is the dominant loss mechanism around 1.5μm, where all other absorption losses are
lowest. The loss at 1.5μm can be reduced to less than 0.2dB km-1, which

25
represents a 1/e absorption length of:

1 = 4.34 (km)
α α dB (1.3.15)

where αdB is the loss in dB km-1. From 1500-1700nm, the loss in silica optical fibres

results in a loss length of α−1 ~22km.

The use of dispersion flattened fibres29 manufactured with multiple cladding


layers to alter the waveguide dispersion profile substantially can reduce the average
magnitude of the GVD over a large wavelength region around λ0. A more interesting

option, however, is that presented by the influence of nonlinearity when high power
pulses are used as signals. These effects are discussed in the next section.

1.4 Optical Fibre Nonlinearity.

In the previous section, only propagation in the linear intensity regime was
considered. Waveguides such as silica optical fibres however, do not generally
respond in a linear manner to a non-zero guided intensity. A change in power at the
input may cause a smaller change in the output accompanied by a qualitatitive change
in the output such as a wavelength shift. Silica has probably the lowest coefficient of
nonlinearity of any transparent material, however the high intensity within a fibre,
coupled with the long confinement length causes an accumulation of the nonlinear
effect, and can result in observable nonlinearity at only moderately high powers. A
beam focussed to a waist of radius ω0 in a bulk media achieves a maximum intensity
I=P/πω02, where P is the focussed power. The confocal parameter28 Z0=πω02/λ

approximately defines the focussed interaction length over which the beam area
doubles from that at the waist. Since nonlinear interactions scale with power and
interaction length, it is useful to quantify the power length product for focussing in a
bulk medium as:

26
πω20 P
IZ0 = P =
πω20 λ λ (1.4.1)

hence IZ0 is independent of focussing. However in dielectric waveguides such as


optical fibres, ω0 is effectively equivalent to the core radius, (adjusted to the mode
field radius) and Z0 is the fibre length before loss becomes significant. Since the fibre

loss length is typically α-1=20km at 1.55μm, the enhancement in the power length
product becomes:

IZ0 (fibre)
≈ πωP
2
1 λ= λ
α P πω20 α
IZ0 (bulk) 0 (1.4.2)

At 1.55μm for ω0=6μm which is a typical mode radius, this enhancement is

~300x106. Despite the low nonlinear coefficient, this enhancement can result in
significant nonlinearity occurring over hundreds of metres of fibre with only a few
milliwatts of peak signal within the fibre. This nonlinearity can have drastic effects on
ultrashort pulse propagation in fibres, and is the major cause of pulse broadening in
linear (normally dispersive) transmission systems. The fundamental nonlinearity to be
included in the evolution equation describing the pulse envelope in a weakly nonlinear
waveguide can be expanded in terms of the electric field E within the pulse. The
induced nonlinear polarisation within the dielectric waveguide PNL can be described

by:

PNL= ε0 χ 1 E+χ 2 EE+χ 3 EEE+... (1.4.4)

where ε0 is the vacuum permittivity and χ(j) (j=1,2,3...) is the jth order susceptibility.

The dominant term in Eq(1.4.3) is the linear term χ(1), from which the
absorption coefficient α and the refractive index n can be deduced. χ(2) is responsible
for nonlinear effects such as Second Harmonic Generation (SHG), and is zero in
centrosymmetric materials such as silica. Only if the symmetry of the host is broken

27
can χ(2) become finite and SHG be experienced. The lowest order nonlinear effects
observed in silica optical fibres generally arise from the χ(3) contribution, responsible
for third harmonic generation (which requires phase matching and is generally not
observed efficiently), four wave mixing and other nonlinear four wave interactions.
Only those processes which can be phase matched are observed efficiently, and most
of the latter nonlinear interactions are normally phase matched through a nonlinear
dependence of the waveguide refractive index on intensity, called the nonlinear Kerr
effect.

Considering χ(3), and ignoring its dependence on ω (which only dominates on a


timescale comparable with the χ(3) response time which is below 10fs30,30a), Eq(1.4.3)
can be simplified to:

PNL= ε0 χ 1 + χ 3 E 2 E (1.4.4)

The intensity dependence of the overall susceptibility and induced polarisation


can clearly be seen. By relating the refractive index to the total susceptibility, it can be
shown that a refractive index exists32:

n = n0 +Δn (1.4.5)

which contains a linear part, and a nonlinear part which is dependent on the local field
intensity through:
χ 3 n0
Δn = E 2
2 (1.4.6)

In real world units

Δn = n2 I (1.4.7)

28
where I is the field intensity expressed in W m-2.

The coefficient n2 has a value 3.2x10-20 m2W-1 for silica and is called the Kerr

coefficient. This is approximately two orders of magnitude smaller than that quoted
for CS2, but the enhancement obtained in long lengths of fibre overcomes this. If two
waves copropagate, the refractive index change experienced by one, n1, is perturbed
by the intensity of the other, I2, and vice-versa, with a resultant change dependent on

the total intensity through:

Δ n1 = n2 I1 + 2I2 (1.4.8)

Although this is a relatively small change in refractive index, a significant


phase advance or retardation can result over moderate lengths of fibre. This relative
phase change is given by:

Δ∅ = Δn k L
= n2 I k L (1.4.9)

for a single wave, where k=2π/λ is the wavenumber and L is the fibre length. This
intensity dependent phase shift results in a time dependent phase shift across a pulse
envelope, and is called Self Phase Modulation (SPM) if only one pulse is involved, or
Cross Phase Modulation (XPM) if two pulses with non-overlapping spectra are
involved. A phase shift across a wave results in a frequency shift determined by the
time derivative of the phase given by:

dω t = - d d∅ t
dt
dI t
= - n2 k L
dt 1.4.10

This frequency shift is sensitive to the wave envelope intensity, and results in a

29
local change in frequency across a pulse. SPM was first observed in fibres having a
hollow core filled with CS231. Only moderate power levels are required in silica based

fibres32,33, as the fibre can be kilometres long. Based on this length dependent
nonlinearity, it is useful to define a nonlinear length LNL(c.f. dispersion length LD in

Eq(1.3.8)) over which the induced phase change becomes comparable with unity:

LNL = 1 = 1
γP0 n2 kI (1.4.11)

where P0 is the peak power of the signal and γ is related to the fibre parameters as

follows:

2πn2
γ=
λAeff (1.4.12)

where Aeff is the mode field area (sometimes called the effective core area). If the

effect of SPM is included in the evolution equation for the pulse envelope propagation
(1.3.10), the result is called the Nonlinear Schroedinger Equation (NLS):

ŽU β 2 Ž2 U
i - + γ U 2U = 0
ŽZ 2 ŽT2
(1.4.13)

Eq(1.4.13) can be compared directly23 with the well known Schroedinger


equation if the coordinate transformation of Eq(1.3.9) is taken into account:

ŽΨ Ž2 Ψ
i + + VΨ = 0
Žt Žx2 (1.4.14)

Clearly the potential V in the Schroedinger equation is analogous to ⎥U⎥2 in the


NLS. This indicates that as the intensity of a pulse increases, the potential in the NLS
deepens or shallows depending on the sign of β2. Negative β2 results in a trapping of

the wave represented by U, which would otherwise spread due to dispersion. If a

30
balance can be struck, a stationary envelope function can be generated, and this
situation is considered later.

If dispersion is ignored by setting β2=0, the evolution equation simplifies to:

ŽU
i + γ U 2U = 0
ŽZ (1.4.15)

The envelope solution of this equation does not undergo any change in shape,
and for an initial pulse envelope function U(Z=0,T)=U0(T) the output pulse can be

shown to be25:

U(Z,T) = U0 (T) exp i n2 kZ U0 (T) 2 (1.4.16)

The instantaneous phase within the pulse evolves according to Eq(1.4.9),


operating on the local intensity across the pulse envelope. The calculated effect of
SPM on the local frequency of a pulse is illustrated in Fig.1.4.1

Fig.1.4.1. Temporal dependence of the local phase change dφ, and frequency shift dω
through SPM for a Gaussian input pulse envelope.

Since n2kZI=γPZ, the effect scales with length and peak power, and has a
characteristic lengthscale LNL. In Fig.1.4.1. the normalisation γP=1 is made and the
propagation length is LNL. The leading edge of the pulse undergoes a frequency

31
downshift (or red shift) and the trailing edge an upshift (or blue shift). Over the central
region of the pulse this leads to an approximately linear frequency 'chirp' on the pulse,
with the largest shift obtained at the inflexion points of the envelope profile. This
chirp is very sensitive to pulse shape, but is qualitatively similar for all "bell" shaped
pulses. The maximum frequency shift Δωmax can be obtained from the maximum phase
shift ΔΦmax from Eq(1.4.9) to obtain:

Δωmax = f Δ
∅max
T (1.4.17)

where f=0.86 for a Gaussian pulse shape34. For an unchirped (transform limited) input
pulse the total output bandwidth is the sum of the initial, plus the change in
bandwidth. It can be shown 34 that the total bandwidth due to SPM is:

Δωmax ≈ 2.8 Δ ∅max+1


τ (1.4.18)

where τ is the FWHM pulsewidth. Therefore SPM spectral broadening depends on


power, length and pulse duration.

A theoretical power spectrum35 calculated for a 10W peak power pulse of 30ps
duration at a 1μm wavelength launched into 500m of fibre with a 100μm2 mode field
area is shown in Fig.1.4.2.

Fig.1.4.2. Theoretical power spectrum through SPM of a 30ps, 10W pulse in a 500m
fibre35.

32
It is clear from Fig.1.4.1. that each particular frequency shift occurs for two
temporal regions within the pulse envelope. Interference in the power spectrum can
occur resulting in the characteristic spectral modulation shown in Fig.1.4.2. If
dispersion is now included in Eq(1.4.13), then over lengths where LD and LNL become
comparable, dispersion operates on the SPM broadened spectrum. In the β2>0 (normal

dispersion) regime, the result is enhanced temporal broadening as dispersion operates


on the new frequency components of the spectrum. A typical experimental spectrum
corresponding to this situation is shown in Fig.1.4.3.

Fig.1.4.3. Experimentally obtained spectrum of SPM using 100ps, 90W signal pulses
in a ~2.2km length of fibre. Note spectral sidelobes are due to optical wave breaking
(see text).

The spectral resolution was insufficient to show spectral modulation from


interference, however new structure became apparent. Dispersion induces a process
called optical wave breaking36. This is seen as two sidelobes on the SPM spectrum in
Fig.1.4.3., and is due to the extremities of the pulse having an opposite sign of chirp to
the centre, as shown in Fig.1.4.1. These extremities have a negative chirp gradient, and
therefore dispersion operates so as to compress the energy in these wings, which
eventually collapse into a temporal singularity and overlap different parts of the
spectrum. Whilst overlapping these frequency components generate the sidelobes
through frequency mixing. The process is analogous to an ocean wave close to the
shore, whose velocity is intensity dependent, hence the name.

33
Temporal broadening through dispersion of a SPM spectrally broadened pulse
can be several orders of magnitude greater than that caused by dispersion alone37.
However the approximately linear chirp which initially only covers about 1/3 of the
pulse through SPM alone can be extended to cover almost all of the pulse through
dispersion. This is accompanied by a "squaring off" of the pulse shape, making the
broad pulse more suitable for high efficiency compression using a negatively
dispersive delay line such as a grating pair38,39. In fibre grating compression, negative
dispersion from a grating pair is used to reverse the effect of normal (β2>0) dispersion

in the fibre. Operating on a pulse whose spectrum has been broadened substantially
through SPM in a fibre, the pulse can be made to compress as the new frequency
components undergo different flight times in the delay line and eventually overlap
each other temporally. The phase delay function of a pair of diffraction gratings
arranged with parallel faces is given by:

∅ω = ∅0 - aω2
(1.4.19)
where a is a constant of the grating compressor related to the grating separation. Since
the frequency chirp is approximately linear across the whole pulse when significant
dispersion accompanies nonlinear SPM in the fibre, the phase must be approximately
quadratic. A quadratic compressor with a phase delay similar to that of Eq(1.4.19) is
capable of almost completely compensating for this chirp, and highly efficient
compression is possible. The output pulse duration is of the order of that given by the
Fourier transform limit of the SPM spectral bandwidth, however, higher order terms in
the chirp also require dispersion compensation , and hence perfect transform limited
pulses are difficult to generate.

Fig.1.4.4. shows the experimental schematic of a fibre grating compressor of


the type used for work described in this thesis. Using pulses of 100ps initial duration,
with peak powers around 50W at 1.3μm, pulses as short as 2ps can be generated with

34
up to 1.5kW of output peak power. This source is very useful for soliton studies, and a
full characterisation of the technique can be found in the references42,42a.

Fig.1.4.4. Scematic of the fibre grating compressor used as a signal source in this
thesis42.

Since β2 is negative above the zero dispersion wavelength λ0, the fibre itself

can behave in a manner exactly analogous to a distributed negatively dispersive delay


line. In the anomalous dispersion regime (above ~1.3μm in a standard fibre) it follows
that the potential analogue in the Schroedinger Equation (1.4.14) becomes deeper and
deeper as the intensity ⎥U⎥2 becomes larger. This has the effect of trapping the energy
within a wave temporally. The distributed negative dispersion operates with SPM
concurrently, compensating exactly for the tendency of the pulse to broaden provided
the power is high enough. Physically, SPM red-shifts the leading pulse edge and blue
shifts the trailing edge. In a negatively dispersive fibre, however, the blue components
can begin to catch up with the red components, and at some particular power level a
balance can be struck between the SPM and the negative dispersion. At this point the
pulse shape is retained as it propagates, and the pulse is a fundamental soliton.

Neglecting fibre loss initially, Eq(1.4.13) (NLS) correctly describes the


envelope pulse evolution in the presence of dispersion β2:

ŽU β 2 Ž2 U
i - + γ U 2U= 0
ŽZ 2 ŽT2
(1.4.20)

35
In order to reduce the units of the equation to those of the Schroedinger
equation, it is useful to impose the following normalisation:

A= U , ξ= Z , τ= T
P0 LD T0 (1.4.21)

where P0 is the peak power in the fibre, LD is the previously defined dispersion length
and T0 is the duration of the input pulse. This yields:

ŽA sgn β2 Ž2 A
i = - N2 A 2 A
Žξ 2 Žτ 2
(1.4.22)

where sgn(β2) is the sign of β2, i.e -1 for anomalous dispersion. The parameter N is

given by:

N2 = LD
LNL (1.4.23)

By defining :

1
LD γτ20 2
u = NA = A= A
LNL β2 (1.4.24)

equation (1.4.22) can now be simplified to yield the well known standard NLS:

2
1 Ž u + i Žu + u 2 u = 0
2 Žτ2 Žξ
(1.4.25)

For anomalous dispersion, this equation can be solved using the inverse
scattering formalism, to yield solitary solutions for input pulses of hyperbolic secant
shape. The simplest solution is the fundamental soliton, whilst further solutions exist
which are integer multiples in amplitude of the fundamental. The fundamental soliton

36
has a solitary or stationary envelope function which is retained during propagation,
whilst higher order solitons exist which are periodic with propagation distance. The
fundamental soliton solution is given by:


u(ξ, τ) = sech(τ) exp
2 (1.4.26)
The importance of this function for pulse propagation lies in the independence
of the pulse envelope sech(t) on propagation distance parameter ξ. Hence if a
hyperbolic secant shaped pulse of peak power sufficient to make N=1 in Eq(1.4.23) is
propagated in a lossless fibre, the pulse will propagate without change in shape over
arbitrarily large distances. It is also important to note that the soliton phase is
independent of local pulse intensity, and thefore no frequency chirp exists (the
fundamental soliton is perfectly transform limited). The phase rotates at a linear rate
equally across the entire pulse. This makes solitons perfect for switching based on
interferometers, as described in chapter 7.

For optical communications, the soliton is ideal, since provided the power is
maintained, dispersion does not dominate transmission. From Eq(1.4.23) the peak
power required to excite the fundamental soliton is given by:

LD
=1
LNL (1.4.27)

and hence the fundamental soliton power is given by:

β2 3.11 β 2
P1 = =
γτ20 γτ2FWHM (1.4.28)

At this peak power level, the L.H.S of Eq(1.4.22) is zero, indicating that the
pulse envelope is stationary. In measurable units,

37
3
3.11 Dλ Aeff
P1 =
4π 2 c n2 τ2FWHM (1.4.29)
where D is the GVD in units of ps nm-1km-1, and the full width half maximum pulse
duration τFWHM =1.763T0, where T0 is the half width (at 1/e fold intensity) of the
pulse. For power levels below P1, the dispersion in the fibre dominates over the

nonlinearity, and broadening always occurs, albeit at a lesser rate if the power is
comparable with P1.

To excite a higher order soliton, the power has to exceed P1 by a factor N2,

where N is the soliton number. i.e. for an Nth order soliton:

PN = N 2 P1 (1.4.30)
When a higher order soliton is excited, the nonlinearity exceeds that required
to balance the dispersion and maintain the pulse shape, and the pulse shape undergoes
compressions and rarefactions at a periodic interval related to the rate of phase
rotation in Eq (1.4.26). The period in units of fibre length is given by5

Z0 = 0.322π cτ = 0.322π τ
2 2 2
2 2
λ D 2 β2
(1.4.32)

If Aeff=80μm2, a fibre with ⎥D⎥= 10ps nm-1km-1 at 1.4μm would require 18W

of peak power to excite a fundamental soliton of 10ps duration. The soliton period for
this 10ps pulse would be 48.6m of fibre, increasing to 4.86km for an input pulse of
100ps duration with adequate peak power. The soliton period is related to the
dispersion length LD by:

Z0 = π LD
2 (1.4.32)

The numerical evolution of the envelope of an n=1,2 and 3 soliton are shown
in Fig.1.4.5., for a propagation length normalised to the soliton period Z0.

38
Fig 1.4.5. Theoretical evolution of the fundamental (N=1) and high order (N=2,3)
solitons with normalised propagation distance Z/Z0 .Note the change in intensity scale
between solutions.

The first observations of soliton propagation5 were undertaken using


transform-limited pulses from a colour centre laser operating around 1.55μm. Higher
order solitons43 were also observed in a similar manner. If the peak power or pulse
shape launched is not exactly matched to the hyperbolic secant solution, it has been
shown theoretically44 that the pulse width and shape adjust accordingly to fit the
soliton shape of appropriate or lower energy, with excess energy shed as a dispersive
pedestal wave. The adjustment normally occurs before Z/Zo=10/π, which is roughly

three soliton periods. This is important when considering propagation experiments


where the pulse intensity envelope is interrogated. Hence any reasonably shaped pulse
can evolve into a soliton, as the solitary wave is the asymptotic solution to a variety of
input waves. Results are presented in this thesis which show amplified noise-bursts,

39
CW signals under modulated gain, amplified modulational instabilities (self-pulsing
instabilities), and amplified sub-fundamental soliton power pulses all evolving into
soliton pulses. This is of major importance for regenerative amplification operating
periodically in an optical communication system, as the regenerated pulses may not
be exact soliton solutions. The fact that noise can develop into pulses is of great
importance, and requires strict control.

High order solitons have been shown to be unstable in optical fibre


transmission, through fragmentation from a process called the Soliton Self-Frequency
Shift (SSFS). This subject is treated in the next section with Stimulated Raman
Scattering (SRS), which concentrates on how nonlinear scattering processes can be
used to amplify pulses which have become attenuated through fibre loss.

So far loss has been neglected. If the fibre length is comparable with the loss
length α-1, which is typically 20km at the lowest loss wavelength around 1.5μm, the
pulse energy will experience significant attenuation through absorption, and loss must
be considered. Loss can be treated as a small perturbation provide it is small over the
evolution length of the soliton defined by the soliton period. This limit is said to be
adiabatic provided the perturbation length is longer than the soliton period, and hence
loss can be incorporated into the NLS in terms of the normalised amplitude in
Eq(1.4.25) to yield:

Žu Ž2 u
i +1 + u 2 u = -iΓu
Žξ 2 Žτ 2
(1.4.33)

where Γ=αLD/2. The term on the R.H.S. is effectively an inverse driving term in the

NLS, and can be treated as a weak perturbation44 to yield a solution for an input pulse
of the form u(0,τ)=sech(τ). The solution undergoes an increase in pulsewidth in the
presence of loss, governed by the ratio of output to input pulsewidth:

40
τ1 = e 2Γξ = e αZ
τ0 (1.4.34)
This indicates that as the propagation length increases, the amplitude of the
soliton will decrease with a subsequent increase in duration in the presence of loss.
Both N=1 and N=2 soliton propagation have been modelled in the presence of loss45,46,
with the conclusion that dispersive broadening occuring on a soliton is less than that
experienced by a non-soliton pulse. Many schemes have been proposed for amplifying
solitons to overcome absorption loss and maintain the signal duration47,48,49.
Stimulated Raman Scattering (SRS) has been demonstrated as a distributed amplifier
in a long distance transmission experiment over many thousands of kilometres49a,50,51.
Erbium doped fibre amplifiers have also been employed in transmission experiments
over 10,000km52.

In chapter 4 of this thesis the first experimental observation of femtosecond


soliton pulse amplification in Erbium doped fibre is presented, demonstrating that the
bandwidth of the system is large enough to support ~200fsec pulses. SRS
amplification in soliton transmission studies has largely been superceded by Erbium
amplifiers, as the lumped gain capability of Erbium has been shown to be more
effective than the distributed Raman gain, which needs to be distributed to allow
sufficient interaction length between pump and signal for energy transfer to take place.
The gain per pump photon is also more efficient in Erbium, with up to 3.3dB/W gain
per unit pump power achieved53, and a variety of efficient pump wavelengths
available.

Raman amplification is utilised throughout chapter 6 to amplify a variety of


signals up to soliton powers. Intra pulse Raman scattering occurs, causing the Soliton
Self-Frequency Shift, resulting in a frequency pulling effect to longer wavelengths.
This limits femtosecond soliton amplification, and results in pulse broadening, since at
higher wavelengths, the dispersion is higher, and hence the fundamental soliton power
increases. However the effect can be used as a tunable source of femtosecond soliton

41
pulses in either single-pass or oscillator arrangements54, although the pulses do have
a significant temporal jitter. Details of the SRS process and Raman amplification are
given in the next section.

1.5 Stimulated Raman Scattering and Raman Amplification.

Two nonlinear inelastic scattering processes dominate propagation of moderate


to high intensity signals in silica based optical fibres, limiting the power at which
signals can be transmitted, and hence the repeater spacing. Nonlinear transmission
involves high intensity pulsed radiation in the form of solitons, and hence extensive
interest is shown in investigating these scattering processes in order to suppress,
manipulate or exploit them.

Both Stimulated Raman Scattering (SRS) and Stimulated Brillouin Scattering


(SBS) are inelastic processes where an initially unexcited medium scatters light to a
different wavelength. Some of the incident energy is absorbed through either
electronic or vibrational excitation of the atom or molecule repectively in SRS, or by
the formation of an acoustic lattice phonon (SBS) the characteristics of which are
determined by the bulk medium properties. Scattering in an initially unexcited
medium generates red-shifted or 'Stokes' radiation, and in an initially excited
(irradiated) medium generates blue-shifted or 'anti-Stokes' radiation.

All of the nonlinear processes discussed in this thesis are governed by the third
order nonlinear susceptibility χ(3), and involve four wave interactions. Those in the last
section left the dielectric medium unexcited, and are elastic. Those in this section
involve an energy exchange which is assumed to be very small compared with the
incident energy, and are inelastic. Both SRS and SBS were discovered early in
fibres55,56,57, and are analogous to the phenomena seen in bulk materials, except for the
exclusion of certain scatttering phenomena which cannot be phase matched in an
approximately two dimensional dielectric waveguide. SRS is caused by the interaction
of the incident field with the optical phonon modes of the medium, and SBS involves

42
the acoustic phonon modes which are much lower frequency, resulting in much
smaller frequency shifts. For both processes, phase matching is automatic for the
stimulated process which occurs through χ(3) at high intensities, since the momentum
deficit is provided by the scattering phonon involved. Although similar in origin, the
dispersion relations applying to SRS and SBS are different, resulting in SBS only
being observed in the reverse direction and only for the narrowest linewidth pump
sources58. Hence SBS is generally not observed in the forward direction in a fibre or
from pulsed sources where the pulse bandwidth is larger than the Brillouin linewidth.
For this reason SBS is not covered here, as the bandwidth of the sources used for
soliton studies (much less than 1ns pulsewidths) always resulted in suppression of the
effect.

The Raman shift in optical fibres is due to an ultra-high frequency optical


phonon mode around 13.2THz which extends as far as 30THz55. Inelastic scattering
imposes a severe limit on the power capacity per channel in an optical transmission
system, and can result in energy being transferred across channels. Spontaneous
Raman scattering (a linear scattering process) initially transfers a small amount of
energy from the incident optical field into molecular vibration of the fibre lattice
structure, resulting in a frequency downshifted (red-shifted) signal being generated.
For an intense pump signal, the effect can be stimulated through nonlinearity resulting
in exponential growth of the Stokes wave, and exponential absorption of the anti-
Stokes wave. If the generated Stokes wave copropagates with the pump, almost
complete pump conversion into the Stokes band can occur. In general, the growth of
the Stokes wave with length is given by the following relation if loss is ignored:

dIs
= g Ip Is
dZ (1.5.1)

where Ip and Is are the pump and Stokes field intensities respectively, and g is the

Raman gain coefficient. This gain is present in both directions relative to the pump.
The Raman gain coefficient, g, is shown as a function of frequency detuning from the

43
pump in Fig.1.5.1. for silica based optical fibres55. The gain coefficient is maximum
for a shift of approximately 440cm-1, and scales inversely with excitation wavelength
(1.32μm for this graph).

Fig.1.5.1. Raman gain coefficient, g, vs frequency shift in silica fibres. Vertical scale
shows magnitude for excitation at 1.32μm.

The lineshape is very broad due to the inhomogenous crystal field of the glassy
fibre, resulting in a spread of the molecular vibrational frequencies into a continuum.
The inclusion of codopants in the fibre core can modify the Raman gain curve, but
dopants used in manufacture have no significant qualitative effect.

As seen in Eq(1.5.1), weak signals within the Stokes band which copropagate
with the pump experience stimulated amplification. This means the process can be
used as a broad-band amplifier , and to produce tunable sources in either single pass or
Raman oscillator arrangements. The amplification is given by:

gI0Z
Is Z = Is 0 e
(1.5.2)

where Is(0) is the Stokes intensity at Z=0, and Z is the interaction length. In practice, if

no signal is injected at the Stokes wavelength, SRS builds up from spontaneous


Raman noise. Eq(1.5.2) shows that distributed amplification of a signal injected
simultaneously at the Stokes wavelength will occur, provided it overlaps with the

44
pump (if the pump is pulsed as is often the case), and that it exceeds the Raman noise
intensity, which will otherwise dominate the process. The FWHM bandwidth of the
Raman gain curve shown in Fig.1.5.1 is in excess of 8THz, and can hence amplify
pulses as short as 100fs in the low loss wavelength region. Once stimulated parametric
Raman scattering occurs in a fibre, a large optical phonon density makes anti-Stokes
scattering possible, from molecules vibrationally excited as a consequence of Stokes
SRS. It evolves according to the following equation:

-gI0Z
Ias Z = Ias 0 e
(1.5.3)

and by analogy with Eq(1.5.2) the anti-Stokes signal intensity Ias(0) is absorbed

exponentially into the pump intensity. Effectively the anti-Stokes becomes a new
pump and the old pump becomes the new Stokes band in this interaction. It then
follows that with sufficient pump intensity the process can cascade to produce
multiple orders of Stokes bands.

Although the SRS process does not have a fixed threshold, a pump power
threshold has been defined57 as that power Pth at which the Stokes band becomes equal

in intensity to the pump:

kAeff
P th ≈
gL (1.5.4)

where Aeff is the mode field area of the fibre, and k is a fibre constant in the region 16-

20. Pump depletion is ignored in this estimate, and so is fibre loss. The value of k is
~16 in the forward and ~20 in the reverse direction to the pump field, and hence
forward Stokes grows first unless the backward Stokes is pre-seeded at the Stokes
wavelength.

For non-polarisation preserving fibre, k is double this value. Pth is

approximately 0.5W Continuous Wave (CW) for a fibre of ~20km, which is the loss

45
length. For pulsed excitation, this peak power is easily obtained and the process is
highly efficent, but for ultrashort pulses dispersion quickly separates the Stokes signal
from the pump through 'walk-off', and the Stokes receives no more gain. The
interaction length Z in Eq(1.5.2) needs to be replaced by the walk-off length ΔL given
by:

ΔL = Δ t
DΔλ (1.5.4)

where Δt is the pump pulse duration, Δλ is the pump-Stokes wavelength separation,


and D is the average fibre dispersion in ps nm-1km-1. It has been shown that for
sufficient pump power, the Stokes signal becomes maximal after 1.5 to 2 walk-off
lengths59,60 which is ~40-60m for 100ps pump pulses at 1.06μm in a fibre with D=25ps
nm-1km-1.

As a result of the high efficiency of the SRS process, a variety of schemes have
been devised to employ it as the gain medium in an oscillator arrangement61. By using
the relatively large bandwidth available to amplify ultrashort pulses, in conjunction
with a negatively dispersive delay line to overcome the effect of normal dispersion on
the Stokes signal from the fibre below λ0, short pulses have been generated directly

from synchronously pumped Raman oscillators62.

If the Stokes field wavelength falls in the anomalous dispersion regime, then
negative disperision can be provided by the fibre itself above λ0. In a variety of

configurations including single pass63,64,65,66, and oscillators66,67,68, soliton pulses as


short as 200fsec have been generated, tunable throughout the Raman band of the pump
pulses, and in some cases the higher order Stokes bands.

Modulational Instability (MI) can act as a precursor to this soliton Raman


continuum generation, which is discussed in chapter 6 of this thesis, provided the
pump wavelength is in the anomalous dispersion regime as well as the Stokes band.

46
The primary importance of soliton Raman generation is that the Raman gain provides
adiabatic amplification of the Stokes pulse, which may consist of noise spikes or MI
peaks. This process is the reverse of that described in Eq(1.4.34), where loss induced
soliton pulse broadening. If the energy of a soliton is increased by amplification, the
duration decreases adiabatically in order to maintain the fundamental soliton
condition. Higher order solitons cannot normally be formed through Raman
amplification, as the energy cannot be injected into the soliton quickly enough for
non-adiabatic compression to occur except for extremely intense pump fields. During
Raman amplification, the soliton pulse normally sees only a small increase in energy
over a gain length relative to the soliton period. This adiabatic form of amplification is
dicussed further in chapter 6, where it is utilised to generate solitons from a variety of
non-soliton signals.

The adiabatic scenario discussed so far operates in the relatively small gain
regime, with αgZ0<0.05, where αg is the gain length for the amplification process and
Z0 is the soliton period. In this thesis results are presented above this limit, where the

pulse energy undergoes dramatic changes in energy over a relatively short distance. In
chapter 6, sub-fundamental soliton power pulses are amplified through large
synchronous Raman gain to reconstruct solitons from dispersive pulses.

A severe disadvantage to SRS exists for the amplification of ultrashort pulses.


If the power spectrum of the pulse is broad enough, it may slightly overlap with its
own Raman gain band, which is finite at very small frequency shifts, as shown in
Fig.1.5.1. As a consequence71, the high frequency components of a pulse can provide
Raman gain for the lower frequency components, in a process that can build up over
long fibre lengths to a continous long wavelength shifting effect. The process is called
the Soliton Self-Frequency Shift (SSFS) and can occur without a pump signal if the
pulse is sufficiently intense. It is for this reason that multisoliton (high order soliton)
pulses generally break up in optical fibres, eventually fragmenting into fundamental
solitons which self frequency shift to various longer wavelengths and generate a

47
continuum. Experimental work to quantify the effect73 has confirmed theory74 which
predicts a shift in mean wavelength of a soliton of duration τ proportional to τ−4 or the
soliton peak power squared. As the pulse shifts to longer wavelength, the fibre
dispersion parameter increases, and the pulse broadens in order to preserve its soliton
nature. The effect is most noticeable when the input pulse spectrum exceeds ~1THz,
either through multisoliton compression which occurs as a high order soliton evolves
or through simple ultrashort soliton propagation. For this reason, solitons with τ>10ps
are normally used in soliton transmission systems, to avoid any SSFS. The SSFS
process can be used in conjunction with spectral selection to produce a source of
femtosecond solitons75, tunable by contol of pulse duration or peak power beyond
1.55μm.

Bandwidth limited gain has been shown both theoretically76 and


experimentally77 to suppress SSFS, and it has been predicted that Cross Phase
Modulation (XPM) and GVD can also have a frequency pulling effect that can trap a
soliton and suppress SSFS78, however it is hoped that more recently proposed
transmission systems incorporating high gain lumped Erbium doped fibre amplifiers
may provide suppression by bandwidth limited amplification, perhaps allowing shorter
signal pulses to be used at a higher repetition rate thus increasing the overall system
bandwidth.

1.6 References.

(1) G.B.Witham. "Linear and Nonlinear Waves", Wiley Interscience,


New York (1974)
(2) A.Hasegawa & F.D.Tappert. Appl.Phys.Lett.23,142(1973)
U U

(3) A.Hasegawa & F.D.Tappert. Appl.Phys.Lett.23,171(1973)


U U

(4) T.Miya, Y.Terunama, T.Hosaku & T.Miyashita. Elec.Lett.15,106(1979) U U

(5) L.F.Mollenauer, R.H.Stolen & J.P.Gordon. Phys.Rev.Lett.45,1095(1980) U U

(6) L.F.Mollenauer & R.H.Stolen. Fiberoptic Tech. April,193(1982) U U

(7) L.F.Mollenauer. Phil.Trans.R.Soc.London Vol A315,437(1985)


U U

(8) R.H.Stolen & C.Lin. Phys.Rev.A17,1448(1978)


U U

(9) C.Cloge. Appl.Opt. 10,2442(1971)

48
(10) A.Hasegawa & Y.Kodama. Proc.IEEE. 69,1145(1981) U U

(10a) K.J.Blow & N.J.Doran. Opt.Comm. 52,367(1985) U U

(10b) N.J.Doran & K.J.Blow. IEEE J.Quant.Elec. QE-19,1883(1983) U U

(10c) K.J.Blow & N.J.Doran. Opt.Comm. 42,403(1982) U U

(10d) D.Anderson. Opt.Comm. 48,107(1983) U U

(11) L.F.Mollenauer & K.Smith. Opt.Lett. 13,675(1988) U U

(11a) L.F.Mollenauer, M.J.Neubelt, S.G.Evangelides, J.P.Gordon, J.R.Simpson


& L.G.Cohen. CLEO CPDP17, Anaheim, CA, USA. May(1990)
(12) A.Hasegawa & W.F.Brinkman. IEEE J.Quant.Elec. QE-16,694(1980) U U

(13) A.Anderson & M.Lisak. Ot.Lett. 9,463(1984) U U

(14) B.Hermansson & D.Yevick. Opt.Comm 52,99(1984) U U

(15) K.J.Blow, N.J.Doran & B.K.Nayar. PD4,Nonlinear Guided Wave


Phenomena, Houston,Texas Feb(1989)
(16) R.H.Stolen, C.Lee & R.K.Jain. J.Opt.Soc.Am.B1,652(1984) U U

(17) E.P.Ippen & R.H.Stolen. Appl.Phys.Lett. 21,539(1972) U U

(18) R.H.Stolen & J.E. Bjorkholm. IEEE.J.Quant.Elec QE-18,1062(1982) U U

(19) E.A.Kuzin, M.P.Petrov & B.E.Davydenko.


Sov.Tech.Phys.Lett. 10,349(1984)
U U

(20) B.Kawasaki, K.O.Hill, D.C.Johnson & Y.Fuji. Opt.Lett. 3,66(1978) U U

(21) N.J.Halas, D.Krokel & D.Grischkowsky. Appl.Phys.Lett.50,886(1987) U U

(22) U.Osterberg & W.Margulis. Opt.Lett. 11,516(1986) U U

(23) J.M.Gabriagues. Opt.Lett. 8,183(1983) U U

(24) B.J.Ainslie & C.R Day. J.Lightw.Tech. LT-4,967(1986) U U

(25) K.J.Blow & N.J.Doran. Lecture notes. NATO Advanced Summer Institute
on 'Nonlinear Waves in Solid State Physics", Erice,Sicily July (1989)
(26) A.W.Snyder & J.D.Love. "Optical Waveguide Theory".
Chapman & Hall (1983)
(27) A.Hasegawa. "Optical Solitons in Fibres". 2nd Ed. Springer Verlag 1990
(28) A.Yariv. "Optical Electronics". Holt-Saunders (1985)
(28a) Bloom et al. Opt.Lett. 4,279(1979)U U

(29) D.Krokel, N.J.Halas, G.Guiliani & D.Grischkowsky.


Phys.Rev.Lett. 60,29(1988)
U U

(30) A.B.Grudinin, E.M.Dianov, D.V.Korobkin, A.M.Prokhorov, V.N.Serkin


& D.V.Kaidarov. JETP Lett. 46,222(1987) U U

(30a) R.H.Stolen, J.P.Gordon, W.J.Tomlinson & H.A.Haus.


J.Opt.Soc.Am. B6,1159(1989)
U U

(31) E.P.Ippen, C.V.Shank & T.K.Gustafson. Appl.Phys.Lett. 24,190(1974) U U

(32) F.Shimizu. Phys.Rev.Lett. 19,1097(1967) U U

(33) R.H.Stolen & C.Lin. Phys.Rev.17A,1448(1978) U U

(34) G.P.Agrawal "Nonlinear Fiber Optics". Academic Press (1989)

49
(36) W.J. Tomlinson, R.H.Stolen & A.M.Johnson. Opt.Comm.54,377(1985) U U

(37) B.P.Nelson, D.Cotter, K.J.Blow & N.J.Doran. Opt.Comm. 54,377(1985) U U

(38) D.Grischkowsky & A.C.Balant. Appl.Phys.Lett. 48,823(1986) U U

(39) A.S.L.Gomes, U.Osterberg, W.Sibbett & J.R. Taylor.


Opt.Comm. 54,377(1985)
U U

(40) E.B. Treacy. Phys.Lett. 28A,34(1968) U U

(41) E.B. Treacy. IEEE J.Quant.Elec. QE-5,454(1969) U U

(42) A.S.Gouveia-Neto, A.S.L.Gomes & J.R.Taylor.


Opt & Quant.ElecLett. 18,423(1986) U U

(42a) A.S.Gouveia-Neto. PhD Thesis. University of London (1987)


(43) R.H.Stolen, L.F. Mollenauer & W.J. Tomlinson. Opt.Lett.18,423(1986) U U

(44) J.Satsuma & N.Yajima. Prog.Theor.Phys.Suppl. 55,284(1974) U U

(45) K.J.Blow & N.J.Doran. Opt.Comm. 42,403(1982) U U

(46) K.J.Blow & N.J.Doran. Opt.Comm. 52,367(1982) U U

(47) A.Hasegawa & Y.Kodama. Opt Lett. 7,285(1982) U U

(48) Y.Kodama. &A.Hasegawa. Opt.Lett.7,339(1982) U U

(49) Y.Kodama. &A.Hasegawa. Opt.Lett.8,342(1983) U U

(50) K.Smith & L.F.Mollenauer. Opt.Lett. 14,751(1989) U U

(51) K.Smith & L.F.Mollenauer. Opt.Lett.14,1284(1989) U U

(52) L.F.Mollenauer, M.J.Neubelt, S.G.Evangelides, J.P.Gordon, J.R.Simpson


& L.G.Cohen. Paper CPD171, CLEO, Anaheim CA, May 1990
(53) M.Nakazawa. In "Optical Solitons: Theory and Experiment". Ed. J.R.Taylor.
In press, Cambridge University Press (1991)
(54) J.R.Taylor,. In "Optical Solitons: Theory and Experiment". Ed. J.R.Taylor.
In press, Cambridge University Press (1991)
(55) R.H.Stolen, E.P. Ippen & A.R. Tynes. Appl.Phys.Lett. 20,62(1972) U U

(56) E.P.Ippen & R.H.Stolen. Appl.Phys.Lett. 21,539(1972) U U

(57) R.G.Smith. Appl.Opt. 11,2489(1972)


U U

(58) D.Cotter. J.Opt.Commun. 4,10(1983) U U

(59) R.H.Stolen & A.M.Johnson. IEEE J.Quant.Elec. QE-22,2154(1986) U U

(60) A.S.L.Gomes, V.L.da Silva, & J.R. Taylor. J.Opt.Soc.Am. B5,373(1988) U U

(61) R.H.Stolen. Fiber and Integrated Optics. 3,157(1980) U U

(62) J.D.Kafka, D.F.Head, & T.Baer. Ultrafast Phenomena V. Springer series in


Chemical Physics. 46,51(1986)
U U

(63) A.S.Gouveia-Neto, A.S.L.Gomes & J.R.Taylor. Opt.Lett. 12,1035(1987) U U

(64) P.Beaud, W.Hodel, B.Zysett & H.P.Weber.


IEEE.J.Quant.Elec. QE-23, 1938(1987)
U U

(65) K.L.Vodapyonov, A.B.Grudinin, E.M.Dianov, L.A.Kulevskii,


A.M.Prokhorov & D.U.Khaidarov. Sov.J.Quant.Elec. 17,1311(1987) U U

(66) M.N.Islam, L.F.Mollenauer, R.H.Stolen, J.R.Simpson & H.T.Shan

50
Opt.Lett. 12,625(1987)
U U

(67) J.D.Kafka & T.Baer. Opt.Lett. 12,181(1987)


U U

(68) A.S.Gouveia-Neto, A.S.L.Gomes & J.R.Taylor. Elec.Lett. 23,537(1987) U U

(69) A.S.Gouveia-Neto, A.S.L.Gomes & J.R.Taylor.


IEEE.J.Quant.Elec. QE-24,332(1988)
U U

(70) A.S.Gouveia-Neto, M.E.Faldon & J.R.Taylor. Opt.Comm.69,325(1989) U U

(71) R.H.Stolen & E.P.Ippen. Appl.Phys.Lett. 22,276(1973)


U U

(72) E.M.Dianov, A.Ya Karasik, P.V.Mamyshev, A.M.Prokhorov, V.N.Serkin,


M.F.Stel'makh, & A.A.Fomichev. JETP Letts. 41,294(1985) U U

(73) F.M.Mitschke & L.F.Mollenauer. Opt.Lett. 11,659(1986)


U U

(74) J.P.Gordon. Opt.Lett 11,662(1986)


U U

(75) A.S.Gouveia-Neto, A.S.B.Sombra & J.R.Taylor. Opt.Comm. 68,139(1988) U U

(76) K.J.Blow, N.J.Doran & D.Wood. J.Opt.Soc.Am. B5,1301(1988) U U

(77) A.S.Gouveia-Neto, A.S.L.Gomes & J.R.Taylor. Opt.Lett. 14,514(1989) U U

(78) D.Schadt & B. Jaskorzynska. J.Opt.Soc.Am. B5,2374(1988) U U

51
52
2. FIBRE RAMAN LASER SOURCES

2.1 Fibre Raman Oscillators

As decribed in the last section, Stimulated Raman Scattering (SRS) represents


a major loss mechanism for intense optical signals in fibres. This is particularly true in
the case of long, high intensity pulses, as the walk-off length for a long pulse can be
hundreds of metres, with the Raman generated at the centre of the pump pulse walking
through the pump pulse and depleting it down to the zero level. Stimulated Brillouin
Scattering (SBS)1 is also important for narrow linewidth pump sources, but was
insignificant in the work described here as the pump pulse durations were always
below 0.5ns. The Raman gain band in standard silica based fibres peaks at a shift of
440cm-1, but extends as far as 1000cm-1, due to the glassy nature of the silica fibre
material. The Raman bandwidth is about two orders of magnitude larger than that
found in crystals due to a breakdown of the wavevector selection rules in the
inhomogeneous crystal field of the fibre2. The spread of molecular vibrational
frequencies results in a spectral continuum, and hence Raman gain is inhomogenously
broadened, since the various phonon frequencies are associated with different sites3.

For a pulsed pump, the forward Stokes wave generally never depletes the
pump completely, due to walk-off. However a backward Stokes signal can see
continuous gain as it walks through the pump , either pulsed or Continuous Wave
(CW). By careful choice of fibre parameters, the group velocities of the pump and
Stokes signals can be matched in the forward direction, either through group velocity
matching around the dispersion minimum wavelength, or by propagation in different
transverse modes.In this way complete pump depletion can result in theory by
allowing the Stokes signal to walk through the pump completely. The actual limit
depends on higher order Stokes generation, and is a complicated problem. Highly
efficient conversion of pump light into the Stokes band is possible, with almost
complete pump conversion in an oscillator arrangement. It is hard to obtain narrow

53
linewidth operation of a Raman oscillator due to the inhomogoneous nature of the gain
for small signals, making them ideal for pulsed amplification. However in the large
signal limit, if pump depletion occurs, the gain becomes saturated and operation at one
laser wavelength can influence the gain at another, resulting in more of a
homogenously broadened system. Thus if the gain is saturated, the whole homogenous
linewidth is available for gain.

If mirrors are coupled to the ends of a fibre the Raman amplifier can form an
oscillator, tunable by the insertion of appropriate tuning elements within the cavity.
Only enough feedback is required to maintain operation above the single pass Raman
threshold, which in a low loss region of the fibre would enable typically 90% output
coupling of the Raman signal. Multiple Stokes orders have been made to oscillate
simultaneously through the use of multiple mirrors4.

In this chapter the construction of two types of Raman oscillator is described.


The first was a CW Raman laser operating at 1.4μm, pumped by a CW Nd:YAG laser
at 1.32μm. The pump power used here was relatively low (~1W), yielding 10mW of
output power tunable over 10nm. The source provided a useful alternative to bulky
and expensive colour centre lasers operating around this wavelength, and could be
readily scaled up for higher output power. The second system was a dispersion
compensated Fibre Raman Ring Laser (FRRL), in which the linear dispersion in the
fibre used to generate the Raman signal was compensated using a negatively
dispersive delay line. The laser was synchronously pumped and harmonically mode-
locked at the repetition rate of the pump laser (100MHz). The output pulses were of
around 1.8ps duration, with up to 20mW of average power available. The system is
ideal for the investigation of soliton effects in fibres as the output wavelength at
1.4μm lies in the anomalous dispersion regime for standard fibres. The output is also
ideal for further Raman amplification as it lies at the Stokes wavelength of any other
pump source at 1.32μm. The laser was used further in chapter 6 for the study of
synchronous amplification of noise bursts into solitons.

54
2.2 A CW fibre Raman Laser at 1.41μm

CW oscillation can occur in a fibre Raman resonator which is formed


essentially by a length of fibre and two mirrors, if the round-trip gain exceeds the
resonator losses. For a typical resonator geometry the round trip loss including that in
the fibre can be as low as ~3.5dB, requiring a single pass Raman gain of little more
than X2. A threshold has been defined (see section 1.3) as the pump power level at
which the Raman Stokes in a fibre of length L grows to the same intensity as the pump
field:

kAeff
P th ≈
gL (2.2.1)

For non-polarisation preserving fibre used here, k=30 in the forward direction.
The Raman gain coefficient g=0.74x10-13mW-1 for a pump wavelength of 1.32μm. Aeff

is the fibre mode field area (typically 100μm2), L is the gain length neglecting loss and
walk-off, which did not occur as the pump in this case was CW. However at the
Stokes wavelength, which was around 1.4μm, the loss in the fibre was approximately
1.5dBkm-1, being enhanced through OH- absorption as shown in Fig.1.3.6. This loss
represented a loss length of αL~3km. The fibre used here was 9km in length,and hence

L in eq(2.2.1) should be replaced by an effective length of 3km. The fibre had a mode
field area of ~54μm2, and hence the calculated Pth for single pass Raman gain was

7.3W.

Only 1W of CW average power was available, and hence the high threshold
pump power was not attainable in a single pass configuration, although CW single
pass generation and oscillation has been reported elsewhere at other wavelengths5,6. In
order to reduce the threshold value, an oscillator was arranged so that the single pass
spontaneous Raman provided a seed for the double pass gain, reducing the threshold
accordingly. Fig.2.2.1. shows the experimental arrangement of the oscillator.

55
Fig.2.2.1. Experimental arrangement of the CW Raman oscillator.

The CW pump signal was derived from a CW Nd:YAG laser operating at


1.32μm, providing an average output power of 2.9W. The pump radiation was directed
through (BS1)(82%T at 1.32μm, 98%R at 1.4μm) and focussed into the fibre (F) via a
X10 uncoated microscope objective (L1). The 9km length of non-polarisation

preserving silica fibre (F) had an 8.3μm core diamter and a Δn of 0.0045. The output
was collected and collimated via (L2), an identical lens to (L1), and directed onto

grating (G) via mirror (M) which was highly reflective around 1.4μm. The grating (G)
was of a holographic type, with 1200 lines/mm, and was blazed for ~70% diffraction
efficiency at 1.4μm for 75o angle of incidence into the -1 diffractive order. The
diffracted signal was then directed of beansplitter (BS1) back into the input via (L1) to

form a ring cavity. Polarisation strainer (P) ensured that the Raman signal emerging
from the fibre was predominantly P-polarised, enabling the maximum gain to be
extracted from the P-polarised pump, whilst minimising the loss from the diffraction
grating. The output from the oscillator was monitored by inserting a ~2% reflecting
pellicle at the fibre output end.

The single pass output spectrum from the fibre for the maximum launched
pump power of 1.2W is shown in Fig.2.2.2(a)., with no detectable Raman signal
around 1.4μm because of the high loss in the system (~13dB total). The spectrum does
however show an interesting signal continuum around the 1.33-1.38μm wavelength
region, the origin and nature of which is unknown. It does however represent an extra
loss to the Raman process at this stage.

56
The total loss in the ring (including that from coupling, fibre absorption which
is high at 1.4μm, and diffraction) was estimated at 18dB, which is a X63 power loss.
The Raman threshold can be calculated from Eq(1.5.2) to give a Raman gain of X63 in
a single pass of the fibre

ln 63 Aeff
Pth = = 505mW
2gL (2.2.2)

To aid alignment initially a dielectric


mirror was placed instead of grating (G). CW
oscillation was obtained for 1.2W of launched
pump power. The spectrum of the oscillator is
shown in Fig.2.2.2(b). By using filters to
select the relevant parts of the spectrum, a
conversion efficiency of 75% from pump to
Stokes was measured, with a 1.7nm Stokes
bandwidth generated around 1.41μm. Under
these conditions, the threshold power was
measured to be 500mW, which was in good
agreement with that estimated. The grating
was reintroduced to allow tunability and a
tuning range from 1.404 to 1.412μm was
obtained with a 0.4nm Raman linewidth. A
typical output spectrum from the tunable laser
is shown in Fig.2.2.3.

Fig.2.2.2. Fibre spectrum for 1.2W


of pump (a) single pass and (b) ring The tuning range could be extended to
oscillator geometry.
cover 2nm of bandwidth by additionally
reinjecting the zero diffractive order generated by the grating, due to the imperfect
polarisation of the fibre output. Up to 40mW of output power could be extracted from
the oscillator at the peak of the tuning range by replacing the 2% output coupler with

57
one of 8% reflectivity, resulting in a 15% increase of pump threshold from 500mW to
~600mW.

Fig.2.2.3. Typical spectrum of CW Raman oscillator with intracavity grating as tuning


element.

The fibre oscillator offered a viable alternative to colour centre sources


operating at this wavelength range which are expensive and require sophisticated
cooling (NaCl colour centres require storage below about 100K). By optimisation of
parameters such as fibre length, which could not be cut, and by incorporating fused
couplers to inject the pump and form the ring cavity, it should be possible to lower the
pump threshold, increase the output coupling substantially and hence the output
power, and increase the tunability to cover the full Raman band.

This experiment offers an introduction to the process of Raman conversion in


the CW oscillation regime. In the next section, a pulse pumped system in the transient
regime is described, synchronously mode-locked by the pump. Walk-off between
pump and Stokes is considered and compensated, as is dispersion in the fibre.

2.3 A Synchronously Pumped Dispersion Compensated Raman Source at


1.41μm.

58
When a Raman oscillator is pumped by intense laser pulses, the Raman shifted
pulses experiences a number of effects which make them suitable for compression
using a negatively dispersive delay line. The Raman pulses are generated at the peak
of the pump pulse, where the intensity and hence Raman gain is highest, within a short
propagation length for intense excitation. The Raman shifted signal propagates faster
than the pump pulse through normal dispersion, and hence walks through the front of
the pump pulse, severely depleting it. The Raman pulses generated are of roughly the
same duration as the pump, and can suffer Self Phase Modulation (SPM) as it
propagates and Cross Phase Modulation (XPM) from the pump. The combined effects
of SPM, XPM from the Raman depleted pump, and considerable normal dispersive
broadening result in a complicated frequency chirp across the Raman pulses, a portion
of which is approximately linear. By arranging the fibre in a synchronously pumped
ring geometry, the reinjected Raman signal can receive further amplification from a
pump pulse injected in synchronism, and oscillation can occur. Since the Raman pulse
is chirped, adjustment of the cavity length synchronises different frequency
components of the Raman pulse with the pump pulse, and the oscillator can be
tuned7,7a.

The Raman signal is assumed to be in the normal dispersion regime (β2>0)

otherwise soliton shaping can occur, which is discussed in chapter 6, resulting in


pulses in the femtosecond regime being generated8,9,10,11,12. Traditionally,
synchronously pumped Raman lasers operating in the normal dispersion regime have
produced output pulses which are of the same order of duration as the pump pulses7,
as the SPM and XPM give rise to enhanced temporal broadening through positive
GVD. To overcome this effect, a negatively dispersive delay line such as a grating pair
can be incorprorated in the cavity, which compensates for positive dispersion in the
fibre, resulting in subpicosecond pulse generation13,7a.

In this section the construction, characterisation of such a system operating at

59
the 1.41μm Raman shifted wavelength of a 1.32μm pump is discussed. The pulses
obtained were ~2 ps duration with peak powers of ~100W. These pulses were
approximately transform limited, and hence perfectly suitable for soliton propagation
studies (see chapter 6),.in contrast to those generated from single pass soliton Raman
generators or soliton Raman fibre ring resonators8,9,10,11,12. The experimental schematic
of the ring laser is shown in Fig.2.3.1.

Fig.2.3.1. Schematic of the dispersion compensated fibre Raman ring laser.

A fibre with a zero dispersion wavelength λ0=1.46μm was used, to ensured

that the pump at 1.32μm and Raman shifted signal at ~1.4μm both experienced normal
GVD (β2>0). A CW mode-locked Nd:YAG laser operating at 1.32μm generating

100ps pulses at a 100MHz repetition rate with an average power of 2W (~200W peak
power) provided the pump radiation. This radiation was directed toward the input end
of the fibre via beam splitter (BS1) (100%R at 1.32μm , 95%T at 1.4μm for 45o). The
pump radiation was focussed into the fibre using an achromatic long working distance
X20 broadband Anti Reflection Coated Objective (ARCO) lens (L1). The Dispersion-

Shifted Fibre (DSF) was 400m long, with a core diameter of 7μm and was single mode
at both the pump and Stokes wavelengths.
The fibre loss was less than 0.6dBkm-1 in the 1.2-1.6μm spectral region, with

60
an OH- absoprtion peak at ~1.39μm where the loss rose to ~2dB/km. The DSF was
tailored to have a dispersion zero wavelength λ0=1.46μm, to ensure both pump and
Stokes were in the positive GVD (β2>0) regime. Fig.2.3.2. shows both the measured

group delay and GVD characteristics of the DSF versus wavelength:

Fig.2.3.2. Measured group delay and GVD vs. wavelength of the DSF. (Courtesy
BTRL)

As shown, the average value of the GVD between the pump and Stokes
wavelengths was ~6.4ps/nm/km. The Stokes shift in this fibre was ~80nm around this
pump wavelength, and hence the relative difference in group delay for these two
wavelengths was approximately 80 X 6.4 = 512ps/km of fibre. This is in good
agreement with the measured 520ps/km group delay difference shown in the second
trace in Fig.2.3.2. The time difference between the pump and Stokes over the 400m
fibre length was therefore ~210ps. This means the Stokes walked forward by ~210ps
through the front of the 100ps pump pulse, in 400m of fibre. The fibre was
approximately 2 walk-off lengths, which was ideal since the first Stokes becomes
maximal after 1.5-2 walk-off lengths.

Approximately 40% pump power coupling efficiency was achieved at the fibre

61
input, and the output radiation was collected and collimated with ARCO (L2) identical
to (L1). The output radiation was directed via aluminium mirror (M1) through a

negatively dispersive delay line, comprising a pair of holographic gratings (G1) and
(G2) arranged single pass with their faces parallel. The gratings had 1200 lines/mm,
and exhibited 75% diffraction efficiency into the -1 diffractive order at 70o angle of
incidence. Mirror (M2) directed the diffracted radiation from the dispersive delay line

through an aperture which allowed the lasing wavelength to be selectively defined,


back through (BS1) and back into the input objective (L1). A 2nm bandpass dielectric

filter centred at 1.4μm was inserted in the cavity to restrict the oscillating bandwidth
of the laser and prevent satellite pulses building up from noise in the cavity. The
output coupler (BS2) had a reflectivity of 20% at 1.4μm.

The fibre coupling block containing lens (L2) was mounted on a translation

stage to allow fine adjustment of the oscillator cavity length, with micron precision.
This allowed synchronisation of the selected Raman signal with a pulse pump from the
pump laser, by maximising the output signal for any particular wavelength governed
by the position of the aperture and the angle of the dielectric filter, which allowed a
small degree of tunability. Since the gratings were most efficient for p-polarised light,
as was the Raman conversion for the p-polarised pump, a fibre polarisation strainer
was inserted at the output end of the fibre.

The output from the oscillator was continually monitored using a vibrating
mirror real time SHG autocorrelator (see appendix), and a 1m scanning spectrograph,
allowing optimisation and characterisation of the sytstem. Whilst gradually increasing
the pump power, the amount of SPM and XPM experienced by the Raman pulses
circulating in the ring varied. To compensate for this variation the distance between
the gratings was adjusted. Through reoptimisation of the cavity length, the shortest
pulses with highest peak powers could be obtained by monitoring the autocorrelation.
Thus the optimum grating separation was found to be 18cm.
The oscillator was essentially unidirectional, as the synchronous pumping only

62
occurred in one direction, and hence the pump threshold for single pass Raman
generation Pth was given by

30Aeff
P th ≈
gL

for non-polarisation preserving fibre, where the mode field area Aeff~70μm2, and the

Raman gain coefficient g=0.74X10-13 mW-1 at 1.32μm. For an interaction length equal
to the walk-off length, i.e. 190m, this yielded a pump power of 82W peak,
corresponding to an average pump power of 820mW. By maintaining the pump power
at 400mW, no single pass Raman generation was detectable around 1.4μm. To
estimate the threshold for oscillation in a ring configuration, an estimate of the loss
was required. However the loss is dependent on factors such as the oscillator
bandwidth, which is highly dependent on pump power through SPM and XPM, and
hence could not be estimated.

Once optimised for the shortest pulses and highest power, the system routinely
generated 2.7ps pulses with up to 20mW of average output power available,
corresponding to 74W peak in the pulses at 100MHz repetition rate determined by the
pump source. The output bandwidth was ~2nm, being limited directly by the
intracavity dielectric tuning filter. For grating separations +/- 10 to 15% of the
optimum 18cm, the pulses became substantially longer. The system could be coarsely
tuned from 1390 to 1415nm by adjusting the horizontal position of the aperture across
the dispersed spectrum emerging from the grating pair, and finely tuned by carefully
tilting the intracavity dielectric filter, which altered the 2nm bandpass centre
wavelength over this range. Each time tuning took place, the cavity length required
reoptimisation by adjustment of the (L2) translation stage. A typical output spectrum

corresponding to operation at ~1.392μm is shown in Fig.2.3.3.

63
Fig.2.3.3. Typical output spectrum of the fibre raman ring oscillator.

The threshold for oscillation was measured to be ~350mW of average pump


power, which corresponded to an estimated single pass oscillator loss of ~30dB, and a
single pass Raman gain of ~0.1dBmW-1 of pump radiation. This was consistent, since
loss through the filter, gratings, aperture and diffraction were considerable. The
temporal behaviour of the output of the oscillator was similar in character to that of
many synchronously pumped systems, in that cavity length adjustment was critical for
ultrashort pulse generation. The cavity length adjustment did not result in time
dispersion tuning7,7a, since the bandpass filter fixed the laser wavelength. Fig.2.3.4.
shows background free autocorrelation traces of the output pulses from the oscillator.

Fig.2.3.4. Background free autocorrelation traces of the output from the dispersion
compensated fibre Raman ring laser. Oscillator cavity length (a) optimised and (b)
mismatched by +/-10μm.

64
In Fig.2.3.4(a). the cavity length was optimised for the best possible output,
resulting in an output pulse duration of 2.7ps, assuming a Gaussian pulse shape. For a
cavity length detuning of +10μm, the laser generated pulses with temporal
characteristics attributed to the autocorrelation function of a noise burst (see
appendix), as shown in Fig.2.3.4(b). This suggests that the oscillator is producing
bursts of noise of ~20-50ps duration, with temporal features as short as a few
picoseconds. Similar behaviour was noted for detunings of -10μm. When operated
close to threshold pump power, it was possible by carefully adjusting the polarisation
strainer in the fibre ring to generate pulses as short as 1.8ps, yielding an optimum time
bandwidth product of 0.54 which is close to transform limited operation, however
more stable operation was obtainable when the pump power was higher, where the
shortest pulses measured were 2.7ps. These pulses, assuming a Gaussian pulse shape,
were approximately twice the transform limited duration for the measured 2nm
oscillating bandwidth.

It is likely that the quadratic compression occuring in the negatively dispersive


delay line was not properly compensating for the chirp imposed on the 1.4μm Raman
signal.This may be the reason for operation above the transform limit. The induced
phase modulation is in fact not perfectly quadratic, as it arises through SPM and XPM,
which is complicated15 by the effects of walk-off and pump depletion, although the
measured chirp16 on a Raman signal has been observed to be reasonably linear over
small portions of the Stokes pulse.

In conclusion, a dispersion compensated fibre Raman ring laser has been mode
locked through synchronous pumping, producing stable 2.7ps pulses with an average
output power of 20mW (74W peak) at a 100MHz repetition rate for a pump power of
400mW. The source was tunable over the range 1390-1415nm, and the output is close
to transform limited. The source was very useful for soliton studies, as the output
wavelength falls in the anomalous dispersion regime of standard optical fibres. The
output wavelength makes the pulses suitable for studying the effects of synchronous

65
Raman amplification, using pump radiation from the same or a similar Nd:YAG laser
operating at 1.32μm. The synchronously pumped oscillator cavity length could also be
misadjusted, to produce noise bursts of similar duration to the pump source,
containing picosecond ultrashort features, which are useful for noise amplification
studies. The source was used extensively to obtain results presented in chapter 6, from
a variety of experiments involving synchronous amplification and soliton generation.

2.4 References

(1) D.Cotter. J.Opt.Comm. 4,10(1983)


U U

(2) B.Schuker & R.W.Gammon. Phys.Rev.Lett. 25,222(1970) U U

(3) from "Fibre Raman Lasers". C.Lin. In "Tunable Lasers". Ed. L.F.Mollenauer &
J.White. Springer-Verlag Topics in Appl. Physics 59 U

(4) R.K.Jain, C.Lin, R.H.Stolen & A.Ashkin. Appl.Phys.Lett. 31,89(1977) U U

(5) C.Lin, R.H.Stolen, W.G.French & T.Malone. Opt.Lett. 1,96(1977) U U

(6) K.O.Hill, B.S.Kawasaki & D.C.Johnson. Appl.Phys.Lett. 29,181(1976) U U

(7) R.H.Stolen, C.Lin & R.K.Jain. Appl.Phys.Lett. 30,340(1977) U U

(7a) E.M.Dianov, P.V.Mamyshev, A.M.Prokhorov & D.G.Fursa. Pis'ma


Zh..Eksp.Teor.Fiz. 45,469(1987)
U U

(8) A.S.Gouveia-Neto, A.S.L.Gomes & J.R.Taylor.


IEEE J.Quant.Elec 24,332(1988)
U U

(9) J.D.Kafka & T.Baer. Opt.Lett. 12,181(1987)


U U

(10) A.S.Gouveia-Neto, A.S.L.Gomes, J.R.Taylor, B.J.Ainslie & S.P.Craig.


Opt.Lett. 12,295(1987)
U U

(11) M.N.Islam, L.F.Mollenauer & R.H.Stolen. "Ultrafast Phenomena V".


Springer series in Chemical Physics. 46,46(1986)
U U

(12) B.Zysset, P.Beaud, W.Hodel & H.P.Weber. "Ultrafast Phenomena V".


Springer series in Chemical Physics. 46,54(1986)
U U

(13) J.D.Kafka, D.F.Head & T.Baer. "Ultrafast Phenomena V".


Springer series in Chemical Physics. 46(1986)
U U

(14) B.J.Ainslie & C.R.Day. J.Lightw.Tech. LT-4,967(1986)


U U

(15) M.N.Islam, L.F.Mollenauer, R.H.Stolen, J.R.Simpson & H.T.Shang.


Opt.Lett. 12,625(1987)
U U

(16) A.S.L.Gomes, V.L. da Silva & J.R. Taylor. J.Opt.Soc.Am. B5,373(1988) U U

66
3. MODE LOCKED DOPED FIBRE LASER SOURCES

3.1 Introduction

Rare earth doped optical fibres have a broad fluorescence linwidth, due to the
homogeneous nature of the glass host crystal field. A high gain can be obtained over
relatively short fibre lengths, provided pumping is efficient, making doped fibre
amplifiers also ideal for incorporation in fibre lasers. The gain can be so high that laser
action is obtainable using only the cleaved fibre facets as reflectors, although this
could lead to problems through amplification of spontaneous emission in long range
communications. Fibre amplifiers of this kind are directly compatible with optical
communications in the linear or nonlinear regime, with the possibility of injecting
pump radiation periodically using fused couplers.

Maximum transmission speeds in fibre based transmission still fall far short of
those determined by the low loss bandwidth, even in the very best soliton transmission
systems, and to increase capacity multiplexing in both time and frequency may be
necessary. In a number of application such as multiplexing/demultiplexing, stable
sources of pulses at ultra-high (100GHz up) repetition rates are required with high
distributive fanout. These sources will effectively be used to derive signal pulses for
switching and clock pulses throughout the system and for synchronisation. A soliton
based system would benefit from a source generating soliton-like pulses, to prevent
the propagation of intrapulse radiation shed in the soliton formation process, which
may result in long range distortion through soliton - soliton interactions. Erbium
doped fibre has a fluorescence band around the region of lowest loss at 1.55μm, and
hence much interest has been dedicated to the development of an Erbium fibre laser
producing ultrasort pulses at ultra-high repetition rates.

Semiconductor diode lasers are ideal sources for use in high bit rate
transmission, since they demonstrate high efficiency and controllability, but they yield

67
highly chirped pulses at even modest gain switching frequencies and low output
powers. The powers available from doped fibre lasers are much higher, and active
mode locking using bulk or 'pigtailed' integrated modulators offers controllability over
output, but repetition rates are limited to the electro-optic speeds. The marriage of
high speed laser diodes as pulse sources with amplification in doped fibre offers an
alternative way of reaching the high repetition rate, high power and the direct
controllability goal. Fibre sources such as Neodymium and Erbium are capable of
producing more than 0.5W of average output power, and since soliton transmission is
curently limited to pulse durations of around 10ps to avoid Soliton Self-Frequency
Shift (SSFS), this would represent peak powers well into the nonlinear regime, making
fibre lasers more than capable of exciting solitons. Fibre amplifiers have optical
bandwidths of around 1THz (e.g. Nd, Er) and are therefore capable of supporting
pulses of this duration.

Since optical fibres can be tailored so that the dispersion minimum wavelength
occurs beyond 1.5μm, a compact, Erbium based laser producing soliton pulses in the
low GVD regime is quite feasible. The process of Modulational Instability (MI) (see
chapter 5) occurring in the anomalous dispersion regime offers the oportunity to
generate pulses in the THz regime. This effect could also be incorporated into an
Erbium fibre laser to produce a source which is perhaps completely passive. MI can
grow from amplitude and phase noise fluctuations, or can be induced by amplitude
modulation from XPM or frequency mixing, and is therefore controllable. The
ultimate modulation rate is limited by high order dispersion terms which dominate
around λo.

In this chapter, three mode locked fibre lasers are presented offering the
possibility of high output power and high repetition rates. The Nd based fibre laser
uses a novel mode locking technique based on interferometrically induced nonlinear
coupling between two cavities. An Er doped fibre laser is presented based on
traditional amplitude modulation mode locking, and has been used in an experiment

68
presented in chapter 5, where modulational instability was induced from the output
pulses. A further system using both an Er doped fibre and a laser diode as a modulator
allows the benefits of both systems to be simultaneously exploited for the first time.

3.2 Active Mode Locking of an Erbium Doped Fibre Laser Through

Amplitude Modulation.

The region of lowest loss in optical fibres occurs in the 1.5-1.6μm wavelength
region, as shown in Fig.1.3.6. Losses of less than 0.2dB/km are obtainable, allowing
transmission over 10-20km with only half power loss. Much interest is shown in the
application of Erbium doped fibre amplifiers in extending the range of transmission
systems based on optical fibres. Very high gain is available1 (up to 46dB pumped with
a 1480μm high power laser diode), and a very broad gain bandwidth2 around the
fluorescence band at 1.55μm.

An Erbium based fibre laser presents an ideal source for investigating


nonlinear effects which affect transmission in fibres. Alternatives exist, such as laser
diodes which can produce many tens of milliwatts of CW output powers, but produce
chirped pulses3 and suffer a significant power loss when gain switched. The pulses are
chirped due to the induced change in refractive index which occurs as the gain
switching current is modulated. Synchronously pumped colour centre lasers operating
around 1.5μm are capable of producing pulses of ~ 8ps directly4, but are limited in use
through cost, are bulky, and normally require cryogenic cooling. Pulses in the 1.5μm
region can also be produced through Raman scattering from shorter wavelength pump
signals or by soliton Raman generation, such as that outlined in chapter 6. The soliton
self-frequency shift can also be used to upshift the wavelength of solitons launched at
lower wavlengths5, however both techniques produce significant jitter, and uncertainty
as to the number, duration and arrival time of the signal pulses. The source described
here was required to produce relatively long (100ps or longer) pulses, with peak
powers in the 1W regime, in order to excite modulational instability in an experiment

69
described in chapter 5, and hence Er doped fibre was an ideal gain medium. Erbium
doped fibres can produce high average powers6, whilst they provide a bandwidth2
capable of amplifying pulses as short as 100fs7. In this section the construction of a
mode locked Erbium fibre laser is described, using active mode locking by amplitude
modulation.

The first reported mode locking of an Er fibre laser8 used a bulk electro-optic
phase modulator, producing 70ps pulses, but with only ~100μW of output power. This
represented only around 10mW of peak power, which would result in a nonlinear
length given in eq(1.4.12) longer than 40km. These pulses are useless for nonlinear
propagation studies, since the fibre loss length is much shorter than this value. Shorter
pulses have been generated9 by incorporating soliton shaping in a fused coupled, fibre
ring oscillator, but the output power was still as low. To facilitate higher power output
in the system decribed here, a fibre which demonstrated high CW average output
power was incorporated into a low loss linear cavity arrangement, pumped with a
small frame 2W Argon Ion laser operating at 514nm. Despite the pump wavelength,
excited state absorption of the pump signal did not dominate10, and high output power
was available (~100mw) The experimental arrangement is shown in Fig.3.2.1.

Fig.3.2.1. Experimental arrangement of the actively mode locked Er fibre laser.

The 514nm pump radiation was directed through beamsplitter (BS) (95%T at

70
514nm, 95%R around 1.55μm) and into a X20 anti reflection coated objective
(ARCO) lens for high transmission in the 1.5μm region. The pump beam was focussed
through a cavity mirror placed in contact with the normally cleaved Er fibre end face,
and into the Er fibre. This 'butt coupling ' technique provided laser output coupling at
the fibre end over the coating range of the mirror (90%T at 514nm, 20%T at 1.56μm).
Of the 2W of pump power available, up to 800mW could be launched into the Er
doped fibre, which was ~3.2m long with an outer diameter of 130μm, a core diameter
of ~15μm. The fibre was highly multimode at the pump wavelength, and the large core
diameter made it weakly multimode at the laser wavelength as well. At 514nm the
absorption was such that 90% of the estimated launched pump radiation was absorbed
over the fibre length. In order to eliminate self laser action off the other fibre end face,
and to suppress sub cavity effects which may destroy mode locking, a quartz wedge
polished at 7o was placed in optical contact to the fibre end with index matching gel.
Radiation exiting the fibre end and passing through the wedge was collected and
collimated with a X10 ARCO lens, coated for ultra low loss over the 1.55μm region.
The beam emerging from the lens was passed through the amplitude modulator, which
was connected to a stabilised source of RF power.

A precision mounted adjustable aperture was inserted between the modulator


and the grating, which formed the cavity end mirror. This holographic grating had 900
lines/mm, and was blazed for high efficiency retrodiffraction around 1.5μm in the
Littro configuration. The grating was mounted in order to enable tuning in a vertical
plane, and cavity length adjustment with micron precision. The amplitude modulator
was an acousto-optic modulator of fused quartz similar to that found in commercial
Nd:Yag lasers, designed to be driven at 49.9MHz. The modulator was driven with 7W
of RF signal derived from a stabilised crystal oscillator, but produced only a very low
modulation depth. The dielectric cavity mirror contacted to the front end of the mirror
compensated for the low modulation depth by increasing the cavity Q, allowing mode
locking to build up over a number of round trips. To ensure the highest diffraction
efficiency from the grating and highest modulation depth in the modulator, a fibre

71
polarisation strainer was incorporated at the fibre end nearest the grating. The aperture
helped restrict the oscillation bandwidth8 by spatially restricting the bandwidth of the
beam retrodiffracted from the grating, and ensured that high order transverse modes
did not oscillate. Stable mode locked oscillation required critical adjustment of the
polarisation strainer, and optimisation of the cavity length.

Initially the laser output from the 20% transmission end was analysed with a
photodiode and sampling oscilloscope providing a combined temporal resolution of
~320ps, as shown in Fig.3.2.2. An isolator was placed in the laser output path to
ensure that reflections from detector apparatus did not destroy the mode locking,
which was extremely sensitive to perturbations. The cavity length was adjusted to give
a round trip of approximately 40ns, to enable 4th harmonic mode locking of the
fundamental modulation frequency, which was ~100MHz.

Fig.3.2.2. Sampling oscilloscope trace of the mode locked Er fibre laser output

The considerable ringing in the signal was characteristic of the impulse


response of the detection electronics, and hence a shorter pulse than the 350ps
measured value was presumed to generating the signal. The measured repetition rate
was ~100MHz, with a slight envelope modulation at 40ns period resulting from
imperfect harmonic mode locking. The optimised output power was over 60mW
average in a predominantly fundamental transverse mode profile, once the aperture
and alignment was correctly adjusted. Higher power could be obtained by allowing

72
hybrid transverse mode oscillation, but the mode locked output was very irregular and
exhibited multiple pulsing.

Fig.3.2.3(a). shows a streak camera trace of the mode locked output, split by an
optical delay of 387ps. The measured FWHM pulse duration was 140ps, which would
require a mode locked bandwidth of 0.02nm assuming a transform limited sech2 pulse.
The CW spectrum obtained with the modulator switched off is shown in Fig.3.2.3(b),
when the laser was tuned as in (a) to 1.565μm. The measured bandwidth corresponded
approximately to the resolution limit of the spectrograph, which was ~0.07nm. With
the modulator switched on again, the mode locked bandwidth deconvolved from the
resolution limit was 0.2nm, as shown in Fig.3.2.3(c). This indicated the laser was
operating approximately 18X the transform limit, probably generating chirped pulses.
Semi-interferometric autocorreleation measurements (see appendix) taken later in
chapter 5 confirmed that the pulses were chirped, with fringe visibility over less than
10% of the pulse profile.

Fig.3.2.3.(a) Streak camera trace, (b) CW and (c) mode locked spectrum of fibre laser
output.

73
Tuning over the range 1530 to 1580nm was possible, without noticable pulse
duration change over the 1555 to 1575nm range of high reflection (peaking at 80%)
for the narrow band dielectric cavity mirror. Outside this range, mode locking was
accompanied by severe relaxation oscillations, indicating that loss in the cavity was
becoming significant. Mode locking could not be obtained without this cavity end
mirror as losses were assumed to be too high, and laser performance may have been
improved by using a broad band cavity end mirror.

For an average output power of 60mW average, the 140ps pulses at 100MHz
repetition rate corresponded to 4.3W of peak power in each pulse. This peak power
was high enough to induce weak nonlinearity within the laser cavity, however the
nonlinear length NNL corresponding to this intacavity power was a few hundred
metres, and the soliton period Z0 for this situation would be in excess of 100km.

Hence no nonlinear pulse shaping effects were expected or measured.

In conclusion up to 60mW average (~4Wpeak) power has been obtained in


140ps pulses tunable from 1555 to 1575nm from an Er doped fibre laser, using a bulk
acousto-optic modulator. The output signal has application in a variety of experiments
investigating nonlinear pulse shaping effects in fibres, such as those discussed in
chapter 5, where modulational instability was induced at around 300GHz.
Substantially shorter pulses would be expected from a similar system with an intra-
cavity phase modulator, which normally results in a much deeper modulation per
round trip, or with a high Q amplitude modulator. The high gain per pass and high
loss at the output coupler make this a highly perturbative system, and hence any pulse
shaping or mode locking must build up in a relatively short number of cavity transits.
Reducing the output coupling in this case compromised the available output power,
whilst making no noticeable difference to the pulse duration. The inclusion of a
nonlinear fibre within the cavity to provide pulse shaping has been shown to generate
shorter pulses9 through soliton shaping, and it is possible that both gain and soliton
shaping effects could be provided by the same fibre This may produce a source of

74
solitons directly from an Erbium fibre laser.

3.3 Active Mode Locking of an Erbium Fibre Laser using an Intra-

Cavity Laser Diode Device.

Relatively high average powers can now be obtained from CW semiconductor


laser diodes, and high modulation rates can be obtained through direct electrical
pulsing. However, current modulation results in a rapid transient occurring in the
carrier density, and a large chirp is imposed across the generated optical pulse.
Generally pulsewidths of 15-40ps can be generated by gain switching laser diodes, and
time resolved spectral measurements have revealed a red shift or negative frequency
chirp12. This chirp varies linearly as the modulation frequency increases, and for high
repetition rate transmission in the anomalous dispersion regime this makes them
unsuitable as the sign of dispersion broadens the pulses drastically. Gain switching
also results in a drastic reduction of the average output power, and hence the
advantages of gain switching for enhancing peak power are only marginal.

Currently, very high average output powers can be obtained from Erbium
doped amplifiers and lasers. The experiment presented in the last section described a
laser producing 60mW of average power, and up to 500mW has been obtained from an
Er doped fibre amplifier at 1550nm13. In fact the output from fibre laser systems is
now comparable with many small frame bulk solid state lasers.

Laser diodes and Erbium amplfiers have previously been integrated14,15, to


produce transform limited pulses from GHz gain switched laser diodes. A distributed
feedback laser diode producing chirped pulses was spectrally windowed using a
narrow band pass Fabry-Perot etalon, and the short, transform limited, low power
pulses were then amplified up to soliton power levels. Integration of this kind allows
utilisation of the high modulation rates available from semiconductor devices,

75
combined with the high average power capability of the Er doped fibre medium. Diode
lasers provide a cheap and simple alternative to bulk or pigtailed modulators used
previously to mode lock Er fibre lasers9,16,17, and require much lower RF drive signals.
Semiconductor lasers can be modulated by direct modulation of the pumping current
at rates exceeding 10GHz, making them extremely important components for high bit-
rate optical communication. They can be monolithically integrated with other
optoelectronic components to form hybrid processing circuits. The output beam has
compatible dimensions to those of typical silica based fibres, and the output
wavelength can be tailored to that of the lowest loss and dispersion.

If no current is injected into a laser diode chip, it exhibits absorption at


wavelengths shorter than that at which it normally emits, corresponding to energy
levels higher than the bandgap energy associated with the laser transition. As carriers
are injected by applying a forward current, more electrons are excited from the
valence (ground state) band to the conduction (upper state) and the material begins to
absorb less, culminating in complete transparency at some current level below which
the device starts to exhibit laser action, since it also has to overcome cavity losses in
order to oscillate. Round trip losses are typically between 30% and 60%, since the
cavity mirrors are normally formed by the cleaved facet of the heterostructure, which
has a refractive index around 3.5.

A very short electrical pulse (<100ps) drive signal can readily be applied to a
laser diode device, allowing the device to operate as a very fast loss modulator, with
open times of approximately the same duration as the typical laser output pulse
durations, which can be as short as 15ps for modulation rates approaching 30GHz. In
this section the application of a GaInAsP ridge waveguide laser diode chip as a
modulator in an Erbium doped fibre laser is described, resulting in mode locking of
the fibre laser.

The experimental configuration employed is shown in Fig.3.3.1. The pump

76
power for the Er fibre laser was once again provided by a small frame Argon Ion laser
producing 2W of CW radiation at 514nm. The pump radiation was directed through
beamsplitter (BS) (95%T at 514nm, 95%R around 1.55μm) and through dielectric
cavity mirror (M) (90%T at 541nm, 50%R around 1.55μm) into a X20 anti reflection
coated objective (ARCO) lens for high transmission in the 1.5μm region.

Fig.3.3.1. Schematic of the experimental arrangement.

Of the 2W of pump radiation, up to 800mW of radiation could be coupled into


the Er doped fibre. The fibre was 3m long and was single mode above 1.17μm, with
an outer diameter of 115μm, and a Δn of 0.008 . At 1.55μm the absorption was
8.8dBm-1, and the absorption at 514nm was such that 90% of the pump radiation was
absorbed over the fibre length.

In order to suppress sub cavity effects and eliminate laser action which readily
occurred from reflection at the fibre ends, quartz discs with a 7o wedge were contacted
to the fibre ends with index matching gel. In this way laser action was effectively
suppressed off the fibre facets. Mirror (M) was mounted on a translation stage to allow
coarse tuning of the cavity length. Radiation emerging from the Er fibre was collected,
collimated and refocussed with two identical ARCO lenses (MO2 &MO3) into the

laser diode chip (LD). The lens coatings were for ultra low loss in the 1.55μm region.

77
Alignment was initially facilitated by amplifying radiation from the laser diode
when driven CW above threshold, in the Er fibre single pass. When driven CW from a
D.C. current supply at approximateley 100mA, the laser diode alone generated
average output powers of up to 30mW. This device, which was used as a modulator
was a GaInAsP ridge waveguide device, 280μm long, with one uncoated facet parallel
to the other, which was ~100% reflection coated around the lasing wavelength. This
highly reflective facet formed one cavity end mirror of the Er fibre laser, in a linear
configuration. Pulses to drive the semiconductor modulator were derived from an
amplified RF frequency synthesizer (RF Osc). This signal was used to drive a step
recovery diode comb generator (Hewlett-Packard 33004A) at ~500MHz, with an
average RF power of 600mW. The derived pulse train consisting of electrical pulses
with ~90ps duration was coupled into the semiconductor modulator via a bias-tee
network in conjunction with the current from a variable DC supply.

The output from the Er fibre laser collected from beamsplitter (BS) was passed
through an optical isolator, to ensure that reflections from detection apparatus did not
disrupt the operation of the laser. The output signal was then analysed with a
photodiode/sampling oscilloscope with a combined temporal resolution of 320ps to
initially obtain stable mode locking. The overall fibre laser cavity defined by dielectric
mirror (M) and the highly reflective facet of the laser diode chip was initially arranged
to have a length seventeen times that of the fundamental, determined approximately by
the 500MHz RF drive frequency supplied to the diode. With no DC bias current to the
diode, laser action in the Er fibre was suppressed, through absorption in the unpumped
laser diode chip. By increasing the DC bias current up to 150mA, laser action was
allowed to occur, dominated by self laser action of the diode at 1550nm. A DC current
threshold for laser action in this configuration of 45mA was measured.

At a 20mA DC current, no output whatsoever was detectable from the laser


diode, nor with 600mW of RF pulses applied via the bias-tee, even with a sensitive
large area detector. Mode locked operation of the Er doped fibre laser was however

78
readily detected with the fast photodiode. For a fixed cavity length, the RF drive
frequency was adjusted in order to obtain the minimum pulsewidth from the Er fibre
laser. At a drive frequency of 500.475MHz, routine operation with 40ps pulse
generation was achieved. Fig.3.3.2. shows a typical output pulse of the 37ps duration
pulses recorded under these conditions with a Hamamatsu OOS-1/IR optical sampling
oscilloscope.

Fig.3.3.2. Optical sampling oscilloscope trace of 37ps pulses from the Er fibre laser
mode locked with an intra-cavity laser diode as a modulator.

A variation in laser diode drive frequency of +/-5KHz lead to a measurable


increase in the output pulsewidth. At optimum cavity length/drive frequency match,
the pulses generated were relatively symmetric with no evidence of envelope
modulation, secondary pulsing or inter pulse radiation. The spectral output of the
optimised mode locked laser is shown in Fig.3.3.3., indicating a FWHM bandwidth of
1nm at a lasing wavelength of 1.531μm.

When the Erbium laser was allowed to operate autonomously by inserting a


100% mirror between the lenses (MO2 & MO3), the CW output spectrum was centred

around the same wavelength, i.e.~1.53μm. This is evidence that the Er fibre gain is
providing mode locked pulses, rather than the output simply consisting of amplified
laser diode pulses, which would give a spectrum centred around 1.55μm. Single pass
amplification of the laser diode pulses in a similar configuration provided ~50ps
pulses centred around 1.55μm. A transform limited sech2 pulse of ~40ps duration at

79
1.531μm requires a mode locked bandwidth of 0.06nm, indicating that operation was
above the transform limit. This was not suprising considering the negatively chirped
nature of pulses produced from a laser diode of this kind, when modulated at high
frequency. The optical sampling oscilloscope trace may have been broader than that of
the pulse duration, if any jitter occurred between the output pulses and the trigger
signal, which was derived directly from the RF oscillator.

Fig.3.3.3. Spectral characteristic of Er fibre laser output pulses shown in Fig.3.2.2.

The measured output power of the laser was ~20mW, which corresponded to a
peak power of ~1W in the 40ps pulses at 500MHz repetition rate. The fundamental
soliton power for a 40ps pulse in a fibre similar to the Er fibre used here was ~1mW
peak, but with a corresponding soliton period Z0>20km. The nonlinear length scale
LNL>500m, thus no soliton or nonlinear pulse shaping mechanisms would have been

strong enough to affect the laser, unless the pulses evolved over hundreds of cavity
transits.

It is possible that the 280μm diode laser device which was effectively placed
intra-cavity in the Er fibre laser was operating as a bandwidth limiting element, since

80
the output facet reflectivity was as high as ~30% on the uncoated side. This possibly
stabilised the laser output as well, which tended to be spectrally unstable if driven CW
without the diode intracavity. For an etalon corresponding to the geometrical length of
the laser diode, the free spectral range Δν is given by

Δν = c
2nL

where n is the semiconductor refractive index which is ~3.5. From this equation,
Δν=153GHz for the device used here, which corresponded to a free spectral range of
approximately 1.2nm at 1.53μm.

The finesse F of a similar etalon, defined as the ratio of the free spectral range
to the FWHM transmission bandwidth is given by

F = π R1 R2
1- R1 R2

where R1 & R2 are the reflectivities of the etalon mirrors. For facet reflectivities of

30% and 100%, typical of those involved here, the finesse F was approximately 2.46,
which would have yielded a transmissive bandwidth of approximately 0.5nm. Hence it
is reasonable to suggest that this element had a bandwidth limiting effect on the
operation of the laser, and would limit the shortest pulse obtainable to 5ps duration, if
dispersion and chirp could have been compensated for in some way, and operation
was transform limited.

Substantially shorter pulse could possibly have been obtained by introducing a


Brewster angle or an anti reflection coating to the laser diode facet. This would have
removed any spectral limiting effects presented by the laser diode chip, and allowed
substantially higher RF modulation and DC bias signals to be applied, without the
possibility of the diode lasing by itself. It is possible however that some form of
dispersion compensation would have been required to compensate for the chirp

81
imposed by the laser diode carrier density fluctuation, by the use of a normally
dispersive fibre or a dispersive delay line.
In conclusion active mode locking of an Erbium doped fibre laser has been
demonstrated using a semiconductor laser diode as a modulator. The technique allows
routine generation of 40ps pulses at a 500MHz repetition rate, limited only by the
electronic driving circuitry. The Er fibre laser has generated average powers of up to
20mW, indicating that this relatively inexpensive method of modulation has the
potential for producing substantially higher peak powers if more efficient pump
schemes are employed, or shorter pulses are generated. This laser could be completely
made up of optoelectronic devices if a semiconductor laser pump was used, and may
find application to the study of high bit rate nonlinear optical communication systems.
The repetition rate of the system could theoretically be extended to the highest
frequency that the semiconductor device could cope with.

3.4 Mode Locking of a CW Neodymium Doped Fibre Laser with Linear

External Cavity Feedback

In the introduction the importance of high repetition rate sources to optical


signal processing and ultrafast transmission was emphasised. Moderately high average
powers are advantageous to increase distributive fanout to various signal processing
elements. Currently the highest repetition rates from CW mode locked lasers
(excluding semiconductors) are limited by electro optic modulation to around 1GHz.
Laser diodes can provide much higher rates, but to date the average powers produced
are relatively low. This shortfall can be corrected by using the technique described in
the last section, using the diode as a modulation source in a fibre laser.

In order to significantly increase obtainable pulse repetition rates, it may be


necessary and more fruitful to rely on passive modulation techniques using nonlinear
effects within the laser medium, such as self pulsing instablilities (Modulational
Instability - see chapter 5). Terahertz repetition rate pulses have been demonstrated

82
through this process18,19, though only over limited duration pulse trains. Up to 100
GHz modulations have been generated in a CW train20, but Brillouin scattering
hampered the process, limiting the modulation depth and significantly attenuating the
signal, which would have required further amplification to make it useful.

Much interest has recently been shown in passive mode locking techniques
employing external nonlinear cavities. The soliton laser21 was the first demonstration
of how an optical fibre coupled to a laser to form an external cavity could dramatically
influence the laser performance, resulting in a considerable reduction of the output
pulse duration. Subsequent research determined that soliton shaping in the external
fibre cavity was not a prerequisite for the process, only SPM induced spectral
broadening21 - 30. The process appears to occur through interferometric addition of
main cavity pulses with spectrally broadened pulses from the external nonlinear
cavity, and can result in pulse shortening through peak enhancement and pedestal
suppression if the cavities are interferometrically stable over a large number of
transits.The technique has been succesfully applied to a variety of systems from colour
centre lasers24 - 27 and Ti:Al2O3 lasers28,29, to Nd:YAG and Nd:Glass lasers30 - 32. One

drawback is the necessary interferometric cavity matching involved, requiring


complex stabilisation by piezo mounted mirrors.

In this section a technique for mode locking a Nd doped fibre laser is presented
which is novel in that it requires no direct modulation at the cavity round trip time, yet
can yield quasi continuous pulse trains of ~50ms duration with repetition rates
determined only by the relative cavity lengths. The technique was first reported in a
CW Ti:Al2O3 laser33 generating pulses as short as 6ps in trains of ~20ms duration,

with pulse repetition rates up to 1GHz. Mode locking is observed when one of the
mirrors in either the laser or the linear external cavity is moved with a velocity of the
order ~0.05m s-1. The term 'linear' refers to the lack of any nonlinear medium in the
external cavity, and as such the mechanism involved which results in pulse formation
has not yet fully been identified, although it is surmised that key elements include:

83
(a) sweeping the length mismatch between the two cavities (and therefore their
resonance) at a rate appropriate for Q-switching the laser,
(b) some nonlinearity in the laser amplifier (probably the optical Kerr effect)
providing pulse compression in a manner similar to Additive Pulse
Modelocking (APM)22,24.

In the experiment described here, a Nd doped fibre laser capable of


independently producing 1W average output power at 1.06μm is mode locked using a
linear external coupled cavity. The experimental schematic is shown in Fig.3.4.1.

Fig.3.4.1. Experimental arrangement of the mode locked Nd doped fibre laser.

A Nd3+ doped single mode fibre (0.5% Nd3+ by weight) of 37cm length
provided the gain medium, with the resonator formed by butt coupling two dielectric
coated mirrors to the end facets of the fibre. Mirror (M1) was 100μm thick, coated for

100% reflection around 1.06μm, and 95% transmission at the pump wavelength which
was 514nm. Mirror (M2) formed the laser output coupler and was ~96% reflective

around 1.06μm. Up to 2.5W of pump radiation was provided by a small frame CW


Argon Ion laser operating at 514nm. The fibre laser was pumped through cavity mirror
(M1) by focussing with a standard uncoated X10 microscope objective. Up to ~50% of

the available pump radiation was coupled into the fibre, which was highly multimode
at the pump wavelength.

The output emerging from the fibre laser through mirror (M2) was collected

and collimated with a X20 broadband ARCO lens, and coupled into the external cavity
via beamsplitter (BS) (97%R around 1.06μm). The external cavity was formed

84
between mirror (M2), beamsplitter (BS) and highly reflective mirror (M3), which was

mounted on an optical shaker unit (Bruel & Kjaer type 4810), which was of a similar
type to that found in a scanning autocorrelator.

The fibre laser had a pump power threshold of ~200mW, and exhibited 99%
pump absorption. With a pump power of approximately 1.25W coupled into the fibre,
an average output power of ~500mW was obtained at 1.06μm. In order to mode lock
the laser, the length of the external cavity (M2)-(BS)-(M3) was approximately matched
to that of the main fibre cavity (M1)-(M2), accounting for the refractive index of the
fibre. Mirror (M3) was then oscillated along the (BS)-(M3) axis, at peak velocities

between 0.03 and 0.18ms-1 to obain pulses whilst the mirror was in motion. The output
pulse train for an oscillation of +/-1.5mm peak to peak displacement and an oscillation
frequency of 38Hz is shown in Fig.3.4.2.

Fig.3.4.2. Oscilloscope traces of the laser output:


(a) Pulse train (5ms/div). (b) Mode locked pulses (200ps/div)

The signal in (a) was obtained with a fast diode and oscilloscope, and shows a
pulse train of ~13ms period, which corresponds to the oscillation period of the shaker.
The pulses were also monitored using a fast diode and sampling oscilloscope with an
overall temporal resolution of ~150ps. The shortest pulses measured were of 300ps
duration (not deconvolved with the 150ps combined temporal resolution) at a
repetition rate of 267MHz, as shown in (b). This repetition rate corresponded to the
inverse cavity round trip time for the Nd fibre cavity, assuming a refractive index of
~1.5. The deconvolved duration of the pulses was 260ps, substantially shorter than

85
those obtained in early experiments using acousto-optic modulators to mode lock Nd
fibre lasers.

A comprehensive characterisation of a Ti:Al2O3 laser mode locked using a

similar technique35 investigated how the pulse duration varied across the generated
pulse trains. The shortest pulses obtained were sampled using an electro-optic shutter
from the centre of the train, where the mirror moved fastest, and yielded pulse
durations approximately half of that integrated over the whole train. It seems
reasonable to asume that the same applies here, and hence pulses as short as 130ps
may be generated at the centre of the train. A study was made of the performance of
the laser as a function of the mismatch between the laser and external cavity lengths.
While mirror (M3) oscillated about an average position, the length of the external
cavity was varied by moving the mirror along its oscillatory axis on a translation
stage. Fig.3.4.3 shows how the measured pulse duration (deconvolved from the
measured 150ps photodiode-oscilloscope resolution) and pulse modulation depth
varied with cavity mismatch length. The modulation depth was defined as the peak to
background ratio of the mode locked pulse train.

Fig.3.4.3. Modulation depth and pulsewidth (deconvolved) vs mismatch between


cavity lengths.

Note that mode locking was observed over a much larger range of cavity
mismatch than that reported for lasers employing APM. As the external cavity length

86
deviated from optimum, the pulse length broadened and the modulation depth
decreased as the pedestal grew. This was a similar effect to that seen in Ti:Al2O3,

however no change in modulation depth was observed for that system. The
performance of the mode locked laser was relatively insensitive to small changes in
either amplitude or frequency of mirror oscillation, or to pump power.

The fibre laser operated on a number of spectral modes between 1.05μm and
1.07μm, both free running and when mode locked by the influence of the external
cavity. The relative intensity of these modes was extremely sensitive to cavity
alignment and fibre strain, whilst the temporal output remained consistently stable
over the resolvable timescales in comparison. Spectra obtained sequentially showed
large changes in the distribution of spectral energy between modes, and hence
conveyed little information as to the mode locking mechanism or laser performance.

To investigate the effect of stronger coupling between the cavities, mirror (M2)

was removed, increasing the output coupling from 4% (transmission) to 96% (residual
facet reflectivity ~4%) . Owing to the stronger coupling, the laser exhibited large
relaxation oscillations which could not be suppressed, even by reducing the
reflectivity of the beamsplitter (BS) or mirror (M3) to reduce the coupling to the

previous value. This suggested that a low loss (high Q) main laser cavity and perhaps
weak coupling between the cavities may have been important for effective mode
locking.

The velocity of mirror (M3) did not appear to be sufficient to directly mode

lock the laser by modulating the effective reflectivity of the external cavity, as the
interference effect would have induced a modulation of the cavity loss in the 100KHz
range. This timescale did, however seem appropriate for Q-switching the laser. The
velocity was also many orders of magnitude less than that required to lock the
longitudinal modes together through Doppler shifting, even if the effect built up over
hundreds of transits. By changing the output coupler of the main cavity (M2) to a 20%

87
transmissivity mirror, thereby increasing the coupling from what appeared stable, Q-
switched and mode locked operation could be observed, as shown in Fig.3.4.4(a) and
(b) on two different timescales.

Fig.3.4.4.Observation of simultaneous Q-switched & mode locked operation on a (a)


Q-switching and (b) mode locking timescale.

The Q-switched envelope duration was ~600ns, with the occurrence of further
relaxation oscillations at a period of ~10ms, which corresponds approximately to the
modulation frequency caused by interference between the two cavities at ~100KHz.

APM may have been responsible for the effects shown in this experiment. The
nonlinearity could have been generated in the fibre laser cavity instead of in the
external cavity, with peak power enhancement caused by massive relaxation
oscillations facilitating SPM. The nonlinear length LNL was less than 5m of fibre for

the intra-cavity mode locked peak power in this experiment, and since the laser had a
relatively high Q factor, it is feasible that some nonlinear broadening occured over a
number of transits. Relaxation oscillations could have been induced through
modulation of the reflectivity of the external cavity at frequencies comparable with the
upper state lifetime of the gain medium. This would explain why mirror oscillation
would only induce mode locking if the mirror velocity was within a given range, with
the exact amplitude or frequency set arbitrarily, since a particular velocity would give
rise to a specific modulation frequency. The minimum and maximum velocities were
0.03 and 0.18 ms-1 respectively.

88
Another mechanism36 which may be responsible for the mode locking
observed here is the influence of the linear external cavity on the initial stages of pulse
evolution in a synchronously pumped system. Any external perturbation can strongly
influence the initial pulse shaping in such a laser, through an independent evolution of
the radiation in each cavity, coupled with a beating between the longitudinal modes of
both.

The relatively modest repetition rate of 267MHz measured here could be


increased in future work by increasing the doping concentration in the Nd fibre and
using a much shorter length. It is possible to achieve laser action over only a few
centimetres of fibre37, and such a laser would be expected to mode lock quite well
using the technique described here. The relative oscillatory cavity mismatch need not
be derived from a shaking mirror, but could be developed electro-optically or by a
Piezo induced fibre strain. Alternatively in a similar Ti:Al2O3 system35, by making the

external cavity an integer fraction of the main cavity length, harmonic modelocking
was achieved, although the generated pulses were much longer and had significant
pedestals.

In conclusion, mode locking has been induced in a Nd doped fibre laser using a
novel technique where linear external feedback is fed back into the gain cavity with an
oscillating mirror. Similar work on Ti:Al2O3 suggests that the system may be scaled to

produce much higher repetition rates35. The technique may also work with other gain
media, and further work using Er doped fibre lasers may lead to the development of a
truly simple and compact source of high repetition rate pulses for use in transmission
systems operating in the nonlinear regime.

3.5 References

(1) Y.Kimura, K.Suzuki & M.Nakazawa. Elec.Lett. 25,1656(1989)


U U

89
(2) C.G.Atkins, J.F.Masicott, J.R.Armitage, R.Wyatt, B.J.Ainslie &
S.P.Craig-Ryan. Elec.Lett. 25,910(1989) U U

(3) R.Bachen. PhD Thesis. University of London (1990)


(4) L.F.Mollenauer, N.D.Vieira and L.Szeto. Opt.Lett. 7,414(1982) U U

(5) A.S.Gouveia-Neto, A.S.B.Sombra & J.R.Taylor. Opt.Comm. 68,139(1988) U U

(6) R.Wyatt. Elec.Lett. 25,1498(1989)


U U

(7) B.J.Ainslie, K.J.Blow, A.S.Gouveia-Neto, P.G.J.Wigley, A.S.B. Sombra


& J.R.Taylor. Elec.Lett. 26,186(1990) U U

(8) D.C.Hanna, A.Kazer, M.W.Phillips, D.P.Shepherd & P.J.Suni.


Elec.Lett. 25,95(1989)
U U

(9) J.D.Kafka, T.Baer & D.W.Hall. Opt.Lett. 14,1269(1989) U U

(10) R.I.Laming, S.B.Poole & E.J.Tarbox. Opt.Lett. 13,1084(1989) U U

(11) H.C.Lefevre. Elec.Lett. 16,778(1980)U U

(12) C.Lin, T.P.Lee and C.A.Burrus. Appl.Phys.Lett. 42,141(1983) U U

(13) J.F.Massicott, R.Wyatt, B.J.Ainslie & S.P.Craig-Ryan.


Elec.Lett. 26,1038(1990)
U U

(14) M.Nakazawa, K.Suzuki & Y.Kimura. Opt.Lett.15,88(1990) U U

(15) M.Nakazawa, K.Suzuki & Y.Kimura. Opt.Lett.15(1990)in press.


(16) J.B.Schlager, Y.Yamabashi, D.L.Franzen & R.I.Juneau.
IEEE Phot.Techn.Lett. 1,264(1989)
U U

(17) A.Takada & A.Miyazawa. Elec.Lett. 26,216(1990) U U

(18) K.Tai, A.Hasegawa & A.Tomita. Phys.Rev.Lett. 56,135(1986) U U

(19) E.J.Greer, D.M. Patrick, P.G.J.Wgley & J.R.Taylor.


Elec.Lett. 25,1246(1989)
U U

(20) H.Itoh, G.Davis & S.Sudo. Opt.Lett. 14,1368(1989) U U

(21) L.F.Mollenauer & R.M.Stolen. Opt.Lett. 9,13(1984) U U

(22) K.J.Blow & D.Wood. J.Opt.Soc.Am. B5,629(1988) U U

(23) M.Merin & M.Piche. Opt.Lett. 14,1119(1989) U U

(24) J.Mark, L.Y.Liu, K.L.Hall, H.A.Haus & E.P.Ippen. Opt.Lett. 14,48(1989) U U

(25) K.J.Blow & B.P.Nelson. Opt.Lett. 13,1026(1988) U U

(26) P.N.Kean, X.Zhu, D.W.Crust. R.S.Grant, N.Langford & W.Sibbett.


Opt.Lett. 14,39(1989)
U U

(27) J.F.Pinto, C.P.Yakymymshyn & C.R.Pollock. Opt.Lett. 13,383(1988) U U

(28) P.M.W.French, J.A.R.Williams & J.R.Taylor. Opt.Lett. 14,696(1989) U U

(29) J.Goodberlet, J.Wang, J.G.Fujimoto and P.A.Schulz.


Opt.Lett. 14,1125(1989)
U U

(30) J.Goodberlet, J.Jacobson, J.G.Fujimoto, P.A.Schulz & T.Y.Fan.


Opt.Lett. 15,54(1990)
U U

(31) L.Y.Liu, J.M.Huxley, E.P.Ippen & H.A.Haus. Opt.Lett. 15,553(1990) U U

(32) C.Spielman, F.Krausz, E.Wintner & A.J.Scmidt. Opt.Lett. 15,737(1990) U U

90
(33) P.M.W.French, S.M.J.Kelly & J.R.Taylor. To be published in Opt.Lett.
(34) I.P.Alcock, A.I.Ferguson, D.C.Hanna & A.C.Tropper.
Elec.Lett. 22,269(1986)
U U

(35) P.M.W.French, D.V.Noske, N.H.Rizvi, J.A.R.Williams & J.R.Taylor.


To be published in Opt.Comm.
(36) S.M.J.Kelly. Opt.Comm. 70,495(1989)
U U

(37) I.M.Jauncey, L.Reekie, J.E.Townsend & D.N.Payne. Elec.Lett. 24,24(1988)


U U

91
4. PULSE AMPLIFICATION IN DOPED FIBRES

4.1 Doped Fibre Amplification

Considerable interest has been placed in optical amplification in rare earth


doped fibres1 - 4, particularly Nd3+ and Er3+, for the development of highly efficient
tunable fibre lasers. Er doping is also of special relevance to fibre based
communications operating around the low loss region from 1.5-1.6μm. Good optical
confinement and direct compatibility with transmission media and laser diodes for
both signal and pump sources has resulted in a large demand for monolithic fibre
amplifiers. The ideal system would have a high saturation energy density, and long
fluorescence lifetime, enabling high average powers to be obtained. In this way,
pulsed amplification could yield large peak gains even at the highest repetition rates.
Solid state gain media offer these qualities, as shown in the last section up to 500mW
of average power was produce from a Nd doped fibre laser, and up to 60mW from an
Er doped fibre laser.

Amplification in Nd doped silica fibres is concentrated on the 1.06μm [4F3/2-


4I 4 4
11/2] transition since the transition at 1.32μm [ F3/2- I13/2] is quenched by an ion-host
interaction in silica glass, i.e. the existence of [4F3/2-4G7/2] excited state absorption8.

Flourozirconate glass has been found to be preferable as a host for operation at this
wavelength9, but since the loss is lower in silica fibre around 1.55μm, Er doped fibre
which exhibits higher gain has dominated research. Existing fibre optical transmission
systems only use a small fraction of the available transparent bandwidth of the fibre
over long distances, since up until now electronic circuits involving a detector,
amplifier and source have been used to regenerate attenuated signals after 20-100km.
To overcome this and other problems associated with high speed electronic
multiplexing, all optical amplification, signal conditioning and information processing
is a necessary requirement. In this way the signal remains optical throughout the
transmission system. Therefore highly efficient monolithic optical amplifiers are

92
required. Raman amplification has been extensively studied as an option, but there are
shortfalls involved, such as the limited maximum obtainable gain per unit length and
the Soliton Self-Frequency Shift (SSFS) (see chapter 1).

The transient response of such doped fibre amplifiers is extremely important.


To avoid SSFS in the transmission line, long signal pulses (>25ps) are normally used
in proposed long haul transmission systems. Although single channel transmission
with pulses of this duration only requires a limited gain bandwidth, the optical
information processing hardware at each end of the transmission line will probably
operate with much shorter pulses over shorter distances. It is hoped that in the future,
the ultimate duration of the 'bits' used in the transmission line will be much shorter,
allowing even higher transmission rates. Ultrashort pulse amplification is considered
here in both Nd and Er doped fibres with picosecond pulse amplification at 1.06μm
and femtosecond pulse amplification at 1.53μm reported for the first time.

4.2 Picosecond Amplification in Neodymium Doped Fibre.

Efficient pump confinement and good mode overlap with pump radiation in the
core results in very high available gain in optical fibres. The doping levels used here
resulted in almost complete pump absorption occurring over the few metre length of
doped fibre, although this is not necessary in Nd since its four level system has no
ground state absorption at the gain wavelength. The purpose of this experiment was to
establish the transient gain available in a typical fibre amplifier based on Nd doped
fibre. Neodymium in a glass host has a 3dB fluorescence bandwidth of more than
10nm around 1.06μm, and is hence capable of pulsed amplification of less than ~200fs
sech2 pulses. Here the pulses amplified were 4ps in duration, generated using the well
known fibre grating compression technique described in chapter 1. A schematic of the
experimental arrangement is show in Fig.4.2.1.

93
Fig.4.2.1. Schematic of the experimental arrangement for pulse amplification in Nd
doped fibre.

A Nd:YAG laser operating at 1.06μm, generating 100ps pulses at 100MHz


repetition rate with an average output power of 5W was used to derive the signal. The
output pulses of this laser were continually monitored with a synchroscan streak
camera. These 100ps pulses were then compressed to ~4ps with a standard fibre
grating compressor10. The average output power of the compressed pulse train was
200-250mW, for 1W of launched average power in the compressor fibre. The fibre
grating compressor incorporated 200m of single mode non-polarisation preserving
fibre, and a dispersive delay line consisting of two parallel 1800line/mm holographic
gratings with 90% diffraction efficiency into the -1 order at a a 75o angle of incidence.
A total grating separation of approximately 1m gave the shortest pulses, which were
continually monitored using a collinear autocorrelator (real time with background, see
appendix). The signal from the Nd:Yag laser was coupled into the compressor fibre
with a standard microscope objective.

Fig.4.2.2.(a) Spectral and (b) Temporal (autocorrelation) characteristics of the pulse


from the fibre grating compressor used as a probe signal in the Nd doped fibre
amplification studies.

94
Fig 4.2.2(a) shows the spectal output of the compressor fibre, showing SPM
broadening to a bandwidth of 0.8nm. This was capable of supporting ~2ps pulses
assuming a Gaussian pulse shape. Fig.4.2.2(b) is the corresponding collinear
autocorrelation of the compressed pulses emerging from the double pass grating
arrangement, with a measured deconvolved pulse duration of ~4ps. This yielded for
200-250mW of average power a peak power of ~500W, and a corresponding energy of
~2nJ per pulse.

The output signal pulses from the compressor were coupled via a variable
neutral density wheel (ND filter) into a conventional GeO2-SiO2 fibre with an 8μm

core, doped with 300ppm Neodymium, with a medium phosphorous content (1-
2%mol% P2O5).This single mode fibre had a loss of ~2dBkm-1 at 1.06μm. At 514nm

however, the fibre became highly multimode, exhibiting an absorption coefficient of


α=1.46X10- 2cm-1. A 2.5m length of Nd fibre comprised the amplifier in a counter-
propagating pump geometry. The pump excitation was provided from a small frame
Argon Ion laser operating at 514nm, with a maximum of 40mW of radiation coupled
into the signal exit end of the Nd fibre with an uncoated X20 microscope oblective
(MO2). The signal from the compressor was coupled into the fibre with an identical
objective (MO1), and separated from the counter-propagating pump at the fibre exit
with a dichroic beamsplitter (BS3) which transmitted all of the 514nm pump and

reflected the amplified signal at 1.06μm.

Fig.4.2.3. Fluorescence gain profile of the Nd doped fibre around 1μm.


Initially the fluorescence spectrum of the Nd fibre was measured for and

95
average launched pump power of ~10mW at 514nm. The fluorecence spectrum around
1μm is shown in Fig.4.2.3. A gain cross section of Nd in glass at 1.06μm of σ=3X10-
20cm2 yields a saturation energy fluence (energy density) of:

∅sat = h ν = hc ≈ 3Jcm-1
2 σ 2σλ

where h is Plack's constant, and λ=1.06μm. Gain saturation results from transitions
between the upper and lower laser levels due to stimulated emmission. This lowers the
population inversion and hence the gain if enough stimulated emmission takes place.
Gain saturation is only significant if more energy than that stored in the gain medium
is extracted within a time interval comparable with the upper state lifetime, which can
be measured. The fluorescence lifetime of the 4F3/2-4I11/2 transition at 1.06μm in the Nd

doped fibre used here was measured by pumping a short length (~20cm) of identical
fibre with ~150ns pulses at a 2KHz repetition rate. These ~3kW pulses were derived
fom a Q-switched, frequency doubled Nd:YAG laser operating at 532nm. The
fluorescence at 1.06μm was separated from the pump band and the fluorescence signal
at 900nm using a monochromator. The signal was detected with a vacuum photodiode
with <5ns response time, and is shown in Fig.4.2.4.

Fig.4.2.4. Fluorescence decay profile of Nd doped fibre at 1.06μm.


The trace indicated an upper state relaxation time of ~430μs. Saturation of the
gain in the fibre amplifier was expected to occur when the signal energy approached

96
the saturation fluence within the gain recovery time. The saturation fluence of 3Jcm-2
corresponded to an average saturation energy of 1.5μJ for the fibre used here, which
had a mode field area of ~50μm2.Since the fundamental transverse mode occupied by
the signal had a questionable overlap with that occupied by the pump, which occupied
a hybrid set of modes, the pump and signal alignment were continuously optimised
throughout the experiment, to maximise the gain experienced.

To attenuate the relatively large average signal power available from the
compressor a neutral density wheel (ND filter) was inserted into the signal path prior
to amplification. A conventional collinear autocorrelator with <100fs resolution and a
synchroscan streak camera was used to directly measure the gain of the amplifier, and
to monitor changes in the amplified pulse shape. Throughout the investigation no
change in the duration of the pulse transmitted through the amplifier was measured at
the power levels used, indicating that the gain bandwidth of the Nd doped fibre
exceeded that of the signal pulse, which was ~1nm. Fig.4.2.5. shows an
autocorrelation trace of the (a) unamplified signal pulse for an average launched signal
power of 4mW and (b) the resultant amplified signal for a launched pump power of
40mW at 514nm. Both traces were recorded on the same vertical scale.

Fig.4.2.5. Autocorrelation traces showing signal at 1.06μm (a) without pump, and (b)
with 40mW of counter-propagating pump radiation at 514nm.

A measure of the gain could be deduced from the ratio of the nonlinear

97
autocorrelation intensities, but was in fact measured more accurately directly using a
synchroscan streak camera. The gain was measured for a fixed maximum launched
signal power of 25mW (corresponding to ~250pJ per pulse) as a function of launched
pump power. The resultant curve is shown in Fig.4.2.6.

Fig.4.2.6.Variation of peak gain with pump power for a constant average signal
power of 25mW.

The maximum peak gain experienced for this signal power was ~1.2 for a
launched average pump power of between 15 and 40mW. Saturation of the gain
readily occurred, due to the high signal intensity injected. Higher gains could be
observed by reducing the signal level by an order of magnitude. To investigate the
gain as a function of signal intensity, the average pump power was maintained at a
maximum level of 40mW, and the signal level (measured by transmission in the
absence of any pump radiation) was adjusted using the ND filter. The resultant trend is
shown in Fig.4.2.7.

Fig.4.2.7. Variation of peak gain at 1.06μm with average signal power for 2.5m of Nd
doped fibre pumped with 40mW of average power at 514nm.

98
For signal pulses of ~4ps at 100MHz repetition rate, a saturated gain of X1.5
was measured for an average signal power of 2.5mW. For signal powers below
~0.2mW the gain increased rapidly from X2 to X10 for 15mW of launched signal
power (0.15pJ per pulse). The observed gain saturation effect is in reasonable
agreement with that predicted for Nd:Glass, since a 2.5mW average signal power
constitutes approximately 1.5μJ of energy in 600μs, which is of the same order as the
measured fluorescence lifetime. The measured gain was substantially lower than the
small signal gain for signals above the 100μW level, hence the saturation behaviour is
attributed to the high signal powers used.

In conclusion, up to X10 gain for 4ps pulses with energies of ~0.2pJ per pulse
(~20μW average) at 100MHz repetition rate has been observed in a Nd doped fibre
amplifier, for 40mW of pump power at 514nm. The importance of gain saturation and
the requirement of a high saturation fluence and long gain lifetime are paramount for
pulsed amplification. Operation at 1.06μm in Nd doped fibre is however of little
consequence to optical communications. Results more relevant to optical
communications are presented in the next section, which deals with Erbium
amplification operating around the 1.55μm region of lowest loss in silica fibres.

4.3 Femtosecond Amplification in Erbium Doped Fibre.

Fig.1.3.6. in chapter 1shows a plot of the measured absorption profile over the
region of highest transparency from 900-1700nm for a Dispersion-Shifted Fibre, with
λ =1.56μm. By careful control of the OH- impurity content during manufacture, the
0

absorption loss can realistically be reduced to less than 0.3dBkm-1 over lengths
greater than 20km. Absorption loss has a broadening effect on solitons, since as the
pulse energy and peak power decreases, the duration must increase to maintain the
fundamental soliton requirement (see chapter 1). High order solitons degenerate into
lower orders through loss a similar manner, unless break up through SSFS occurs first.
Soliton propagation was predicted early in the 1970's, but fibre technology had to

99
significantly advance for losses to be <1dBkm-1 before soliton effects could be
experimentally observed within an absorption length.

A long distance soliton based transmission system requires amplification to


ensure that pulse broadening through fibre loss does not dominate. To overcome
absorption, which can be as low as 0.2dBkm-1 in a state of the art fibre, 0.2dBkm-1 of
gain is required. For a typical transmission span equal to the loss length α-1~21km, the
total loss and hence total lumped gain required would be 4.2dB.This magnitude of
gain can easily be provide by an Erbium doped fibre amplifier, although the repetition
rate of the pulse train undergoing amplification places an upper limit on the repeater
spacing required due to gain saturation. Gain saturation readily occurs at relatively
low average powers for solid state amplification of high repetition rate signals.

Erbium amplification has been extensively studied for use in optical


transmission6,11 - 16 and 30dB of gain over a bandwidth in excess of more than 10nm
FWHM has been demonstrated at 1550nm7. This wavelength also coincides with the
emission wavelength of many common InGaAsP laser diode, enabling Er gain to be
used to amplify small amplitude modulated pulse trains up to soliton power levels.
Laser diode pumping of an Er fibre amplifier has also been achieved, enabling
convenient and efficient pumping at 1480nm17.

The full gain bandwidth (5dB) of the best Er doped fibre is in excess of
25nm18, and is therefore capable of amplifying pulses as short as 100fs (assuming
sech2 shape). In the experiment presented here, ~200 fs soliton pulses were amplified
in a single mode Er doped fibre for the first time.

A single mode fibre with an Erbium concentration of 200ppm, a core diameter


of 4.56μm and an outer diameter of 115μm was used. The fibre was 10m in length,
with a core-cladding index difference Δn=0.014. The spectral fluorescence profile of
the fibre was measured for excitation at 514nm, and is shown in Fig.4.3.2.

100
Fig.4.3.2. Fluorescence spectrum of Er3+ doped fibre. The bandwidth of a 200fs
soliton is shown as a bar for reference.

The peak of the fluorescence appeared around 1.53μm, with a secondary peak
at 1.55μm. For reference, the bandwidth required to support or amplify a 200fs soliton
is shown as a bar for reference. Although the spectrum was measured with a low pump
power to ensure that laser action did not occur from the fibre ends, the exact
fluorescence lineshape is strongly dependent on dopant level, fibre length, overall
absorption an pump power19. Thus the gain bandwidth for a specific pump power
depends on the fibre parameters. The energy level splitting of the Er3+ ions in the
random glass host crystal field results in a very broad 1.55μm [4I15/2 - 4I13/2] transition.

Two fluorescence peaks normally exist around 1535 and 1552nm, with the
1552nm peak only dominant for high pump powers. The fluorescence spectrum shown
in Fig.4.3.2. is perhaps more representative of amplified spontaneous emission, and
probably shows less FWHM bandwidth than that available for amplification. For even
higher pump powers, the gain spectrum normally encompasses both peaks, resulting in
>20nm gain bandwidth18.

101
Fig.4.3.3. Schematic of the cascade soliton Raman ring laser.

The signal source use in this experiment was an optimised cascade Raman
soliton laser20, which was synchronously pumped by a mode locked Nd:YAG laser
operating at 1.32μm. The schematic of the soliton Raman laser is shown in Fig.4.3.3.
The fibre used in the ring laser was 600m long, with a dispersion zero wavelength
λ0=1.46μm, and hence was anomalously dispersive only at the second Stokes Raman

band wavelength around 1.5μm for the 1.32 μm pump. The principle behind Raman
ring oscillators was introduced in chapter 2, and in this case, the disperion
compensation arises through soliton shaping from anomalous dispersion in the fibre.
More details of the processes involved in Raman soliton generation19 are given in
chapter 6.

Fig 4.3.4. shows spectra of the output of the soliton Raman ring laser. The
spectrum in Fig 4.3.4(a) corresponded to single pass Raman generation, and indicates
only the first Stokes generation at around 1.41μm for the power levels being used.
(The narrow line around 1.34μm was attributed to Raman gain induced in a weak
second line from the Nd:YAG laser.). Fig 4.3.4(b) show the fibre output spectrum in a
resonator configuration, optimised for operation at 1.5μm, in the second Stokes
Raman band. Third Stokes generation was apparent, simultaneously occurring with the

102
second Stokes. By fixing the pump power at around 400mW average, the ring laser
provided in excess of 150mW average output power in a broad band around 1.5μm.
Within this broad band several single solitons existed, the number and temporal
separation of which were not particularly controllable, as a result of the soliton Raman
continuum evolution from noise.

Fig.4.3.4. Spectra from the cacade soliton Raman ring laser for an average pump
power of 500mW at 1.32μm, (a) without and (b) with feedback.

Intensity autocorrelation of the ring laser output was the integration of all of
the pulses, which were of an average duration of ~200fs. By spectrally selecting a
portion of the spectral continuum, near transform limited output was obtained. In this
case, spectral selection around 1.53μm derived a source of solitons with ~200fs
duration and ~80mW average power, at the fluorescence peak of the Erbium fibre (see
Fig.4.3.2.). This spectrally selected signal was used as a signal for Erbium
amplification.

A scanning autocorrelator with <50fs temporal resolution was used to


continuously monitor the signal pulses from the Raman ring laser. Fig.4.3.5. shows a

103
background free (non-collinear) autocorrelation of the signal from the ring laser,
spectrally selected around 1.53μm, indicating an overall duration of 200fs. The
Erbium doped fibre was pumped in a counter-propagating pump geometry similar to
that used for Nd doped fibre in the last section, with 1W of CW Argon Ion pump
radiation at 514nm available. Up to 300mW of average pump power at 514nm was
coupled into the exit end of the fibre.

Fig.4.3.5. Background free autocorrelation trace of 1.53μm soliton pulses from


Raman ring laser.

The pedestal component in the background free autocorrelation was less than
half a percent of the total intensity, indicating that in excess of half the average power
was contained in the soliton pulses. A corresponding transform limited 200fs soliton
pulse would have a spectral bandwidth of 12.3nm, as indicated in Fig.4.3.2. as a
reference bar. It is however important to note that the signal pulses used here have
bandwidths in excess of twice this value, as more than one soliton is present within an
envelope of spectrally separated signals. The complete spectrum, including all solitons
present, contributes to the autocorrelation envelope.

These signal solitons were simultaneously launched into the Er doped fibre,
and amplified on application of the counter-propagating pump at 514nm. Fig.4.3.6.
shows the autocorrelation trace of the pulses emerging from the Er doped fibre.

104
Fig.4.3.6. Autocorrelation trace of the (a) unamplified and (b) amplified soliton
pulses. Both vertical scales are the same, but zero levels differ for clarity.

For the trace in Fig.4.3.6(a), the pump power at 514nm was adjusted for
lossless propagation of the soliton signal, determined by measurement of the average
signal power at 1.53μm. This step was necessary because the three-level laser system
in Erbium would have resulted in complete signal absorption of the signal from the
ground state in the abscence of pump radiation over this fibre length. In Fig.4.3.6(b).
the launched pump power was increased to 250mW at 514nm. A clear increase in
signal intensity was evident with no increase whatsoever in the soliton pulsewidth
from 200fs.

To measure the degree of gain saturation occurring in the amplifier, the gain
was measured for various average signal powers, using a variable neutral density
wheel to attenuate the soliton signal power from the ring laser. The pump power was
fixed at 300mW in the 10m length of fibre during this process and the resultant trend
in the gain is shown in Fig. 4.3.7.

105
Fig.4.3.7. Gain vs average soliton signal power for a fixed pump power of 300mW.
For signal powers above ~5mW average, severe saturation of the gain was
evident, as shown in Fig.4.3.7., with the gain falling to almost unity for 50mW of
signal (corresponding to ~2.5kW launched peak power). For signals of about 1mW
average power, a gain of 2.6 was exhibited, corresponding to a net gain of ~4.2dB.

It was necessary throughout the experiment to operate with signal power levels
of the 1mW level, since the fundamental soliton power in the Er fibre was ~22.5W
peak (0.45mW average). Below this power level, significant dipersive broadening
dominated, resulting in broader measured pulse durations in the absence of gain. The
spectral bandwidth of the Raman soliton laser signal was as large as ~30nm (limited
only by the spectral filtering used beyond the ring laser). It is therefore possible that a
higher gain was present, since the measured bandwidth of the amplifier was less than a
half of that of the signal at the pump powers used here. If bandwidth limited gain was
operating, however, some temporal reshaping of the amplified pulses would be
expected, and no such effect was observed, within the ~50fs resolution acheivable. For
high signal levels, the low gain measured was a result of amplifier saturation, similar
to that recorded for picosecond amplification in Nd doped fibre in the last section.

A ~4.2dB gain has been observed in an Er doped fibre amplifier for 200fs
soliton pulses at a ~1mW average power level (~50Wpeak). Similar gains have also
been reported when operating the system with 120fs signal pulses. The low gains
achievable were a result of the relatively large bandwidth spread of the input signal,

106
and amplifier saturation at higher signal powers. No temporal dispersive broadening
was evident in the amplified soliton signal. More gain (up to ~20dB) has been
measured for smaller signal powers, however significant temporal broadening was
observed below an average power of ~0.5mw, which corresponds to the fundamental
soliton power in the Er fibre for a 200fs soliton pulse.

The principle involved in this experiment could be extended for incorporation


in non-dispersive reamplification of attenuated pulses in all optical communications.
Soliton shaping could enhance the amplification process and enable soliton
regeneration from sub-fundamental soliton or dispersive pulses. This would require
the amplificaton process to occur over a length scale much longer or shorter than the
soliton period, since non-adiabatic amplification over the soliton period could lead to
significant pulse distortion occurring.

4.4 References

(1) M.C.Brierley, P.W.France & G.A.Miller. Elec.Lett. 24,540(1988) U U

(2) M.C.Farries, P.R.Morkel &J.E.Townsend. Elec.Lett. 24,709(1988) U U

(3) D.C.Hanna, R.M.Percival, I.R.Perry, R.G.Smart, P.J.Suni, J.E.Townsend &


A.C.Tropper. Elec.Lett. 24,111(1988)
U U

(4) D.C.Hanna, R.M.Percival, I.R.Perry, R.G.Smart, P.J.Suni, J.E.Townsend &


A.C.Tropper. Elec.Lett. 24,1223(1988)
U U

(5) R.J.Mears, L.Reekie, I.M.Jauncey & D.N.Payne. Elec.Lett. 23,1026(1987) U U

(6) E.Desurvire, J.R.Simpson & P.C.Becker. Opt.Lett. 12,888(1987) U U

(7) C.R.Giles, E.Desurvire, J.R.Talman, J.R.Simpson & P.C. Becker.


CLEO'88 PD9. OSA technical digest Vol7,473(1988) U U

(8) I.P.Alcock, A.I.Ferguson, D.C.Hanna & A.C.Tropper. Opt.Lett.11,709(1986) U U

(9) M.C.Brierley, C.A.Miller. Elect.Lett.24,438(1988)


U U

(10) A.S.L.Gomes, U.Osterberg & J.R.Taylor. Optr.Comm. 54,377(1985) U U

(11) R.J.Mears, L.Reekie, S.B.Poole * D.N.Payne. Elec.Lett. 21,738(1985) U U

(12) S.B.Poole, D.N.Payne & M.E.Fermann, Elec.Lett. 21,737(1985) U U

(13) E.Snitzer, H.Po, I.Hakimi, R.Tuminelli & B.C.Macollum.


Opt.Fibre.Comm.Conference. Anaheim, USA. PD2(1988)
(14) Y.Kimura, K.Suzuki & M.Nakazawa. Elec.Lett. 25,1657(1989)U U

(15) K.Suzuki, Y.Kimura and M.Nakazawa. Appl.Phys.Lett. 55,2573(1989) U U

107
(16) M.Nakazawa, K.Suzuki & Y.Kimura. IEEE.Phot.Tech.Lett. 2,216(1990)
U U

(17) M.Nakazawa, Y.Kimura, & K.Suzuki. Elec.Lett. 25,199(1989)


U U

(18) T.J.Whitley & T.G.Hodgkinson. 14th Europ.Conf.on Opt.Comm.


Brighton, UK. IEEE Conf.Publ. 292,1,58-60.
U U

(19) Y.Kimura & M.Nakazawa. J.Appl.Phys. 64,516(1988)


U U

(20) A.S.Gouvei-Neto, A.S.L.Gomes, J.R.Taylor, B.J.Ainslie & S.P.Craig.


Opt.Lett. 12,927(1987)
U U

108
5. MODULATIONAL INSTABILITY

AND CROSS PHASE MODULATION

5.1 Modulational Instability

In Eq(1.4.13) of chapter 1, the Nonlinear Schroedinger Equation (NLS) was


introduced:

ŽA β 2 Ž2 A
+ γ A A = - αA
2
i - 2
ŽZ 2 ŽT 2
(5.1.1)

where a loss term has been included from Eq(1.4.33) on the R.H.S. For the special
case where the loss coefficient α=0, this equation has been shown to support solitary
wave solutions of hyperbolic secant envelope pulses, whose shape does not change
with propagation provided the conditions set in Eq(1.4.27) are satisfied, i.e. that the
dispersion length equals the nonlinear length.

An interesting alternative exists to ordinary soliton shaping operating on pulses


in the anomalous dispersion regime. This is found in the scenario where CW or quasi-
CW fields are propagated. The situation is comparable32 to launching pulses of such a
long duration that the fundamental soliton power in Eq(1.4.28) is very low, and hence
a very high order soliton is excited, with an extremely long soliton period. If we
assume initially that the field A(0,T) is CW, or completely independent of time,
Eq(5.1.1) can be solved to obtain:

A(Z,T) = P0 ei γ P 0Z
(5.1.2)

This time independent solution is consistent with CW light propagating


unperturbed (except for attenuation if loss is included), apart from acquiring a power
dependent phase shift as it propagates. This intuitive behaviour represents the steady

109
state solution, but a more important effect becomes apparent if small perturbations are
considered. If a perurbation in amplitude a(Z,T) is imposed on A(Z,T) the propagation
can be evaluated by substitution into Eq(5.1.1) with α=0 to obtain:

Ža β 2 Ž2 a
i = +γ P 0 (a+a* )
ŽZ 2 ŽT2
(5.1.3)

If a general solution is assumed1:

a(Z,T) = a1 cos (kZ-ΩT) + i a2 sin(kZ-ΩT)


(5.1.4)

with k and Ω representing the wavenumber and frequency of the perturbation


respectively, Eq(5.1.3) yields a non-trivial solution provided the following dispersion
relationship between k and Ω is satisfied:

2 1/2
k = ± 1 β 2 Ω Ω + sgn β 2 Ωc
2
2 (5.1.5)

2 4 γ P0 4
where Ωc = =
β2 β 2 LNL
(5.1.6.)

and sgn(β2)=+/-1, depending on the sign of the dispersion parameter β2. The nonlinear

length has already been defined in chapter 1 as:

LNL = 1
γP 0 (5.1.7)

and the magnitude of Ωc is therefore dependent on power and dispersion magnitude.


For normal GVD (β2>0), k remains real, and the solution generally remains stable.
However for anomalous GVD (β2<0), k becomes imaginary for Ω<Ωc, and an

imaginary k results in an exponential relationship between the perturbation magnitude


and the propagation length Z from Eq(5.1.4). This solution is therefore inherently

110
unstable for anomalous GVD, and in fact results in an exponential growth of the
perturbation a(Z,T) with propagation length. The process is referred to as
Modulational Instability (MI), resulting for a sufficiently intense pump field at ω0 in:

(i) spontaneous modulation growth from the steady state for the autonomous CW
field in the presence of noise fluctuations, and
(ii) growth of any seed modulation imposed on a CW carrier at ωo, provided the
seed modulation frequency ωs<Ωc given in Eq(5.1.6).

Modulational instability has been found to occur in other systems such as fluid
dynamics and plasma physics, and is commonly referred to as the self-pulsing or
Benjamin-Feir instability2. The conditions necessary to observe MI are similar to those
required to observe soliton propagation in fibres. An interaction between anomalous
group velocity dispersion and SPM can provide exponential gain with length for any
amplitude or phase fluctuations induced on a CW or quasi-CW carrier.

The gain is frequency dependent, and is given by considering the complex


dispersion relation in Eq(5.1.5). Since

a(Z,T) = (a1 +a2 )2 e i kZ-Ω T (5.1.8)


the gain is given by the magnitude of the imaginary part of the complex wavenumber,
k. The gain at frequency ωo+Ω is given by:

2 2 1/2
g Ω = 2Im(k) = β 2 Ω Ωc - Ω

= 2 γ P0 = 2
LNL (5.1.9)

This implies that any amplitude and/or phase modulated signal becomes
unstable for modulation frequencies Ω<Ωc. Fig.5.1.1. is a graph of the spectral

dependence of the gain for three peak power levels (1W,2W & 4W) with parameters
β2=-20ps2km-1 and γ=2W-1km-1, which are appropriate for a standard silica fibre at

1.55μm.

111
Fig.5.1.1. Frequency dependence of MI gain for three peak power levels, with
β2=-20ps2km-1 and γ=2W-1km-1.

The gain reaches a peak for two equidistant sideband frequencies +/-Ωmax,

where

2 γ P0 1/2
Ωmax = ± Ωc = ±
2 β2 (5.1.10)

and has a magnitude which is linearly power dependent through:

2
gmax = g Ωmax = 1 β 2 Ωc = 2γ P 0
2 (5.1.11)

If fibre loss becomes significant, the gain is offset, resulting in a smaller


integrated growth of the modulation. The frequency range over which the modulation
can grow is typically 0.1-1THz, corresponding to the growth of spectral sidebands
separated by ~1-10nm in the low loss region around 1.5μm. For the effect to become
measurable, the power must be sufficently high for the gain to overcome fibre loss,
which requires peak powers of typically ~100mW over a fibre loss length. However as
modulational frequencies are very high, detection through autocorrelation is normally
used over much shorter fibre lengths. Typical peak powers required to observe the
effect through autocorrelation are therefore ~1W or above.

112
It should be noted that the bandwidth of a typical 300fs sech2 pulse at 1.5μm is
approximately 8nm. To exhibit MI the modulation period would have to be of the
order of 30fs, which would require extremely high powers, and hence ultrashort pulses
do not normally exhibit MI at moderate peak powers. The process is best observed
with CW pump signals of high intensity. As the pump power is increased, the gain
peak moves further from the carrier frequency, and higher harmonic modulation peaks
can develop at integer multiples of the fundamental modulation frequency. The
process eventually saturates when the CW pump signal is sufficiently depleted to
reduce the peak power at the carrier frequency. Almost 100% pump depletion is
possible.

MI is the lowest threshold nonlinear effect to be observed, if Brillouin


scattering is neglected. Considerable attention has been paid to the process of induced
modulational instability, in which a pump field induces spectral broadening and MI on
a signal at another wavelength through Cross Phase Modulation (XPM). The process
is decribed in more detail in section 5.3. The resultant XPM induced broadening seeds
MI in the signal, even if the signal is itself not sufficiently intense enough to induce
MI on its own. The pump field is in effect providing a seed amplitude or phase
fluctuation in the weak signal, and the threshold for the MI process is hence
considerably lowered.

Modulational instability can be explained in terms of a degenerate Four Wave


Mixing (FWM) interaction, in which two carrier pump photons decay to create one
Stokes and one anti-Stokes photon within the MI sidebands, such that ωstokes= ω0-Ω
and ωanti-stokes= ω0+Ω. The efficiency of the FWM process is highly dependent on phase

matching, which is only achieved within the MI gain sidebands. The phase matching is
provided through anomalous GVD operating on the SPM broadened signal. From the
FWM perspective, it is easier to explain how the process can also be induced by
injection of a probe signal within the gain sideband of a CW carrier. If a probe at

113
frequency ω1=ω0+Ω, copropagates with a CW carrier signal which is of insufficent

intensity to produce spontaneous MI, the probe may experience an exponential


growth given by Eq(5.1.9), provided Ω<Ωc. The process causes a similar effect as

spontaneous MI3,4, however the frequency shift of the sideband is no longer power
dependent, and is fixed at the probe frequency. This process is referred to as FWM or
probe induced MI.

The observed effect of MI on an intense carrier is the growth of a sinusoidal


modulation on the CW or quasi-CW carrier
field, irrespective of how the process is
initiated. This modulation can be observed best
using autocorrelation techniques, as the
frequencies are typically very high. The
modulation sidebands can be observed
spectrally, with the sideband separation
corresponding exactly to the inverse
modulation frequency. A typical MI signal can
be seen both temporally and spectrally in
Fig.5.1.2.
Fig.5.1.2.Modulational instability
observed for 100ps pulses at
1.32µm in 500m of fibre, both The MI shown in Fig.5.1.2. was
spectrally and temporally.
produced by propagating 100ps pulses at
1.32μm into 500m of fibre with a peak power
of 6W (60mW average). The autocorrelation (inset) indicated a modulation period of
450fs (corresponding to a ~2THz modulation frequency). Brillouin scattering was
effectively suppressed for the 100ps pump pulses and did not therefore dominate. MI
is a fundamental problem for the propagation of long pulses in the soliton regime, if
the peak power is sufficient to provide gain at a modulation period less than the pulse
duration.

114
The gain from modulational instability can be used as an amplifier, since a
probe signal within the MI sideband of an intense carrier will exponentially grow at
the expense of the carrier. The sidebands produced through MI can also become
amplified through Raman scattering, and act as a precursor to the production of
Raman solitons, significantly reducing the threshold for the process. This is dicussed
in chapter 6. Since gain saturation through carrier depletion eventually slows down the
growth MI sidebands, the generated modulations can then undergo soliton shaping,
resulting in the generation of a train of solitary waves, which when amplified can grow
into many single solitons.

Since modulational instability can be induced through XPM, the pump field
which induces nonlinear broadening in the relatively weak signal need not propagate
in the anomalous GVD regime. Coupling through XPM is independent of the sign of
dispersion. In this way the peak power of an intense pump signal, regardless of its
dispersive nature, can provide the nonlinearity necessesary for a weak signal
copropagating in the anomalous GVD regime to undergo soliton shaping. Sub-
fundamental soliton signals can undergo nonlinear pulse shaping, provided they are in
the anomalous GVD regime, and that a strong pump is present, copropagating to
provide XPM. In this chapter a series of experiments are presented which investigate
MI, induced MI through XPM and FWM, and XPM induced pulse shaping in the
anomalous GVD regime.

5.2 Modulational Instability from a Mode Locked

Erbium Fibre Laser Source.

Experimental observation of MI requires propagation of an intense CW or


quasi-CW field on a timescale of the expected modulation. That illustrated in
Fig.5.1.2. used ~100ps pulse from a Nd:YAG at 1.32μm laser to provide a high peak
power of ~6W, and similar techniques have been applied elsewhere2. The effect has
more recently been obtained with a pure CW pump field from a Nd:YAG source5,

115
despite significant Brillouin depletion of the relatively narrow band source. Such
pump sources are suitable due to a high output power, and a wavelength which is close
to the zero dispersion wavelength of standard fibres at ~1.3μm.

The extremely low absorption loss in optical fibres around 1.55μm has meant
that much interest has been placed in Erbium doped fibre amplification, as discussed
in chapter 4. The medium is directly compatible with soliton systems, with improved
performance over long distances and at higher bit rates. In order to generate ultra-high
repetition rate pulse trains around this wavelength, nonlinear effects such as MI offer a
means of generating pulses at up to 5THz6. By shifting the zero dispersion wavelength
λ0 of the fibre to around 1.55μm, it should be possible to reduce the anomalous GVD

sufficiently to generate extremely high rates, in accordance with Eq.(5.1.6). However,


although as the dispersion decreases the modulation frequency theoretically
approaches infinity, higher order dispersion terms have been found to dominate
around λ0, resulting in a levelling off of the modulation frequency5. Stimulated

Brillouin Scattering (SBS) presents a dominant loss mechanism for intense CW beams
whose bandwidths are much less than one Brillouin linewidth (~10GHz in silica). In
fact reflectivities of 50% are typical5. Pulsed systems with large bandwidths are
therefore preferred for generating MI.

This chapter describes an experiment to determine the power level required to


generate MI at the wavelength of an Erbium fibre laser. This was used to determine
the likelihood of obtaining continuous MI within an Erbium doped fibre laser cavity,
to produce a source of pulses from MI. Since discernable MI would have to dominate
over other nonlinear effects, a mode locked pump source was used. The laser had an
output bandwidth in excess of 2GHz, which reduced the SBS gain sufficiently as to
make it unmeasurable. A fibre with a small effective core diameter reduced the power
requirement for MI by increasing the intensity, and reducing the effective GVD by
dispersion shifting λ0 to around 1.55μm.

116
The source of radiation used in this experiment was decribed in detail in
section 3.2. This Er fibre laser was mode locked with an acousto-optic modulator, and
produced pulses as short as 140ps at a repetition rate of 100MHz, with an average
power output of 60mW. The output was tunable from 1530 to 1580nm, with short
pulses generated over the range 1555 to 1575nm. Outside this range, the reflectivity of
the output coupler fell below ~ 15%, and the laser did not exhibit mode locking. The
experimental schematic is shown in Fig.5.2.1.

Fig.5.2.1. Schematic of experimental arrangement for generating MI from a mode


locked Er doped fibre laser.

Essentially, the output pulses from the laser were focussed into 1.2km of
Dispersion-Shifted Fibre (DSF) via an isolator to avoid reflections from the fibre or
lenses destroying the laser mode locking. The laser was tuned by alignment of the
intra-cavity grating to the wavelength region of low anomalous dispersion in the fibre,
close to λ0~1.54μm. The DSF had a 5μm core diameter, a Δn of 0.017 and a loss of

less than 0.4dBKm-1 at 1.55μm. At 1.559μm, by optimisation of the cavity length, the
laser generated 150ps pulses at 100MHz repetition rate, as displayed on a synchroscan
streak camera with an S-1 photocathode. Up to 60mW of average output power was
available corresponding to around 4W peak power, of which up to 30mW average
could be coupled into the DSF with a X10 ARCO objective. The GVD at this
wavelength was estimated to be D=1.5ps nm-1km-1.

A more useful version of Eq(5.1.6) is given by6:

117
2 2 c n2 P
Ωc = 3
λ Aeff D (5.2.1)

where all symbols have their usual meaning. Given that λ=1.559μm, P=2W peak,
Aeff=33μm2, and D=1.5ps nm-1Km-1, Ωc was predicted to be 430GHz. Hence under

these conditions, MI gain was expected for frequencies below 430GHz, peaking at:

Ωc
Ωm = ≈ 300GHz
2 (5.2.2)

Temporal
evidence of this
modulation through
autocorrelation would
therefore exhibit a period
of 3.3ps. Since the pump
pulses used here were of
~150ps duration, they
appeared CW as far as
MI was concerned.

Fig.5.2.2. Spectral output of the DSF showing MI Fig.5.2.2. shows the


sidelobes for a launchedaverage power of 30mW.
spectral output of the
DSF for an average
launched power of 30mW at 1.559μm.

Clear sidelobes due to MI were apparent, with a spectral sideband separation


of 2.25nm. The central portion of the spectrum showed two more peaks, with a peak to
peak bandwidth of ~1nm. This was probably a result of SPM (see section 1.1) from
the high peak power pulses propagating over a long length of fibre. Even with SPM
included, no more than a few picoseconds of temporal reshaping was expected, since

118
the dispersion length for the pump pulses was quite long.

The peak power of the signal emerging from the DSF was no more than ~2W ,
making it very difficult to observe any temporal modulation by autocorrelation.
Instability in the pump laser when generating pulses as short as 150ps caused severe
jitter in the spectral sideband separation, which tended to wash out the temporal
manifestation of the MI. To improve the signal to noise ratio of the system, slow scan
autocorrelation was used with a scan time of ~1ps per second of scan, with a total
capture time of ~30s. The resultant semi-interferometric trace obtained showed a
marked increase in contrast due to the enhancement of the fringes within the transform
limited portion of the input pulse. The pump pulses were significantly chirped, and
hence the fringes only occurred over the central portion of the autocorrelation trace.
Fig.5.2.3. shows the resultant modulated semi-interferometric autocorrelation trace of
the DSF output under the same conditions as those for the spectra of Fig.5.2.2.

Fig.5.2.3. Semi-interferometric autocorrelation trace of the DSF output showing MI.

A clear temporal modulation with a 3.6ps period was evident, consistent with

119
the inverse sideband separation in the spectrum of Fig.5.2.2. Note that the interference
fringes were not perfectly resolved, and had a spacing resulting from signal aliasing
with the sampling oscilloscope resolution at this scan rate. The measured 3.6ps
modulation period was within good agreement with that predicted at 3.3ps, with any
discrepancy probably due to either pump depletion, or under estimation of the mode
field area Aeff.

In order to obtain autocorrelation measurements, the laser was adjusted to


produce as much peak power as possible, with 150ps output pulses. However
operation under these conditions was extremely unstable over periods in excess of 1
minute, and to investigate the dependence of the MI period with pump power, a more
stable regime was used where the laser produced pulses in excess of 500ps duration.
The stability increased dramatically, enabling spectra to be sequentially measured.
Fig.5.2.4. shows the dependence of the MI sideband separation on pump power in a
series of spectra.

Fig.5.2.4. Spectra illustrating MI sideband frequency dependence on average pump


power.

A strong power dependence clearly existed, in accord with Eq(5.2.1). Within


experimental error, the measured sideband separations were in agreement with those

120
predicted, indicating modulation periods (frequencies) of approximately 11.5ps
(88GHz), 16.5ps (61GHz) and 27ps (37GHz) for launched powers of 15mW, 9mW
and 5mW respectively. These frequencies spanned the range 37-277GHz, and
therefore represented a source of modulations at a tunable repetition rate around
~1.56μm. If these modulations were amplified, they would have been expected to
develop into deep solitary waves and eventually form solitons. This proved that MI
could be induced from peak powers present within an Er doped fibre laser, providing a
useful ultra-high repetition rate source over this wavelength range, for use as clock,
signal or synchronisation pulses in an all optical communications sytem.

In conclusion, an actively mode locked Er fibre laser has been shown to induce
MI in pulses of a few hundred picoseconds, with characteristic modulation frequencies
in the range 37-277GHz determined by the launched pump peak power (or pulse
duration). The process could hopefully be induced intra-cavity in a fibre laser, to
produce deep modulation of the laser filed, or extra-cavity to produce modulations
which could be amplified further into solitons. Further research could produce a CW
pumped source of ultra-high repetition rate pulses if a high power Er fibre amplifier7
were used as the gain medium. Higher modulation frequencies should be obtainable
with higher pump powers or lower GVD, by tuning as close as possible to λ0, although

the latter procedure has been found to level off the modulation frequency to between 5
and 10THz as higher (4th) order dispersion dominates, instead of increasing to infinity
as Eq(5.2.1) predicts.

5.3 Cross Phase Modulation

In section 1.1 Self Phase Modulation (SPM) was introduced, where the
intensity dependent refractive index present in a fibre results in a variation in the phase
across an optical pulse, and the development of a positive frequency chirp. If two
waves copropagate in an optical fibre, they can perturb the refractive index seen by
themselves and by the other wave in exactly the same way, resulting in Cross Phase

121
Modulation (XPM). Neglecting coupling through Stimulated Raman Scattering (SRS)
or Stimulated Brillouin Scattering (SBS) and coupling processes such as phase
matched Four Wave Mixing (FWM), the coupling through XPM can result in exactly
the same effect as SPM if the two pulsed waves copropagate perfectly. The resultant
phase modulation is a result of both SPM and XPM for each of the two waves. In
accordance with the treatment of SPM given in chapter 1, and neglecting absorption in
the fibre, the coupling can be shown to be related to an intensity dependent refractive
index change:

n = n0 + Δ n (5.3.1)

where n0 is the unperturbed or linear refractive index experienced at low intensities,

and Δn is given by:

Δ n = n2 E1 2 + 2 E2 2
(5.3.2)

where the nonlinear refractive index coefficient n2=3.2X10- 20m2W-1. The first term in

this equation is similar to that given in Eq(1.4.7), and results in the familiar SPM of
the wave by itself. The second term is responsible for XPM, and is dependent on the
intensity of the copropagating field intensity. Note how the coupled term is twice as
effective8a. The effect is sometimes referred to as Induced Phase Modulation (IPM)
when the coupling occurrs between two signals derived from the same source wave,
e.g. XPM from a pump wave to a Stokes wave through SRS9,10.

XPM imposes severe limitations on the information capacity of multichannel


nonlinear transmission systems , and can result in pulse distortion and cross-talk11. It
is manifest as a phase modulation similar in aspect to SPM, and is easily detected in a
weak signal which copropagates with an intense pump field12. The effect is
substantially reduced in counter-propagating geometries because the fields (if pulsed),
only interact for a very short accumulated distance. The exact nature of the effect is

122
complicated by pump - signal walk-off13,14, resulting in spectral asymmetry and even
cancellation of the local temporal phase shift. With walk-off, signal downchirp
induced by the leading edge of the pump pulse can be rapidly followed and cancelled
by the upchirp induced by the trailing edge. Similarly, no spectral broadening may
result from the interaction of a very short signal with a very long, intense pump if no
walk-off occurs. Since the intensity of the pump is slowly varying on the signal
timescale, little phase shift with time is experienced, and no chirp is induced. With
similar pump and probe pulse durations and no walk-off, the probe experiences a
phase modulation through XPM similar, but twice in magnitude to that experienced by
the pump pulse through SPM. The spectral effect is exactly similar to SPM, apart from
discrepancies which may arise from the experimental nature of the signals. For
example a CW signal would yield a large CW spectral component, which would
remain unmodulated between the copropagating pump pulses.

XPM is most pronounced in a weak signal when it copropagates at the exact


group velocity of the pump (perfect tracking), and the signals are of comparable
duration14. Fig.5.3.1. shows the predicted phase shift (left) and spectra (right) arising
from XPM in a signal pulse copropagating with an intense pump pulse of the same
duration.

123
Fig.5.3.1. Phase shift (left ) and spectral broadening experienced by a signal pulse
for (a) perfect tracking,(b) initial coincidence followed by walk-off and (c) complete
pulse-signal walkthrough.

For the case where pump and signal pulses perfectly track with each other (a),
the resultant effect on the signal is indistinguishable from SPM. In (b) the two pulses
initially coincide in the fibre, and then the pump walks off through the signal either
forwards or backwards. The resultant phase shift is clearly influenced by the sign of
curvature of the pump pulse edge which walks off through the signal, resulting in a
strongly asymmetric spectral broadening. If the pump walks completely through a
signal pulse of finite and similar duration as shown in (c), the effect of the two
intensity edges produces no net phase change, on the timescale of the signal pulse. The
spectrum undergoes no net change, although some slight spectral shaping may occur
through incomplete compensation or interference during the interaction. In the work
that follows, the influence of an intense pump on a weaker signal through XPM is
investigated for the cases of perfect tracking and walk-off. XPM can provide a means
of spectrally broadening a low level signal incapable of inducing spectral broadening
in itself, in order to perform dispersive compression with gratings, or dispersion
compensation through soliton shaping in the fibre itself. This latter technique is

124
responsible for the results presented in chapter 5.5 where pulses are generated from a
CW laser diode signal.

5.4 Induced Modulational Instability from Pulses in the

Normal Dispersion Regime.

In section 5.1, MI was introduced, noting that the effect can only
spontaneously arise in the anomalous dispersion regime. This is true for MI which is
phase matched through SPM, and grows through soliton shaping into solitary waves.
Since the intensity of a signal is of importance only to induce nonlinearity, there exists
the possibility of inducing MI in a signal through the influence of a strong
copropagating pump field, which spectrally broadens the signal through XPM. In this
section, sub-picosecond solitary waves are experimentally derived at 1.32μm (just
within the anomalous GVD regime for a standard fibre) through the process of XPM
induced MI3. The MI is induced through copropagating an intense pump pulse at
1.06μm (which is well within the normal GVD regime), of approximately the same
duration as the signal pulse.

The conditions necessary for MI to occur are similar to those required for
soliton propagation. The signal must have sufficient peak power in order to overcome
the tendency for the modulation to collapse through dispersive broadening. The
necessary peak power can be provided by a copropagating pump, injected
simultaneously with a weak signal, even if the signal is of insufficient intensity to
induce MI by itself. By varying the power and/or frequency of the external modulation
provided by the pump, solitary waves of controllable width and repetition rate could
be produced4. The coupling between pump and signal through XPM can be
detrimental and lead to cross- talk in multichannel communication systems11, however
here it allowed a signal in the normal dispersion regime to provide the high intensities
required in the interaction. MI has been shown to occur for signals propagating in the
normal dispersion regime, with the required phase matching provided solely through

125
SPM and XPM, or by propagation in different fibre modes15,16, however the effect was
closer to phase matched degenerate FWM than soliton shaping.

In this experiment, to ensure that the pump and signal propagate with similar
propagation constants17, the fibre was chosen so that the minimum dispersion
wavelength λ0 was between the pump and signal wavelengths. The group velocities

were made more similar in this way, although they still differed considerably and
hence walk-off still occurred. The pump pulses experienced SPM , and transferred a
chirp to the signal pulses of twice the magnitude via XPM, provided the two signals
did not walk-off. In the anomalous GVD regime, the chirped signal underwent soliton
compression, and developed MI from amplitude or phase fluctuations in the pump
field. The experimental arrangement is shown in Fig.5.4.1.

Fig.5.4.1. Schematic of the experiment for demonstrating induced MI through XPM.

The signal pulses were derived from a CW mode locked Nd:YAG laser
operating at 1.32μm, generating 100ps pulses at 100MHz repetition rate with average
(peak) powers of up to 1.8W (180W). Although previously described theoretical MI
considerations have been based on CW signals, a pulsed source is nearly always used
to avoid loss through SBS, which did not occur here as the pump and signal
bandwidths were too broad. The high peak powers obtained with pulsed excitation
also facilitated autocorrelation, which was required to measure temporal modulations
on the timescale of the MI. However on this timescale, both the pump and signal

126
pulses were in fact approximately CW. The pump pulses were provided by a similar
Nd:YAG laser operating at 1.06μm, producing 100ps pulses with average (peak)
powers of up to 5W (500W). Power fluctuations in both lasers were less than 3%peak,
and the laser pulsewidths were continuously monitored by sampling both input and
output beams from the experiment on a synchroscan streak camera. This also
facilitated synchronisation of the pulses, which were temporally overlapped to within
20ps at the fibre input with an optical delay line.

The fibre used for the nonlinear interaction (F2) had a minimum dispersion
wavelength λ0=1.27μm, and hence was weakly anomalously dispersive for the probe

signal at 1.32μm, and normally dispersive for the pump at 1.06μm. The fibre had
GVD parameters |D|=5ps nm-1km-1 at 1.32μm, and |D|=26ps nm-1km-1 at 1.06μm.
Since the GVD parameters and hence group velocities still differed considerably, the
XPM interaction length was limited to the walk-off length which was computed as
follows: The propagation time delay between two signals with wavelength separation
Δλ is given by18:

Δ T = L D λ Δλ
c λ (5.4.1)

where L is the fibre length, λ is the average wavelength and D(λ) is the relative
normalised dispersion parameter in dimensionless units given by:

D λ = D cλ (5.4.2)

where D is the GVD paramter in ps nm-1Km-1. This yields:

ΔT = D Δ λ L (5.4.3)

For the 100ps pump and signal pulses, ΔT=200ps, |D|=(26-5)=21ps nm-1Km-1,
and Δλ=255nm. The interaction length L was therefore 37m for the pulses to walk-off

127
by 200ps. The length over which MI develops is characterised by the nonlinear length
LNL=(γP0)-1 from Eq(5.1.7). For fibre (F2) used here, Aeff=96μm2, and hence for a
typical launched pump power of 1W peak, LNL=212m, which was over 5 times the

walk-off length. To increase the interaction length for the peak powers used in this
experiment, the 100ps pulses were initially broadened in fibre (F1) which was 2km
long, with a minimum dispersion wavelength λ0=1.48μm. The combined action of

SPM (which caused ~10nm spectral broadening) and the large positive GVD
parameter of 20ps nm-1km-1 broadened the pump pulses to a duration of approximately
500ps. The maximum launched power into fibre (F1) was maintained at 400mW using
a X10 microscope objective (MO1) to ensure that Raman scattering did not occur18,10,

which would have severely distorted the pump pulses and resulted in a highly
nonlinear induced phase shift of unknown quality in the signal, via XPM. The
resultant 500ps, 8W peak power pump pulses emerging from fibre (F1) had a SPM
broadened spectrum, with a linear chirp19 of approximately 0.02nm s-1. ΔT in
Eq(5.4.3) was now 500+100=600ps for the pump - signal interaction, yielding a walk-
off length of L=112m. This walk-off length was much closer to the nonlinear length
(212m), increasing the chance of generating MI. The use of 1.2km of fibre for (F2)

also meant that if total phase synchronism was lost between pump and signal pulses,
the pulses would still interact somewhere in the length of fibre (F2).

Both the 1.06μm pump pulses from (F1) and the 1.32μm signal pulses were
simultaneously launched into fibre (F2) using a dichroic beamsplitter (BS2) and X20
achromatic microscope objective (MO3). The radiation emerging from the fibre was
collected and collimated with an identical objective (MO4) into a 1m scanning

spectrograph and a scanning, background-free (non-collinear, see appendix)


autocorrelator, providing simultaneous spectral and temporal resolutions of 0.1nm and
<30 fs respectively. Less than 50mW average (1W peak) power of the 500ps, 1.06μm
radiation was coupled into the nonlinear interaction fibre (F2) to ensure once again that

no Raman conversion took place. The pulses continued to broaden dispersively in this
fibre from the 500ps, 10nm bandwidth to 800ps in the presence of a GVD parameter

128
of 26ps nm-1Km-1 at 1.06μm. The average power of the 1.32μm radiation launched
into fibre (F2) was maintained at 20-25mW (2-2.5W peak), which was below the

observed threshold for modulational instability to be observed in this fibre4,20. This


was confirmed by examining both the spectral and autocorrelation outputs for the
1.32μm radiation launched alone, which showed no characteristic sideband generation
or modulation, and no Raman coversion, just slight spectral broadening due to SPM.

The spectral output from fibre (F2) around the 1.32μm signal wavelength both

with and without the pump radiation at 1.06μm is shown in Fig.5.4.2. for various
levels of pump power.

129
Fig.5.4.2 .Variation of fibre (F2) output spectrum with pump power (a)0, (b)20mW
(0.4Wpeak), (c)30mW (0.6W), and (d)40mW (0.8W) at 1.06μm. The signal power was
25mW at 1.32μm.

With no pump power present (a), for 25mW average signal power at 1.32μm
(~2.5W peak) only slight SPM induced spectral broadening was evident. This is
consistent with that predicted from Eq(1.4.10), i.e.~1.1nm, provided the uncertainty in
launched pulse duration was accounted for. When 20mW (0.4W peak) of 1.06μm
chirped pump radiation was introduced simultaneously, a rapid growth of sidebands
was evident (b), separated from the central peak by ~8.7cm-1 (260GHz, or 1.5nm). A
further increase of the pump power at 1.06μm to 30mW (c) and 40mW (d) caused the
sideband intensity to increase accordingly. The sidebands in turn generated secondary
sidebands at a frequency separation of ~17.8cm-1 from the central peak. A five fold
increase in the intensity of the secondary peaks occurred for a 30% increase in pump
power, illustrating the rapid exponential nature of the MI gain process predicted
theoretically1. The fundamental modulation increased steadily with pump power, as
predicted in Eq(5.1.6).

Owing to a XPM coefficient twice the magnitude of SPM21, spectral


broadening of the central feature was evident at high pump powers. The measured

130
bandwidth of the central feature was 1nm without pump radiation, increasing to
1.35nm for 40mW of launched average pump power (0.8W peak) at 1.06μm. This was
evident in the spectra shown in Fig.5.4.2(a) & (d). The 0.35nm bandwidth increase
was in good agreement with that predicted (0.37nm) for an interaction length of 112m
in fibre (F2).

Fig.5.4.3. Fibre (F2) output spectra for an


average signal power of 20mW at 1.32μm and
simultaneous pump powers of (a) 0 and (b)
30mW average at 1.06μm.

Fig.5.4.3. shows output spectra recorded


under similar conditions, except that the 1.32μm
Fig.5.4.4. Autocorrelation trace
corresponding to the power levels in signal average power was reduced to 20mW,
Fig.5.4.3(b). MI through XPM was
apparent temporally, lead ing to resulting in less SPM broadening to 0.66nm, as
520fs pulses separated by 3.5ps
period, superimposed on a ~100ps shown in (a). When 30mW of average power at
pedestal. Zero intensity level is
indicated. 1.06μm was introduced, XPM broadened the
central feature to 0.8nm, and MI sidebands
appeared separated by approximately 9.7cm-1 (~290GHz or 1.7nm).

Corresponding time resolved measurements made by autocorrelation indicated


no temporal structure present on the broad 100ps signal pulses at 1.32μm for 20mW

131
launched signal power alone. In the presence of the 1.06μm pump pulses, a distinct
deeply modulated, sub-picosecond pulse structure appeared on the autocorrelation
trace as shown in Fig.5.4.4. The duration of the individual structured pulses was
measured (assuming sech2 profiles) to be 520fs, with a temporal period between the
pulses of 3.5ps, which was in good agreement with the inverse of the 290GHz
measured frequency separation of the MI sidebands shown in Fig.5.4.3(b).

Decreasing either launched pump or probe power, over a fixed fibre length,
resulted in a reduction of the temporal structure and an increase in the modulation
period. This was consistent with theoretical prediction11, and was indicative of optimal
compression of the modulation occurring earlier in the fibre for higher pump powers
with, as a result, the generation of more satellite pulses over the rest of the gain length
in the fibre. The femtosecond structure shown in Fig.5.4.4. was superimposed on a
long ~100ps background pulse represented by the input signal pulse. A modulation
depth at the centre of 60% maximum was indicated on the trace, indicating that the
pulse structure was 100% modulated6a,6b. This was confirmed by the large proportion
of energy present in the modulated sidebands of the spectrum in Fig.5.4.3(b).

The effect of walk-off could have been rectified by appropriate choice of


source wavelengths to ensure perfect group velocity matching of the signals, and may
have resulted in complete modulation of the signal pulse down to the baseline.

In conclusion, MI has been demonstrated producing pulses of ~520fs at up to


~290GHz repetition rates, for the first time through XPM of a weak signal pulse in the
anomalous dispersion regime, by a strong pump pulse in the normal dispersion regime.
This experimental result is of direct relevance to the production of short pulses from
sources at wavelengths where traditional techniques of mode locking or pulse
compression are not applicable. The effects demonstrated here also impose severe
restrictions on the density and peak power of multichannel communication systems.
When information is transmitted by Wavelength Division Multiplexing (WDM)

132
channels may suffer from cross-talk induced MI. Coherent detection would be
rendered unusable if XPM altered the incoming phase of transmitted information.

For an M-channel system, the maximum power in each channel is restricted to:

P < 0.05 α
γ M-1 (5.4.4)

where α-1 is the fibre loss length and γ is the nonlinear coefficient defined in
Eq(1.4.12). This is consistent with experimental data obtained with laser diode signals
propagated over 15km of fibre22,23, which resulted in an imposed power limitation of
1mW peak to restrict the induced phase shift below ΔΦ=0.1. Since the soliton power
scales with γ-1, this suggests that WDM in a soliton transmission system may present a
problem, since low values of γ would result in a higher fundamental soliton power P0.

5.5 Picosecond Pulse Generation through Cross Phase Modulation.

In section 5.3, Cross Phase Modulation (XPM) was introduced. It was shown
theoretically that XPM between copropagating waves with non-overlapping spectra
could lead to spectral broadening of one wave by the other (and vice versa). For the
case where both waves are pulses of approximately similar durations, the effect is
indistinguishable from SPM, apart from phase and spectral asymmetry resulting from
walk-off between the pulses. If dispersion begins to dominate over the fibre lengths
involved, i.e.

T ps < L β 2

where L is the fibre length, T is the pulse duration and |β2| is the the GVD in ps2km-1,

the asymmetric spectral reshaping can also lead to temporal reshaping. In this chapter,
XPM is applied to a low intensity CW laser diode signal in the anomalous GVD
regime, induced through copropagation with an intense 100ps pulse in the normal
dispersion regime. In this way, picosecond pulses are generated from the weak CW

133
diode laser signal through the temporal reshaping which results from spectral
broadening through XPM by the intense pump4. The effect relies primarily on the
soliton-like shaping which occurs when a spectrally broadened pulse propagates in the
anomalous GVD regime, but the spectral broadening arises through the high intensity
of the pump radiation in the normal GVD regime rather than SPM arising from the
signal.

For a CW signal, under appropriate conditions, picosecond pulses can develop


through XPM from a pulsed pump source11. Due to the presence of the intense pump
pulse, the XPM coupling through the intensity dependent refractive index causes a
time dependent phase shift across the signal, which is proportional to the intensity of
the pump pulse envelope. This results in a frequency downshifting of the leading and
upshifting of the trailing edge of the signal, and the development of a positive chirp.
The edges referred to are the parts of the signal within the envelope of the pump, since
those parts outside where the pump field is zero remain unperturbed. In the anomalous
GVD regime, this chirped signal will tend to compress temporally through soliton-like
shaping, with the eventual production of a pulse. The duration of the pulse depends on
the linearity of the induced chirp, and the resultant spectral bandwidth. The process is
enhanced when both pump and signal copropagate with equal group velocities. In a
single mode fibre this can be achieved by ensuring that both pump and signal lie either
side of the minimum dispersion wavelength λ0, so that their group velocities are the

same.

Fig 5.5.1. illustrates the effect of pump - signal walk-off on the spectrum of the
signal for the situations where the group velocities of the pump and signal are equal
(Vp=Vs), and for the situations where (Vp<Vs) and (Vp>Vs). These theoretical

predictions are based on XPM between a strong pump pulse and a weak signal pulse
of the same duration, which initially coincide in the fibre.

134
Fig.5.5.1. XPM spectra for signal walk-off through the front of the pump (Vp<Vs),
perfect tracking (Vp=Vs), and through the rear of the pump pulse (Vp>Vs).

In the first case (Vp<Vs) the signal pulse walks through the front of the pump

pulse, and hence only receives a net red shift resulting from the leading edge of the
pump pulse. The opposite occurs for the case where (Vp>Vs). For the perfect case
tracking where (Vp=Vs), the spectrum remains symmetric and indistinguishable from

SPM. For complete walkthrough of the pump and signal, which can only occur if the
pulses initially do not overlap, the net phase change is zero, as the signal first
experiences a net red shift followed swiftly by a net blue shift.

For the situation envisaged here, however, the theoretical predictions in


Fig.5.5.1. are opposite to that expected, due to the CW nature of the signal. The
scenario can be visualised by noting that with walk-off, the CW signal experiences
cancellation of the initially induced chirp, with the only surviving phase shift
occurring at one end of the pump pulse.

In the region of anomalous GVD, spectral broadening can lead to significant


temporal reshaping25. Asymmetric spectral broadening can lead to asymmetric
temporal pulse shaping, with the development of rapid oscillations at one end of a
pulse whilst the other end is largely unaffected. These oscillations are due to optical
wave breaking, which was dicussed in chapter 1. The effects are only noticeable if the
XPM walk-off length becomes comparable with the fibre dispersion length LD.

Symmetric spectral broadening with a linear, positive chirp should lead to soliton

135
pulse shaping of the CW signal area which overlaps with the pump pulse.

The experimental arrangement was relatively simple, involving simultaneously


coupling the ouput of a CW semiconductor diode laser and a CW pumped mode
locked Nd:YAG laser operating at 1.06μm into a length of single mode optical fibre.
The beams were combined and coupled into the fibre using a dichroic beamsplitter and
a broadband X20 ARCO. The output from the fibre was interrogated both spectrally
with a 1m scanning spectrograph and temporally with a photodiode. The CW laser
diode was a grating tuned, InGaAsP device which permitted tuning over the 1.43-
1.56μm spectral range, with a maximum average output power of 500μW. A
conventional, CW pumped mode locked Nd:YAG laser provided the pump pulses,
which were of 100ps duration at a repetition rate of 100MHz, with peak powers of up
to 700W.

The fibre length was 1.2km, with a manufacturers calculated minimum


dispersion wavelength λ0=1.273μm. The mode field area was approximately 96μm2,

and the fibre was single mode above 1μm. Fig.5.5.2. is a plot of the measured group
delay versus wavelength with the computed GVD parameter for the fibre used in this
experiment. From this plot, it was estimated that the pump pulses at 1.06μm would
perfectly track at the same group velocity with a signal propagating at 1.53μm. This
permitted efficient overlap of the pump and signal over the entire fibre length,
producing symmetric spectral broadening. The GVD parameter at 1.06μm was 26 ps
nm-1km-1 (normal GVD) and at 1.53μm was 16 ps nm-1km-1 (anomalous GVD). For a
maximum launched pump power of 500mW average (50Wpeak), which represented a
power density of ~50MWcm- 2, the magnitude of the expected spectral broadening in
the 1.53μm signal was over 30nm, assuming perfect pump signal tracking and
neglecting Raman conversion in the pump. The maximum launched signal power was
60μW average, hence effects from SPM in the signal were negligible over the fibre
length involved.

136
Fig.5.5.2. Measured group delay (upper) and calculated GVD (lower), indicating
perfect tracking between the signal at 1.53μm and the pump pulses at 1.06μm.

Nonlinear spectral broadening of the CW signal at 1.53μm could only occur


over the duration of the pump pulses. Since the duty cycle of the 100ps pump pulses at
a 100MHz repetition rate was 1%, more than 99% of the spectrum at 1.53μm was
dominated by the unmodulated CW diode signal. To reduce the relative intensity in
the CW portion of the spectrum and improve the signal to noise ratio, a mechanical
chopper was placed in the pump beam, and a lock-in amplifier was used to measure
the transmitted modulated diode spectrum at the frequency of the chopped pump
beam. (This was especially necessary where walk-off occurred, since only a small
spectrally broadened component was measured). The input laser diode spectrum was
single line, with a spectral width limited by the 0.1nm resolution of the spectrograph.
Fig.5.5.3. is the measured output spectrum of the fibre at the diode wavelength,
indicating a spectral broadenening due to XPM, for 480mW of average pump power
and 60μW of laser diode signal power.

137
Fig.5.5.3. Spectrally broadened laser diode signal through XPM from 480mWof
average power in 100ps pulses at 1.06μm.Diode laser tuned to (a)1459.5nm and
(b)1541.5nm.

In (a) the laser diode was tuned to 1459.5nm, resulting in enhanced broadening
occurring on the short wavelength side of the spectrum. Long wavelength
enhancement occurred for tuning to 1541.5nm, as shown in (b).The asymmetric
broadening was related to the influence of relative walk-off between the pump and
signal in the fibre, since in (a) (Vp<Vs), and in (b) (Vp>Vs).

A direct similarity was evident in these spectra to those predicted in Fig.5.5.1,


if the CW nature of the signal was taken into account. Much theoretical consideration
has been given to the effect of walk-off on XPM14,11,13. The results shown in Fig.5.5.3.
have a marked similarity to numerical simulations based on the effects of XPM and
walk-off in the presence of significant dispersion13. Fig.5.5.4. shows the theoretically
predicted spectrum for a similar case where dispersion occurs over a length scale ten
times that of the walk-off length for the XPM interaction, for a pulsed signal of the
same duration as the pump. The spectra are shown both with and without the inclusion
of GVD, for the case where the pump pulse travels faster than the signal (Vp>Vs).

138
Fig.5.5.4. Spectral broadening occurring in signal pulses through XPM, with and
without the presence of significant dispersion. The asymmetry results from walk-off
where (Vp>Vs).

The spectra shown in Fig.5.5.3 showed a marked similarity to these


predictions, indicating that dispersion was playing a significant role in producing
asymmetry in the broadened signal. The results differed only in the sign of asymmetry,
due to the CW nature of the signal, and in the presence of a significant CW component
resulting from areas of the signal which were modulated, but experienced no net phase
shift as they passed through the pump pulse. This CW component would not have
been present for ideal lock-in detection if no walk-off had occurred. The spectral
broadening varied as a function of pump power according to the graph show in
Fig.5.5.5. for a launched signal power of 60μW and a signal wavelength of 1530.5nm,
which was regarded as close to the optimum wavelength for perfect tracking.

Fig.5.5.5. Variation of spectral broadening width with pump power through XPM, for
a fixed signal power of 60μW at 1530.5nm.

139
The bars in Fig.5.5.5. relate to the bandwidth or spectral extent of the diode
signal broadened by the corresponding average pump powers shown. A trend in
increased spectral width was apparent, with little shift in the centre wavelength,
indicating that near perfect tracking was achieved for a signal near 1.53μm as
predicted by Fig.5.5.2.

For the highest launched average pump powers (500mW) significant Raman
conversion occurred from 1.06μm to 1.12μm, interfering with the XPM process by
depleting the pump pulse envelope. This was particularly noticeable for lower signal
wavelengths (e.g.1452nm), at which pump powers sufficient to generate Raman
conversion produced significant spectral broadening up to three times that normally
measured when Raman conversion was absent. The resultant broadening was also
asymmetric, being enhanced to the long wavelength side. This can be explained
qualitatively by the pump travelling slower than the signal, as shown in Fig.5.5.2. for a
signal wavelengh of 1452nm. During Raman conversion the pump pulse would have
become depleted and the Raman component would walk-off through the front of the
pump pulse, inducing a distinctly asymmetric chirp onto the pump through XPM13,10,
resulting in an enhancement on the long wavelength side. This chirp, if transferred
through XPM onto the laser diode signal, would have been responsible for the
observed spectral asymmetry at high pump powers.

Fig.5.5.6. Variation of spectral broadening through XPM with signal wavelength, for
60μW of signal and a fixed 480mW of average pump power.

140
Fig.5.5.6. illustrates the effect of group
velocity mismatch and hence walk-off between the
pump and signal on the XPM process by variation
of the signal wavelength. The average signal power
was maintained at 60μW, and the pump power at
480mW.

A discernable increase in the spectral


broadening took place for a signal wavelength of
1512nm, indicating that this was close to the
wavelength for which perfect tracking with the
pump pulse occurred. By tuning the laser diode near Fig.5.5.8. Top to bottom:
PUMP- pulses monitored at
1512nm, a more symmetric spectrum with enhanced 1.06µm. CW-Carrier wave
modulated at the centre
broadening was obtained without the use of a lock- wavelength region of the signal
spectrum. LSB-Lower spect ral
in amplifier, as shown in Fig.5.5.7. A broad spectral sideband pulses. USB-Uppe r
spectral sideband pulses.
pedestal was generated in accordance to that Resolution 2ns limited by
amplifier bandwidth of
predicted by theory11, however the resolution of ~500MHz.

the spectrograph was compromised by the


requirement for sensitivity, and 0.1nm
resolution was insufficient to resolve the
predicted interference structure.

Assuming a 1.2km interaction length, the


~4nm spectral broadening was of the same order
of magnitude as that expected, although Raman
depletion of the pump was probably responsible
for reducing the effective peak power of the
pump pulses.
Fig.5.5.7. XPM induced sp ectrum
obtained for a signal wavelength of
In the time domain, it was not possible to 1.512µm, with 480mW of av erage
pump power.

141
obtain autocorrelation of the weak output signal to measure the generated pulsewidths.
The average signal power was around 60μW, and of this only a small fraction (~1%
by duty cycle) was attributed to the XPM signal. The pulses generated through
anomalous dispersive compression of the spectrally broadened (positively chirped)
signal could therefore only constitute a maximum of ~0.6μW of average power. It was
however, possible to display the pulses on a conventional storage oscilloscope. An AC
coupled Germanium photodiode with ~100ps rise time was used to measure the
temporal distribution of the spectrally resolved signal. Two, 15dB gain, broadband
(500MHz) RF amplifiers were placed after the photodiode to improve the signal to
noise ratio. The spectral resolution was reduced ~1nm to obtain higher sensitivity. The
rejection ratio of the spectrograph prevented pickup of the pump signal when tuned to
the signal wavelength, and no modulation of the diode signal was observed in the
absence of the pump pulses in the fibre. Modulation required the presence of both
signals, and hence resulted from the XPM process. Fig.5.5.8. shows the spectrally
filtered signals monitored with the photodiode and oscilloscope with a combined
resolution of 2ns. The top trace shows the pump pulses at 1.06μm, for phase reference.
The trace marked CW (carrier wave) was obtained by spectrally resolving the central
region of the spectrally broadened signal in Fig.5.5.7., and indicates that depletion of
the carrier signal was occurring simultaneously with modulation by the carrier pulse.
This is consistent with XPM forcing energy away from the central CW region of the
spectrum periodically. The lower traces marked LSB and USB correspond to the lower
and upper sideband components of the spectrum respectively. Clearly the sidebands
are created in synchronism with the pump pulses, within the experimental error
invloved.

Assuming that the operational parameters involved here are reasonably close to
those used in the theoretical model which this experiment was designed to emulate11,
then pulses of only a few picosecond in duration may have been formed through the
XPM process. The fibre used in this experiment was substantially longer than that
which is optimum for short pulse generation, which has been shown theoretically11 to

142
lead to short pulse fragmentation at the expense of the central pulse. Both higher
sensitivity and temporal resolution would be required to confirm this predicted
behaviour.

In conclusion, this is the first observation of picosecond pulse formation from a


weak CW laser diode signal through XPM induced by a Nd:YAG laser. The effect has
been optimised and investigated with respect to pump - signal walk-off by tuning the
laser diode wavelength through that value which resulted in perfect tracking with the
pump pulses. Stimulated Raman amplification (see chapter 6) or the use of Erbium
doped fibre amplifiers should enable the energy of the pulses to be increased and the
pulsewidths to be measured using a streak camera or autocorrelator. This technique
may be directly applicable to the production of ultrashort pulses from mode locked or
gain switched laser diode systems17.

5.6 2THz Pulse Train Generation through Induced

Modulational Instability.

Induced Modulational Instability (MI) can occur when XPM induces spectral
broadening in a relatively weak signal propagating in the anomalous GVD regime, as
described in section 5.4. The effect can also be induced if a probe wave is launched
simultaneously with a pump wave, frequency shifted from the pump so as to fall
within the MI gain region. The process can be envisaged as occurring through
frequency mixing27 between the probe wave and the relatively intense pump, with the
probe acting as a source of seed photons instead of the process evolving from
spontaneous noise. If |Ω| is less than Ωc given by Eq(5.1.6.), then MI will amplify a
probe wave at frequency ω1, where ω1=ω0+Ω, and ω0 is the frequency of the intense
pump field. Physically, two photons from the pump at ω0 are parametrically converted
into one at ω1 and one at 2ω0-ω1, analogous to degenerate four wave mixing3,4.

The process can only be induced if the pump falls in the anomalous dispersion

143
regime, unlike XPM induced MI which has been shown to produce similar phenomena
under conditions where group velocity mismatch is facilitated by use of
birefringence5,28,16. Through application of an external probe signal copropagating
with a pump of insufficient intensity to produce MI by itself, MI can give rise to the
generation of ultrashort soliton-like pulses at repetition rates governed solely by the
MI sideband detuning from the carrier4. This should allow the generation of pulse
trains with repetition rates up to ~10THz, before frequency level-off through the
influence of higher order dispersion occurs5. These modulation rates are up to an order
of magnitude higher than those obtainable electronically or electro-optically.

As decribed in section 6.2., MI can drastically reduce the required threshold


powers for soliton Raman generation in optical fibres, with the MI sideband falling
within a region of significant Stokes gain and evolving into fundamental soliton
structures which undergo soliton self-frequency shift20,29,30. In the experiment
described here, a train of 130fs pulses is generated at picosecond repetition periods
through probe induced MI, using a CW laser diode injected on the anti-Stokes side of
a Nd:YAG laser carrier pulse. The experimental arrangement is shown in Fig.5.6.1.

Fig.5.6.1. Schematic diagram of the experimental arrangement.

The pump or carrier frequency was provide by a CW mode locked Nd:YAG


laser operating at 1.32μm. On the timescale of the expected THz MI the 100ps pump
pulses produced from this laser represented a quasi-CW field. A maximum average
output power of 2W at 100MHz repetition rate provived up to 200W of peak power,
but launched pump powers were maintained below the level which generated

144
spontaneous MI.

The CW probe signal was provided by a CW InGaAsP Fabry-Perot type


semiconductor diode laser. The chip was 200μm long, with plane uncoated facets
providing a multi-longitudinal mode output centred around 1.307μm, with an average
output power of ~1mW. Both pump and signal lasers, with orthogonally polarised
output beams, were combined in a conventional air-spaced polarising beamsplitter (P),
and focussed into and out of fibre (F) using standard X10 ARCO lenses M01 and M02.

The fibre sample (F) was 500m in length, single mode around 1.3μm, with a minimum
dispersion wavelength λo=1.3μm. The fibre was therefore anomalously dispersive at

the pump wavelength around 1.32μm, enabling MI to be observed with relatively low
power single beam pump pulses. The signal emerging from the fibre was interrogated
using a 1m spectrograph with 0.1nm resolution, and a standard background free (non-
collinear - see appendix) autocorrelator with less than 50fs resolution.

By maintaining a low average pump power in the fibre, self induced or


spontaneous MI was kept to a minimum over the 500m length of fibre. This enabled
the effect of the probe seed to be exploited as far as possible. Fig.5.6.2(a) shows the
recorded spectrum of the signal emerging from the fibre for an average launched pump
power of 70mW at 1.32μm alone.

145
Fig.5.6.2. Output spectra from 500m of fibre for input of (a) 70mW average power at
1.32μm alone and (b) with simulataneous injection of 300μw CW radiation at
1.307μm from laser diode.

At this pump power level, MI sidebands were just detectable spectrally, as


shown by the sideband growth at a ~52cm-1 (~10nm) frequency separation from the
central carrier frequency. The bands were visible spectrally at 1.308 and 1.330μm,
however no stable modulation was detectable temporally on the broad 100ps pulse
profile of the carrier by autocorrelation. The laser diode spectrum consisted of several
longitudinal cavity modes, depending on driving current, separated by 1.2nm
(210GHz) centred around 1.31μm, as shown inset in Fig.5.6.2(a) on the same vertical
scale as the other spectra. The diode spectrum therefore lay on the anti-Stokes (lower
wavelength) side of the Nd:YAG pump laser wavelength.

Fig.5.6.2(b) shows the spectrum obtained when 300μW average power of CW


laser diode radiation was simultaneously coupled into the fibre, with 70mW average
power (~7W peak) of pump radiation from the Nd:YAG laser. Parametric four wave
mixing could clearly be seen, with excitation of a Stokes sideband. Rapid exponential
gain of the sidebands was evident3, if the relative intensity of the sidebands with the
pump, and the modest probe signal was considered.

146
Stimulated Raman amplification of the Stokes sideband did not dominate for
the relatively low pump powers used here by seeding on the anti-Stokes side of the
pump wavelength. Interference of the frequency components generated spectrally gave
rise to the evolution of pulse trains temporally. An autocorrelation trace of the output
corresponding to the spectra of Fig.5.6.2(b) is shown in Fig.5.6.3.

Fig.5.6.3. Background free autocorrelation trace of the induced MI indicating the


evolution of a train of 2THz pulses.

In Fig.5.6.3(a), the trace was taken over the approximate FWHM duration of
the pump pulse, which was ~100ps. On the broad pump pulse profile, trains of
femtosecond pulses separated by ~4.6ps had evolved. The period corresponded well
within experimental error to the 4.7ps modulation period inferred from the 1.2nm
spectral separation of the longitudinal modes of the laser diode. Around the peak of
the autocorrelation trace a 50% modulation depth was apparent. This suggests that the
generated pulse trains were almost 100% modulated6a,6b. Fig.5.6.3(b) shows the
autocorrelation trace detail of one of the pulse trains on an expanded time and
intensity scale. The 490fs measured pulse separation agreed remarkably well with the
475fs separation inferred from the 12nm wavelength difference between the 1.32μm
pump and 1.307μm probe wavelengths.

Assuming sech2 pulse profiles4, the individual pulses in the train were
measured to have 130fs duration, as shown in Fig.5.6.3(b). This duration would be
dependent on pump power, wavelength, dispersion magnitude and fibre interaction

147
length. The asymmetric nature of the sidebands in Fig.6.5.2(b). was due to slight
Raman conversion providing preferential gain on the long wavelength side of the
pump. This is a typical characteristic precursor to soliton Raman continuum
generation (see section 6.2), however no Raman shifted components were observed
over the entire spectral range.

The autocorrelation trace shown in Fig.5.6.3. infers the generation of a train of


130fs pulses, whith durations shorter than the 490fs modulation period. This indicates
that some pulse formation through soliton shaping is taking place4. This behaviour is
ascribed to high order soliton propagation31 over lengths longer than the soliton
period, which is of the order of 100m for these pulses in this fibre. MI has, in fact,
been shown to be qualitatively very similar to very high order soliton breakup32. A
490fs modulation period infers a repetition rate of 2THz. It may be possible to produce
CW trains at this rate if an Erbium or Raman oscillator was phase locked using this
signal as a control train. If the diode laser source were tunable, producing a larger
frequency separation between pump and probe signals, higher pulse repetition rates
may have been possible.

In conclusion, parametric degenerate four wave mixing due to probe-induced


MI in a Nd:YAG laser pulse carrier signal has been demonstrated using a weak CW
laser diode signal. The probe operated on the anti-Stokes side of the pump in order to
minimise the strong Raman conversion effect. Trains of femtosecond pulses at THz
repetition rates, with very high modulation depths have been generated. This
technique provides a simple method for generating controllable, ultra-high repetition
rate pulse trains at rates more than an order of magnitude higher than those obtainable
through conventional electronic means.

5.7 References

(1) A.Hasegawa & W.F.Brinkman. IEEE J.Quant.Elec. QE-16, 694(1980)


U U

148
(2) K.Tai, A.Hasegawa & A.Tomita. Phys.Rev.Lett. 56,135(1986) U U

(3) A.Hasegawa. Opt.Lett. 9,288(1984) U U

(4) K.Tai, A.Tomita, J.L.Jewell & A.Hasegawa. Apll.Phys.Lett. 49,236(1986) U U

(5) H.Itoh, G.Davis & S.Sudo. Opt.Lett. 14,1368(1989) U U

(6) S.Sudo, H.Itoh, K.Okamoto & K.Kubodera. Appl.Phys.Lett. 54,993(1989) U U

(6a) M.J.Potasek. Opt.Lett. 12,717(1987) U U

(6b) M.J.Potasek & G.P.Agrawal. Phys.Rev. A36,3862(1987) U U

(7) J.F.Massicott, R.Wyatt, B.J.Ainslie & S.P.Craig-Ryan.


Elec.Lett. 26,1039(1990)
U U

(8) F.Ito, Kitayama & H.Yoshinaga. Appl.Phys.Lett. 54,2503(1989) U U

(9) J.Manassah. Appl.Opt. 26,3747(1987) U U

(10) A.S.L.Gomes, V.L.da Silva & J.R.Taylor. J.Opt.Soc.Am. B5,373(1988) U U

(11) B.Jaskorsynska & D.Schadt. Elec.Lett. 23,1090(1987) U U

(12) R.R.Alfano, Q.X.Li, T.Jimbo, J.T.Manassah & P.P.Ho.


Opt.Lett. 11,626(1986)
(13) D.Schadt & B.Jaskorsynska. J.Opt.Soc.Am. B4,856(1987) U U

(14) M.N.Islam, L.F.Mollenauer, R.H.Stolen, J.R.Simpson & H.T.Shang.


Opt.Lett. 12,625(1987)
U U

(15) G.P.Agrawal. Phys.Rev.Lett. 59,880(1987) U U

(16) P.D.Drummond, T.A.B.Kennedy, J.M.Dudley, R.Leonhardt & J.D.Harvey


Opt.Comm. 78,137(1990)
U U

(17) B.Jaskorsynska & D.Schadt. IEEE J.Quant.Elec. 24, 2117(1988) U U

(18) R.H.Stolen & A.M.Johnson. IEEE J.Quant.Elec. QE-22, 2154(1986) U U

(19) A.S.L.Gomes, A.S.Gouveia-Neto & J.R.Taylor. Elec.Lett. 22,41(1986). U U

(20) A.S.Gouveia-Neto, A.S.L.Gomes & J.R.Taylor.


IEEE J.Quant. QE-24,332(1988) U U

(21) R.H.Stolen & J.E.Bjorkholm. IEEE J.Quant.Elec. QE-18, 1062(1982) U U

(22) A.R.Chraplyvy, D.Marcuse & P.S.Henry. J.Lightw.Tech. LT-2,6(1984) U U

(23) A.R.Chraplyvy & J.D.Stone. Elec.Lett. 20,996(1984) U U

(25) G.P.Agrawal, P.I.Baldeck & R.R.Alfano. Opt.Lett. 14,137(1989) U U

(27) A.Hasegawa. Opt.Lett. 14,288(1984) U U

(28) J.E.Rothenburg. MC1. Topical Meeting on Integrated Photonics Research


U U

Hilton Head, SC. USA. March 1990. Phys.Rev.A. in press.


(29) A.S.Gouveia-Neto, M.E.Faldon & J.R.Taylor. Opt.Lett.13,1029(1988) U U

(30) A.S.Gouveia-Neto, M.E.Faldon & J.R.Taylor. Opt.Comm. 69,325(1989) U U

(31) R.H.Stolen, L.F.Mollenauer & W.J.Tomlinson. Opt.Lett. 8,186(1983) U U

(32) M.Nakazawa, K.Suzuki, H.Kubota & H.Haus. Phys.Rev.A39. 5768(1989) U U

149
6. SOLITON RAMAN GENERATION

AND AMPLIFICATION

6.1 Soliton Raman Generation

If relatively long (~100ps) pulses are propagated in standard single mode


fibres, efficient Stimulated Raman Scattering (SRS) can yield tunable sources of
radiation, with the possibility of producing Raman solitons <100fs if the Stokes band
falls in the anomalous GVD regime, as discussed briefly in chapters 1&2. The process
initially grows from amplified quantum noise, via either Modulational Instability (MI),
cross and self phase modulation induced pulse shaping or a combination of both. SRS
represents a major loss mechanism for high power pulses3,4, especially if the duration
is long, and the wavelength is close to that of minimum GVD λ0, where walk-off

between pump and Stokes pulses is minimised. The relatively broad bandwidth of the
Raman Stokes band (~100nm) extends from near 0 to over 1000cm-1 (peaking at
~440cm-1 in silica fibres), enabling amplification of pulses as short as ~100fs duration.
The pulse shaping mechanism responsible for generating a soliton Raman continuum
is however far more subtle than that involved when solitons are excited directly, by
launching transform limited pulses of appropriate peak power.

A consequence of there being finite Raman gain at near zero detuning from the
pump wavelength is the possibility of amplifying MI pulses, (see chapter 5) which
may develop on a long pump envelope if it falls in the anomalous GVD region. The
existence of MI, which is by no means a prerequisite to soliton Raman generation, can
substantially reduce the threshold powers required. Instead of amplifying MI, a small
signal injected within the Stokes Raman gain band could pre-seed the conversion
process6, experiencing considerable gain. Noise bursts or sub-fundamental soliton
power (linearly dispersive) pulses could also be amplified in a similar manner,
eventually achieving power levels sufficient to form fundamental solitons. In this
chapter a series of experiments are described demonstrating how soliton pulses form

150
the stable product of any amplified signal provided enough energy is present, and how
they can be regenerated at a point after which dispersion has reduced the intensity well
below the fundamental soliton power.

It has been demonstrated1,2 that in the region of anomalous GVD, provided


sufficient power is converted into the Raman Stokes band, soliton shaping
mechanisms can result in the generation of a spectral continuum of ultrashort
femtosecond pulses, giving rise to a spectral continuum of pulses7. This continuum
provides a tunable source of radiation, but with considerable uncertainty in timing.
Pulses as short as ~190fs can be generated without pedestals8, limited ultimately by
the response time of the SRS process9. In the following section, the continuum
generation process is dicussed.

6.2 Modulational Instability as a Precursor.

As a result of a combination of the optical Kerr effect and anomalous GVD, it


was shown in chapter 5 that a CW wave will experience exponential gain for
amplitude or phase fluctuations on the envelope, resulting in Modulational Instability
(MI). This process can be observed at very low peak powers, below the threshold for
any significant Raman conversion. Since there is finite Raman gain at very small
frequency detunings from the pump wavelength (see Fig.1.5.1), pulses formed through
MI can act as a precursor seed for the SRS process, substantially reducing the
threshold for conversion. In fact Raman soliton generation also readily develops if the
pump and Stokes bands straddle the minimum dispersion wavelength λ0 in the fibre,

since the pump and Stokes will not significantly walk-off, resulting in considerable
conversion efficiency, but this is surprisingly less efficient than if MI acts as a
precursor. The pump does not undergo MI in this case, since it falls in the normal
GVD region, however the influence of XPM on the Stokes pulse leads to considerable
soliton shaping.

151
SRS wavelength conversion of the pump radiation begins either at the Stokes
MI sideband, the peak of the Raman gain, or that wavelength which walks off least
from the pump, depending on the exact characteristics of the fibre and pump signals.
Any number of pulse-like structures developing from amplified MI undergo soliton
compression as their energy increases, with an associated decrease in pulse duration in
the presence of gain10.

By virtue of the random nature of the process which essentially evolves from
noise, more than one soliton can develop per pump pulse, at no fixed wavelength. For
relatively high powers, pump pulse fragmentation occurs rapidly through Raman
enhancement of the MI signals12. These modulations are themselves solitary waves
and can be relaunched to produce solitons in their own right using Raman gain for
amplification13. The bandwidth of the amplified and compressed soliton fragments
becomes sufficient to produce intra pulse Raman scattering, and each soliton self-
frequency shifts (SSFS) to a wavlength, dependent on the soliton duration.

SSFS is a peak power and hence pulsewidth dependent effect, such that:

dω = - D
dz τ4

where D is the dispersion parameter of the fibre, and τ is the soliton pulse duration.
The power dependency arises from the decrease in soliton duration occuring through
Raman amplification. The SSFS is extremely sensitive to soliton duration τ, and
different pulses therefore shift to longer wavelengths by varying amounts. The spectral
development with pump power, of the output of a 500m length of fibre (λ0=1.3μm),

for a launched pump signal of 100ps pulses from a Nd:YAG laser at 1.32μm is shown
in Fig.6.2.1.11

152
Fig.6.2.1. Development of spectral Raman continuum generation with average pump
power11, showing evolution from MI. Intensity sensitivity is constant throughout.

Here the process was initiated by MI, as the pump propagated in the
anomalous GVD regime. For 60mW of launched average power, MI sidebands were
apparent at a frequency separation of 67cm-1 from the pump wavelength,
corresponding to a modulation frequency of ~2THz, or a period of 500fs. Background
free autocorrelations (see appendix) used to confirm this effect showed a broad pump
envelope with this modulation superimposed. Fig.6.2.(a)-(h) shows the temporal
development of the MI as it evolved into solitons, corresponding to the previous series
of spectra.

Fig.6.2.2. Variation of recorded background free autocorrelation trace with average


pump power11 for (a) 60mW,(b) 80mW,(c) 100mW,(d) 120mW,(e) 140mW,(f)
160mW,(g) 240mW & (h) 300mW. Different intensity scales were used between traces.

The modulation frequency, Ω, for which maximum gain occurs is related to the
pump power P0 through the following relationship (from Eq(5.1.10)):

153
Ωα 2 n2 P0
D
where n2 is the nonlinear refractive index coefficient and D is the dispersion

parameter. As the pump power increased , the expected modulation frequency


increased. The effect is shown in Fig 6.2.1., although could only be seen on the Stokes
(long wavelength) sideband. This effect was seen temporally in Fig.6.2.1(a),(b) & (c).
On the anti-Stokes side there was in fact a small decrease, solely due to pump
depletion by Raman conversion. As the pump power increased further, the Stokes shift
increased continuously until the radiation took on a spectrally continuous nature up to,
and beyond 1.5μm from the 1.32μm pump wavelength. Temporally, as shown in
Fig.6.2.2, the increase in pump power reduced the number of modulation peaks
contributing to the autocorrelation, until a single pulse dominated the trace. The
contrast ratio between pulse and pedestal components also dramatically increased with
pump power.

Since the autocorrelator had a finite spectral acceptance window of ~200nm, a


pedestal was detected associated with the remaining pump fragments, which could be
filtered. The shortest obtained pulses are shown in Fig.6.2.3. for a pump power of
240mW.

Fig.6.2.3. Background free autocorrelation of Raman solitons generated for a pump


power of 240mW11, spectrally filtered to pass wavelengths above 1.35μm.

Despite filtering to prevent pump fragment detection, a significant 13%


intensity pedestal was still present, with only 12% of the total energy present as the
pulse structure. The remaining pedestal arose from dispersive radiation. Since more
than one soliton developed through MI, the essentially random pulse spacing also

154
contributed to the autocorrelation pedestal through cross-correlation.

The centre of the pump pulse had the highest peak power, and hence
discontinuities or noise were most likely to give rise to MI in this region. The pulses
generated would receive Raman gain, at a longer wavelength than the pump, and
therefore travel slower than the pump as they were further from λ0. This resulted in the

signal pulses walking through to the back of the pump pulse. These then collided with
solitons in the early stages of formation, generated near the back of the pump pulse,
with any instantaneous enhancement of peak power and bandwidth during collision
resulting in further SSFS.

Whether soliton Raman generation is preceded by MI or develops from noise


on the pump pulse, the result is a randomly spaced ensemble of solitons, of around
100-200fs duration (depending on pump power and fibre length), contained in a
spectral continuum. It is however not possible to maintain production of a single n=1
soliton, with instabilities in the pump pulse envelope causing large changes in the
wavelength, duration and hence arrival time of the generated pulse. This is because of
the evolution of the process from noise. Solitons at higher wavelengths are generally
of longer duration, since the dispersion and hence fundamental soliton power is
higher, but greater spectral separation is generated in contrast to those close to the
pump wavelength, with lower pedestals. There is however a large timing jitter
associated with each generated soliton in the continuum, from one pump pulse to
another14, of the order of 5-10ps for 100ps pulses from a Nd:YAG source.

6.3 Semiconductor Laser Diode Signals as a Precursor.

If a small external signal is injected within the Raman gain band of a pump
pulse in an optical fibre, this precursor signal would be expected to seed the Raman
conversion process in a similar manner to that produced when MI occurs. Problems
associated with the solitons being derived from spontaneous noise, or from the soliton

155
shaping that ensues from amplification of MI can in this way be overcome. To
facilitate this, a semiconductor laser diode, tunable over the Raman gain band, could
provide the seed signal which would substantially reduce the Raman conversion
threshold at one specific wavelength region. The exact wavelength at which the
conversion is initiated would be defined, and provided the pump power is controlled,
multiple pulse generation could be suppressed. The schematic for such an
experiment15 is shown in Fig.6.3.1.

Fig.6.3.1. Experimental arrangement for laser diode seeding of Raman soliton


conversion.

A mode locked Nd:YAG laser operating at 1.32μm with an average (peak)


power of up to 2W (200W) provided 100ps pump pulses at a 100MHz repetition rate.
The CW seed signal was provided by a 500μm long InGaAsP semiconductor laser
diode in an external cavity arrangement. One facet of the laser diode (the output facet
was uncoated) was coated with an antireflection coating to suppress laser action
without the addition of an external cavity reflector at the driving currents used here.
The external cavity was formed using a X20 ARCO lens, which collected and
collimated the laser diode fluorescence into a beam incident on a diffraction grating,
which was blazed for efficient diffraction in the Littro configuration over the tuning
range of the diode. By aligning the feedback from the grating back into the diode chip,
up to 1mW of average output power could be obtained in a single longitudinal mode,
over a tuning range from 1.33 to 1.38μm. A standard air spaced polarising
beamsplitter (P) enabled combination of the P-polarised pump pulses and the S-
polarised CW diode signal. This also offerred a degree of protection for the diode chip

156
from the high power pump pulses, which caused damage in the diode facet if
relaxation oscillations occurred in the pump laser.

A standard X20 microscope objective (MO1) focussed both pump and signal

beams into 1.98km of standard single mode fibre with a dispersion minimum at
λ0=1.34μm, and a loss of 0.5dBkm-1 around 1.3μm. An identical objective (MO2)

collected and collimated the fibre output beam into both a 1/2m scanning
monochromator and a background free (non-collinear - see appendix) autocorrelator
with resolutions of 0.2nm and ~20 fs respectively. To investigate the temporal
constituents of the broad spectral features generated, narrow bandpass dielectric filters
were inserted before the autocorrelator. Fig.6.3.2(a) shows the spectrum recorded for
an average launched pump power of 160mW(~16Wpeak) at 1.32μm in the fibre.

Fig.6.3.2. Spectra of fibre output for (a) 160mW average pump power at 1.32μm
alone, (b) as in (a) plus 60μW CW diode laser signal at 1.36μm, (c) 180mW pump
power alone, and (d) as in (c) plus 70μW CW diode laser signal at 1.36μm.

No MI was evident at any level of pump power, since the pump wavelength
here was in the normal GVD regime. This resulted, for equivalent fibre and input
pulse parameters, in a higher threshold for Raman soliton generation17. This higher
threshold was beneficial in the experiment, reducing the chances of producing
spontaneous Raman solitons from fluctuations of the pump pulse profile, at power
levels which would produce little gain in the laser diode signal. A low intensity
Raman component centred at 1.38μm was evident in (a) (the Raman gain peak occurs
around 1.4μm in fused silica fibres16), but at this power level (160mW) no temporal

157
soliton evolution was evident in the most sensitive autocorrelation measurements,
indicating that any radiation in the Raman component was of insufficient energy to
form solitons.

The laser diode signal was tuned ~20nm lower than the spontaneous Raman
peak to 1.36μm in order to define the wavelength of soliton Raman operation. When
60μW of laser diode signal was simultaneously coupled into the fibre with 160mW of
pump radiation, a large amplification of the diode radiation at 1.36μm took place, as
shown in Fig.6.3.2(b). It was possible to continuously tune the laser diode from 1.33 to
1.38μm, and obtain gain centred at the seed wavelength in a similar manner.

Fig.6.3.2(c) shows the spectral output for


180mW of injected pump power alone. At this
pump power level, a significant Raman
component was generated at ~1.38μm, and a
distinct soliton-like structure was present in an
autocorrelation trace of the output. By placing a
Fig.6.3.3. Typical autocorrelation
10nm band pass filter before the autocorrelator trace (background free) of Raman
solitons at 1.36µm, through
centred at 1.36μm, it was possible to completely seeding with 70mW of CW d iode
laser signal at 1.36µm for 180mW
suppress this structure completely, indicating of pump power at 1.32µm.

that no such structure was present at the seed


wavelength of 1.36μm. When the seed signal was once again introduced at a 70μW
power level as shown in Fig.6.3.2(d), the process was once again primarily initiated at
1.36μm, and soliton structures developed at and beyond this wavelength as evidenced
through autocorrelation.

For the 160mW pump power case (b), the largest pulse/pedestal contrast ratio
was obtained. Below this level, the gain in the laser diode signal fell off rapidly. The
threshold for amplifying a 100μW laser diode signal was measured by spectrally
resolving the band centred at 1.36μm using a 10nm bandpass filter, and was estimated

158
at 140mW of average pump power. The autocorrelation trace corresponding to the
situation shown spectrally in Fig.6.3.2(d) is shown in Fig.6.3.3., with a 10nm bandpass
filter centred at 1.36μm placed before the autocorrelator.

The trace indicates the presence of a 250fs pulse on a 3% intensity pedestal,


with ~28% of the total energy present in the pulse. Beyond the 10nm bandpass filter,
the average power contained in this amplified laser diode signal was 6.8mW,
corresponding to 1.8mW (~73W peak) being attributed to the pulse. The estimated
fundamental soliton power for a 250fs pulse at a wavelength of 1.36μm, with a
dispersion parameter |D|=5ps nm-1km-1 and a fibre mode field area of ~38μm2 is ~63W
peak. Hence within experimental error, the pulse shown in Fig.6.3.3. can be said to be
representative of a fundamental n=1 soliton. Note that in Fig.6.3.2(b). the spectral
width of the main Raman component was insufficient to support a 250fs soliton. The
narrow band part of the Raman signal was CW on a femtosecond timescale, and
therefore contributed to the autocorrelation pedestal.

The Fourier transform pulse duration limit imposed by the 10nm bandpass
filter would have the effect of limiting the measured pulse duration to about 200fs, if a
sech2 profile appropriate to soliton pulses is assumed. To ensure that the measured
pulses were not in fact shorter, the filter was replaced by a similar filter with 18nm
bandpass, yielding the same measured pulse duration. Provided that the pump power
was maintained at a180mW +/-5mW, the output pulses remained approximately n=1
solitons, tunable from 1.33 to 1.38μm. The spectra of Fig.6.3.2(d) indicates a FWHM
Raman bandwidth of ~20nm, which could theoretically support a 100fs soliton.
However no such short pulses were ever measured, using an unfiltered autocorrelator.
The reason for this is attributed to there being more than one soliton present in this
spectrum. The long wavelength shifted band around 1.38μm was in fact due to the
SSFS, and was emphasised for higher pump powers.

Fig.6.3.4 shows the measured temporal pulsewidth trend at 1.36μm for

159
increasing pump power, with a fixed 100μW signal power.

Fig.6.3.4. Variation of the measured soliton pulsewidth at the 1.36μm seed wavelength
with pump power, for a fixed seed signal power of 100μW.

For pump powers above 180mW, the pulsewidth rapidly increased as


spontaneous unseeded solitons were generated at the Raman gain peak around
1.38μm. These solitons collided with and broadened the solitons generated through the
seeding process during walk-off. As the pump power was increased further, more
spontaneous solitons were generated, and the broadening process became exaggerated
as the collisions occurred earlier in the fibre. This suggests that a narrow band of
pump powers existed over which the number and wavelength of the seeded solitons
could be controlled. Each soliton collision which occurred at higher powers gave rise
to a redistribution of energy between the two colliding pulses during coalescence. For
pulses at different wavelengths, one could provide Raman gain for the other, albeit
over a very short interaction distance. The spectrum of solitons during coalescence is
also much broader than the individual components, and hence SSFS is more likely to
occur. Energy loss of any kind results in soliton broadening, and so does SSFS, since
at higher wavelengths the dispersion parameter and hence fundamental soliton power
is higher. Although each interaction may have caused little broadening, multiple

160
collisions occurring over long lengths of fibre could have resulted in the ascribed
broadening behaviour.
For example, in the fibre used here, the dispersion parameter |D| ~5ps nm-1km-1
at 1.36μm. For two 250fs solitons spectrally separated by as much as 20nm, the
interaction length would correspond to ~2.5m. The Raman conversion threshold given
by Eq(1.5.4) over 2.5m would yield a peak power value of ~60W, which is an
attainable peak power in the case of a seeded interaction. There is therefore sufficient
peak power for energy conversion and broadening in the results presented here.
Broadened pulses which fall below the fundamental soliton power disperse away,
contibuting to the measured autocorrelation pedestal. This effect is illustrated in
Fig.6.3.5.

Fig.6.3.5. Variation of the measured soliton autocorrelation pedestal intensity as a


percentage of the peak intensity with pump power. The seed signal was at 1.36μm with
100μW average power.

Non-soliton components of the spectrum around 1.36μm and cross-correlations


between multiple solitons also contributed marginally to this pedestal as the power
was increased, in a similar manner to the soliton Raman continuum generation
discussed in section 6.2. The role of the laser diode seed signal appeared only to fix

161
the initial wavelength at which Raman gain occurred, reducing the conversion
threshold for one particular wavelength. Fig.6.3.6. shows how the seed signal power
level had very little influence on the generated soliton pulsewidth. The seed signal
only appeared to provide an alternative to growth from noise, or discontinuities arising
from imperfect pump pulses or MI, which normally self selects a wavelength for
maximum growth. The laser diode provided a more predictable and controllable seed
wavelength for the process, but the seed power made very little difference.

Fig.6.3.6. Variation of the measured soliton pulsewidth with laser diode seed signal
power at 1.36μm, for a constant pump power of 160mW.

For much higher launched pump powers, cascade into higher Raman Stokes
bands occurred, as indicated in Fig.6.3.7. The launched pump power level was
increased to 190mW average at 1.32μm, in optimised 80ps pulses.(24W peak). In the
absence of any laser diode seed signal, soliton generation occurred at 1.38μm, with no
significant component at 1.36μm, as shown in Fig.6.3.7(a).

162
Fig.6.3.7. Measured spectra with (a) 190mW average pump power (80ps pulses) at
1.32μm alone, and (b) with the addition of 160μW CW laser diode signal at 1.36μm.

When 160μW of 1.36μm CW laser diode radiation was added, severe pump
depletion occurred, and soliton generation was seeded at 1.36μm. This yielded 370fs
pulses within a 10nm, 1.36μm band, on a 30% intensity pedestal. The seeded pulses
experienced considerable SSFS, combined with generation of second Stokes radiation
at 1.48μm. By spectrally resolving this band, 220fs pulses on very low (~1%)
pedestals were measured by autocorrelation. This band was confirmed as a second
Stokes band since no variation of the centre wavelength occurred with pump power, as
would be expected from a signal undergoing SSFS.

In conclusion, by using a CW laser diode as a seed signal, efficient generation


of single solitons has been obtained at a predetermined wavelength tunable across the
Raman gain bandwidth. Single fundamental soliton operation could be routinely
obtained by maintaining the pump power in the fibre below the level for which
significant spontaneous Raman soliton generation occurred. This produced a source of
n=1 solitons of ~250fs pulse width, with low pedestals (3% of intensity) lasting
approximately the duration of the pump pulses (~100ps). Since the wavelength of the
pulses is fixed, the uncertainty in the pulse arrival time has probably been reduced to a
certain degree, however the use of a pulsed seed signal from e.g. a gain switched laser
diode would reduce this pulse to pulse jitter considerably, provided the jitter in the
source was <5ps.

163
6.4 Stimulated Raman Amplification of Noise Bursts into Solitons.

In section 2.2, the construction and characterisation of a synchronously


pumped, dispersion compensated fibre Raman ring laser operating around 1.4μm was
discussed. This ring laser was capable of producing 20mW(~100W peak) of average
output power in ~1.8ps pulses, from a 400m length of standard dispersion-shifted
(λ0=1.46μm) fibre, synchronously pumped with ~500mW of radiation at 1.32μm, in

100ps pulses. For a cavity length mismatch from the optimum of +/-10μm, the ring
laser generated pulses with the temporal structure characteristic of a burst of noise, as
shown in the autocorrelation trace of Fig.6.4.1.

Fig.6.4.1. Background free autocorrelation of the ring laser output for a cavity length
detuning of +10μm. Similar structure occurred for equivalent negative mismatch.

Although early experiments relied on launching transform limited pulses in


order to excite solitons, a pulse with any reasonable shape should evolve into a
soliton18, with any surplus energy appearing as dispersive noise in the fibre. In this
section, an experiment is described which investigates the effect of amplifying noise
components up to soliton powers, to demonstrate how soliton generation is the stable
limit for a variety of input conditions when propagation occurs in the anomalous GVD
regime. The results are of direct relevance to soliton transmission systems which

164
invlove periodic, highly pertubative gain, since they show that noise can develop, with
sufficient gain, into non-dispersive soliton signals, which may lead to considerable
system error.

To investigate the effects of noise-burst transmission in optical fibres, the


signal from the detuned ring laser described above were launched into a length of
dispersion-fibre, whilst synchronous Raman gain was used to amplify the noise
features up to soliton powers. Since the highly perturbative regime encountered in
soliton Raman generation (see section 6.2) results in evolution from MI or
discontinuities in the pump pulse into solitons, it is reasonable to expect that short
noise spike components of the ~100ps noise bursts generated by the ring laser would
themselves grow in intensity and form solitons.

The autocorrelation trace shown in Fig.6.4.1. represents (see appendix) that of


a finite burst of noise, with ~66ps envelope duration, containing noise features whose
average duration corresponds to the duration of the coherence spike, i.e. ~2.8ps. The
wavelength of the noise signal was centred at 1.39μm, with a bandwidth of ~2nm,
limited by the intra-cavity filter in the ring laser. Fig.6.4.2. shows the experimental
arrangement providing synchronous Raman gain for the noise signal.

Fig.6.4.2. Schematic of the experiment for amplifying noise bursts into solitons.

165
Beam splitter (BS1) (25%R at 1.32μm) divided the pump beam form the mode
locked Nd:YAG master pump laser. Approximately 400mW of the 1.8W average
(180Wpeak) power available at 1.32μm was directed via beamsplitter (BS1) and an
optical delay line onto beamsplitter (BS2) to provide a sycnchronous pump for the

noise signal. The 20-40mW average power noise signal at 1.39μm was derived from
the detuned ring laser, which in turn was pumped by the remaining 1.2W of pump
radiation passing through beamsplitter (BS1). Both 1.32μm pump and 1.39μm noise
signal were combined with dichroic beamsplitter (BS2) and focussed with a standard
X10 microscope objective (MO1) into 500m of fibre with a minimum dispersion
wavelength λ0=1.38μm. The fibre had a small anomalous dispersion parameter

|D|=2ps nm-1km-1, and a mode field area of ~64μm2 at 1.4μm. The output from the
fibre was collected and collimated by an identical lens (MO2) into a scanning

background free autocorrelator with 50fs resolution, and a 1/2m scanning


monochromator. The delay line in the pump beam was adjusted so as to maximise the
gain experienced by the noise signal through Raman amplification by the pump, by
monitoring the spectrum at 1.39μm. This was arranged by ensuring that the signal
slightly lagged the pump pulse temporally, as the signal had a slightly higher group
velocity and would therefore walk through the pump pulse. This delay was then fixed
throughout the experiment, once optimised by observation of the autocorrelation at
1.39μm.

The synchronous pump power was adjusted by insertion of a variable neutral


density filter wheel into the pump beam alone. For a maximum launched average
pump power of 80mW, no soliton generation evolved from the pump pulses alone,
since this power level was insufficient for single pass Raman soliton evolution to
occur. To avoid the possibility of soliton shaping occurring from the noise signal alone
in the absence of gain, the power level of the signal at 1.39μm was maintained below
1mW average in the amplification fibre. Under these conditions, an autocorrelation
similar to that shown in Fig.6.4.1. was obtained, indicating that no soliton shaping was

166
occurring at these relatively low signal powers. The autocorrelation trace of the signal
output from the fibre is shown in the presence of synchronous gain in Fig.6.3.4., for
(a) 45mW and (b) 65mW of pump power at 1.32μm. In (a), the trace of the
unamplified noise signal in the absence of pump power is shown on a similar scale for
reference near the baseline.

Fig.6.3.4. Background free autocorrelation traces of the Raman amplified noise signal
(1mW average power) at synchronous pump pwers of (a )45mW and (b) 65mW
average at 1.32μm.

In (a) clear amplification of the noise signal was evident, with an estimated
gain from the autocorrelation of ~3.5, assuming no change in pulse duration. The short
coherence spike component of the trace began to dominate the measurement, and
some compression of this component occurred from 2.8ps to 1.16ps. This was
consistent with the ultrashort structures in the noise signal undergoing soliton
compression through amplification, as some of the structures approached a peak
power level close to the fundamental soliton power. For 65mW of average pump
power shown in (b), a further decrease in the FWHM of the spike occurred to 820fs,
and the trace began to exhibit features typical of a pulse with a pedestal, rather than a
bursts of noise. The pulse like structure became more significant, and the
autocorrelation peak/pedestal ratio increased from the 2:1 associated with a noise burst
to 5:1. For the highest launched synchronous pump power, which was 80mW average
at 1.32μm, the autocorrelation trace shown in Fig.6.3.5. was obtained.

167
Fig.6.3.5. Background free autocorrelation trace of the amplified noise signal for
80mW pump power.

The FWHM duration of the pulse structure was now as low as 620fs, above a
pedestal which extended over the pump pulse duration, accounting for ~12% of the
autocorrelation peak signal. By consideration of the respective areas of the
autocorrelation function, it was estimated that ~22% of the energy at 1.39μm was
contained in the short pulse component. By measuring the average spectral power in
this band, the average power in the pulse structure was estimated to be 1.8mW. This
band experienced an average power gain of approximately 8 times.

The fibre used for amplification in this experiment had a dispersion parameter
|D|=2ps nm-1km-1 at 1.4μm. For a pulse of 620fs duration at 1.39μm, the fundamental
soliton power is ~5.7W peak. This would correspond to an average power of 0.36mW
at 100MHz, with a predicted soliton period of 93m. From these calculations, it is
apparent that more than one soliton was present in the amplified noise signal shown in
Fig.6.3.5. In this case, ~5 solitons were generated, and hence due to the random nature
of the soliton evolution, the autocorrelation trace represented an integrated cross-
correlation between these pulses, contibuting to the pedestal intensity. Relaunching the
amplified noise signal into a second fibre of over 1km length, and similar
characteristics to the first yielded similar though slightly longer pulses at the outptut.
Since this was significantly longer than the soliton period (Z0~93m), which also
characterised the sub-fundamental soliton dispersion length LD from Eq(1.4.32), the

168
amplified noise signal had definitely partially evolved into soliton pulses. The pedestal
on the autocorrelation was retained, indicating that some sub-fundamental soliton
power structure was still present after propagation over 1km, as was expected.

Spectrally, an important feature of the noise amplification process was the


complete absence of any SSFS, which would normally be associated with the
propagation of femtosecond solitons. The most reasonable cause of this was that the
solitons evolve in the prescence of considerable narrowband gain from the Raman
process19, which has been shown to suppress SSFS. The spectral width of the input
noise signal originally corresponded to the inverse of the coherence spike duration,
i.e.~2.3nm, however in the presence of Raman gain and soliton shaping the spectrum
broadened to 3.8nm bandwidth. This is consistent with the noise spikes evolving into
solitons of a shorter duration, but this value yields a time-bandwidth product of
ΔνΔτ=0.37, which is in excess of that expected for soliton pulses (ΔνΔτ=0.314 for
sech2 pulses). This can only be due to the evolution of more than one soliton, and the
effect of SSFS. This is consistent with the measured autocorrelation, indicating that
several pulses are present.

In conclusion, femtosecond soliton generation by synchronous Raman


amplification of noise bursts with envelope durations of ~100ps and noise components
of ~3ps has been demonstrated, illustrating the attractive property of the soliton
solution in the anomalous GVD regime. This result has serious consequences for noise
propagating in long distance soliton based transmission systems, and indicates that
amplified spontaneous noise could grow in the presence of relatively low gain into
soliton pulses, possibly causing spurious errors. A method of suppressing low
intensity intra-pulse noise is presented in section 7.2, which may result in the
suppression of such a possibility.

169
6.5 Stimulated Raman Amplification of Dispersive Pulses

into Solitons.

The distributed loss intrinsically present in an optical fibre can be compensated


for by the application of gain in a similar distributed mode. In the limit where the
soliton number n is more or less unperturbed from unity during the transmission and
amplification process, the scheme is said to be adiabatic. Quantitatively, this can be
expressed as the limit where the characteristic length of the amplification period or
gain length (α-1) is much longer than that of the soliton period (Z0), such that
αZ0<0.0521. This has been shown to result in no net change in pulse shape over

6000km of transmission fibre21a,, with zero net signal loss and less than 8% energy
fluctuations. To reduce the Raman gain perturbation still further, quasi-CW gain can
be provided by periodic bidirectional pumping via dichroic fused couplers, resulting in
very small pulse energy fluctuations. In this way the total Raman gain becomes the
sum of two exponential functions, one decaying whilst the other grows.

An alternative experimental situation is presented here, to investigate the effect


of higher gain (αZ0>>0.05) on a pulse whose intensity is far below the required

fundamental soliton power. It is similar in principle to a system of amplification


employing very high gain and periodically spaced repeaters or lumped amplifiers, and
has been extensively investigated elsewhere10. The gain employed here is at least
double that required in lumped amplification systems under current investigation,
operating in the regime where αZ0>0.3.

The experimental configuration for this investigation was similar to that


presented in the last section (6.4) for amplifying noise signals. In this case, the Raman
ring laser (characterised in chapter 2) was optimised to produce near transform limited
pulses of ~1.8ps duration with a ~2nm bandwidth around 1.39μm.. Average powers of
up to 40mW were available, representing a pulse peak power of ~200W peak at
100MHz repetition rate. The ring laser output signal was first launched into a test fibre

170
to ascertain the fundamental soliton power. The fibre used here was 140m long, with a
minimum dispersion wavelength λ0=1.27μm, and an anomalous GVD parameter

|D|=15ps nm-1km-1 at 1.4μm. The mode field area of the fibre was estimated to be
~80μm2 at 1.39μm, resulting in a predicted fundamental soliton power of 8.1W peak
(~1.5mW average) for the ~1.8ps signal pulses. The predicted soliton period was
calculated to be 105m, and hence signal pulses would exhibit significant dispersion
below the fundamental soliton power. The output from the optimised Raman ring laser
is shown in the background free autocorrelation trace of Fig.6.5.2(a).

Fig.6.5.2. Autocorrelation traces of (a) output signal pulses from Raman ring laser
and from amplification fibre for (b) 12mW and (c) 2.5mW of average launched signal
power. In (b) & (c) the signal pulses are shown as an outer trace for reference.

This trace indicates a signal pulsewidth of ~1.8ps. Fig.6.5.2(b) shows the same
input signal pulse (outer trace) and the autocorrelation trace of the output from the
fibre for a launched average signal power of 12mW in the absence of any pump
power. The trace shows a clear high order soliton compression to 1.3ps, indicating that
the signal power exceeds the fundamental soliton power, and is in fact exciting a high
order soliton. This measured duration is considerably longer than the optimum pulse
duration occurring through high order soliton compression (~110fs), because this
compression occurs at a point ~27m from the input to the fibre.

171
By reducing the launched signal power to 2.5mW average, the autocorrelation
trace show in Fig.6.5.2(c) was obtained. This shows that the input and output signal
pulses are of approximately the same FWHM duration, i.e. ~1.8ps, with only a slight
envelope disparity in the wings of the trace due to a deviation of the launched signal
pulses from the ideal sech2 shape. This similarity in FWHM duration persisted for
average powers as low as 1.5mW, in good agreement with the predicted fundamental
soliton power.

For the amplification investigation, the test fibre was replaced with a 600m
length of fibre with a dispersion minimum wavelength λ0=1.38μm, and an anomalous

GVD parameter |D|=3ps nm-1km-1. The mode field area for this fibre was 95μm2,
yielding a predicted fundamental soliton power of 1.97W peak, or 350μW average for
the 1.8ps signal pulses at 100MHz, with a predicted soliton period of 525m. Once
again, below this power level, noticeable dispersion was expected to dominate over
the 600m fibre length. The calculated walk-off length between the 1.8ps signal pulses
at 1.39μm and the 100ps pump pulses at 1.32μm was ~400m, indicating that the gain
length and soliton period were of similar magnitude, with αZ0~1. The gain process

was thus still approximately adiabatic, and hence the soliton was expected to adjust its
duration to match its amplified energy accordingly. High order solitons would only be
excited for αZ0>>1, thus the gain length would have to have been comparable with the

soliton period, which for the pulses used here is practically impossible to achieve
using Raman gain without causing loss to higher order Stokes bands.

A maximum of 80mW of average pump power in 100ps pulses at 1.32μm was


simultaneously launched into the 600m amplifier fibre, in synchronism with the signal
from the Raman ring laser. The delay line was adjusted to optimise the interaction
length between the signal and pump pulses, by maximising the gain whilst monitoring
real time autocorrelation. Fig.6.5.3(a) shows the fibre output for an average launched
signal power of 250μW from the Raman ring laser alone.

172
Fig.6.5.3. Autocorrelation traces of the output from the amplification fibre for an
input signal of 1.8ps duration at 250μW average power (a) without amplification and
(b) with 80mW of synchronous pump power at 1.32μm.

A clear dispersive broadening of the signal pulses occurred in the absence of


any gain, with a resultant FWHM output duration of 5.5ps. In Fig.6.5.3(b). the output
signal is shown amplified by 80mW of synchronous pump power at 1.32μm, for an
equivalent input signal level. Compression of the pulse envelope occurred to 700fs
duration, with an amplified average signal power (measured using narrow bandpass
filters) of 3.5mW at 1.39μm. This represented an average power gain of 14 times, or
~12dB. On reducing the pump power to ~20mW, the output pulse duration increased
to ~1.5ps, approximately equivalent to that of the input signal pulse, and hence
corresponding to ~n=1 soliton propagation. This yielded an average output power in
the 1.39μm band of 500μW (3.3W peak), which is in reasonable agreement with the
predicted fundamental soliton power for a 1.5ps pulse, which is 2.83W peak.

Theoretical investigation of amplified pulse propagation in the anomalous


dispersion regime26 predicts that solitons compress as their energy increases according
to the relation:

d τ-1 = α τ-1
dz

where α is the gain coefficient of the amplification process (Raman in this case, which
is proportional to pump power) and τ is the pulse duration. The gain saturates during
Raman amplification, limiting α, and therefore limiting the minimum pulsewidth
obtainable through Raman amplification. Thus the limitiation in the compression

173
achieved in this experiment is ascribed to gain saturation. Gain from e.g. Erbium
doped fibre amplifiers could perhaps provide an even higher compression factor, and
over shorter gain lengths enable investigation in the non-adiabatic regime with
αZ0>>1, where high order soliton excitation would be expected.

The absence of any SSFS effects in the spectral output of the amplifier was
once again ascribed to the prescence of considerable narrow band gain from the
copropagating pump field, which has been shown to suppress the effect through
frequency pulling20. The amplified pulses were verified to be solitons by relaunching
into ~3km of fibre, with only slight temporal broadening occurring due to coupling
losses and fibre parameter differences.

In conclusion, the reconstruction of dispersed sub-fundamental soliton power


pulses has been demonstrated through synchronous Raman gain. Stable soliton
evolution has been shown to occur throughout the gain process, without breakup due
to the high gains involved, despite the comparable gain length and soliton period. The
technique is directly appropriate for amplification of laser diode signals up to soliton
powers, with the added advantage of pulsewidth reduction through soliton
compression. The resilient properties of pulses in the soliton regime has been
demonstrated in the presence of strong perturbations such as high gain.

6.6 References

(1) V.A.Vysloukh & V.N.Serkin. Pis'ma Zh.Eksp.Teor.Fiz. 38,170(1983) U U

(2) E.M.Dianov, A.Ya.Karasik, P.V.Mamyshev, A.M.Prokorov, V.N.Serkin,


M.F.Stelmakh & A.A.Fomichev. Pis'ma Zh.Eksp.Teor.Fiz. 41,42(1985) U U

(3) R.G.Smith. Appl.Opt. 11,2489(1972)


U U

(4) R.H.Stolen. C.Lee & R.K.Jain. J.Opt.Soc.Am. B1,652(1984)


U U

(5) R.H.Stolen & E.P.Ippen. Appl.Phys.Lett. 22,276(1973)


U U

(6) A.S.Gouveia-Neto, A.S.L.Gomes & J.R.Taylor.


IEEE J.Quant.Elec. QE-24,332(1988)
U U

(7) E.M.Dianov, A.Ya.Karasik, P.V.Mamyshev, A.M.Prokorov, V.N.Serkin,


M.F.Stelmakh & A.A.Fomichev. JETP Lett. 41,294(1985)
U U

174
(8) A.S.Gouveia-Neto, A.S.L.Gomes & J.R.Taylor. Opt.Lett. 12,1035(1987) U U

(9) R.H.Stolen, J.P.Gordon, W.J.Tomlinson & H.A.Haus.


J.Opt.Soc.Am. B6, 1159(1989)
U U

(10) Y.Kodama & A.Hasegawa. Opt.Lett. 7,339(1982) U U

(11) A.S.Gouveia-Neto, M.E.Faldon & J.R.Taylor. Opt.Lett. 69,325(1989) U U

(12) M.Nakazawa, K.Suzuki, H.Kubota & H.A.Haus. Phys.Rev. A39,5768(1989) U U

(13) A.S.Gouveia-Neto, M.E.Faldon & J.R.Taylor. Opt.Lett. 13,1029(1988) U U

(14) U.Keller, K.D.Li, M.Rodwell & D.H.Bloom.


IEEE J.Quant.Elec. QE-25,280(1989)
U U

(15) E.J.Greer, D.M.Patrick, P.G.J.Wigley, J.I.Vukusic & J.R.Taylor.


Opt.Lett.15,133(1990)
U U

(16) R.H.Stolen & E.P.Ippen. Phys.Rev.Lett. 48,805(1982) U U

(17) A.S.Gouveia-Neto & J.R.Taylor. Elec.Lett. 24,1544(1988) U U

(18) A.Hasegawa & Y.Kodama. Proc.IEEE. 69,1145(1981) U U

(19) F.M.Mitschke & L.F.Mollenauer. Opt.Lett. 11,659(1986) U U

(20) K.J.Blow, N.J.Doran & D.Wood. J.Opt.Soc.Am. B5,1031(1988) U U

(21) L.F.Mollenauer, R.H.Stolen & M.N.Islam. Opt.Lett. 10,229(1985)


(21a) L.F.Mollenauer & K.Smith. Opt.Lett. 13,675(1988) U U

(22) A.Hasegawa. Opt.Lett. 8,650(1983)


U U

(23) A.Hasegawa. Appl.Opt. 23,3302(1984)


U U

(24) L.F.Mollenauer, J.P.Gordon & M.N.Islam.


IEEE J.Quant.Elec. QE-22,157(1986)
U U

(25) L.F.Mollenauer & R.H.Stolen. Opt.Lett. 10,229(1985) U U

(26) K.J.Blow, N.J.Doran & D.Wood. J.Opt.Soc.Am. B5,381(1988) U U

175
7. NONLINEAR OPTICAL LOOP MIRROR PULSE PROCESSING

7.1 Introduction

The optical Kerr effect was introduced in chapter 1 which, to a first order
approximation, induces a nonlinear phase shift in the field across a pulse in the
presence of an intensity dependent refractive index change. The local phase of a pulse
undergoes a shift which is dependent on the local intensity, and as the intensity rises
and falls this results in a change in phase. A change in phase results in a frequency
chirp across the pulse, referred to as Self Phase Modulation (SPM). If the phase shift
is induced through the presence of another pulse of sufficient intensity, then Cross
Phase Modulation (XPM) occurs (see chapter 5).

Hence the phase shift (φ) across an intense pulse in an optical fibre becomes
intensity (I) dependent . Over a length of fibre L with coefficient of nonlinear
refractive index n2, Eq(1.4.9) gives:

Δ ∅ = n2 I k L (7.1.1)

where k=2π/λ is the wavenumber for a pulse centred at wavelength λ. If in the first
instance a CW wave is considered, Eq(7.1.1) shows that over moderate lengths of
fibre, a unity phase shift can be achieved for relatively low average powers. The
length at which Δφ becomes unity is in fact the nonlinear length LNL defined in
Eq(1.4.11). At 1.06μm, a fibre of 100m with a mode field area Aeff=38μm2 would

induce a π phase shift for 6.4W of input peak power. If two beams with different
relative intensity dependent phases overlap and interfere, the intensity of the resultant
field would clearly be sensitive to the parameters in Eq(7.1.1). By fixing the length,
for example, an intensity dependent switch could be constructed.

176
A number of both theoretical and experimental devices have been proposed for
switching signals in a fibre communications system, based on routing and logic
operations1 - 4. The devices rely on nonlinearity through the optical Kerr effect, which
provides a very fast response time in the order of 10fs. However the pulsed response
of such devices is not ideal, since the change in envelope intensity across a pulse gives
rise to a range of phase shifts, and hence signal pulse fragmentation. In section 2 of
this chapter the operation of such a device based on a Nonlinear Optical Loop Mirror
(NOLM) is described, both for non-soliton and soliton signal pulses4, illustrating how
a NOLM can provide complete pulse switching without fragmentation. In section 3, an
experiment is described which demonstrates how optical pulses can be 'cleaned up'
using the intensity discriminating nature of the NOLM.

7.2 Intensity Dependent Transmission through a

Nonlinear Optical Loop Mirror

A nonlinear fibre Sagnac interferometer or Nonlinear Optical Loop Mirror


(NOLM) consists of a length of fibre which is terminated at the ends with a single
common fused coupler. Fig.7.1. shows the schematic configuration of a NOLM

Fig.7.2.1. Schematic representation of the Nonlinear Optical Loop Mirror (NOLM)

The fused coupler is normally of the twisted evanescent type, with a power
splitting ratio α:(1-α), where {0<α<1). Due to the symmetric nature of the device, the
common optical path length for an input field E1, split by the coupler into fields E3 and
E4, is the same. This is similar in concept to an anti-resonant ring interferometer5, and

avoids the requirement for interferometric alignment of the path length difference.

177
This results in a linear response similar to that of a mirror for α=0.5 exactly, with an
incident field E1 reflected back as E01, with zero output in E02.

If the symmetry of the loop is broken in any way, so as to introduce a relative


phase advance of E3 with respect to E4 or vice-versa, the loop mirror will begin to
transmit light through field E02, depending on the relative phase shift of the two fields

in the loop. This is achieved by adjusting the splitting ratio of the coupler, to be
different from α=0.5, so that fields E3 and E4 have different intensities. This intensity

difference between the counter-propagating fields results in a differential phase


difference, which is dependent on the intensity of the incident beam E1. This results in

an overall intensity dependent transmissivity/reflectivity in the NOLM. The two fields


in the loop achieve a relative phase difference4 given by:

Δ ∅ = 2π n2 L E3 2 - E4 2
λ (7.2.2)

For an evanescent coupler with power splitting ratio α:(1-α), the fields in the
loop are described as follows:

E3 = α E1 + i 1-α E2
E4 = i 1-α E1 + α E2
(7.2.3)

Now for a single input Ein, E2=0. The two fields generated after propagation around

the loop and before recombination in the coupler are:

2
E3 = α Ein exp i α Ein 2π n2 L/λ
2
E4 = i 1-α Ein exp i 1- α Ein 2π n2 L/λ (7.2.4)

These equations represent the through coupled and cross coupled fields derived
from Ein, with the addition of a phase factor induced through the nonlinear phase shift

(SPM). The phase difference between the two fields increases as α deviates from 0.5.

178
The total transmitted intensity through the loop mirror is given by :

2 2
E02 = Ein 1 - 2α 1-α 1+COS 1-2α Ein 22π n2 L/λ
(7.2.5)

The input signal is transmitted through the NOLM when |E02|2 = |Ein|2, i.e. when
(1-2α) |Ein|2 2π n2 L/λ=π . Hence, for any value of α, 100% transmission through the

NOLM is obtained if

2π n2 L/λ Ein = mπ
2
m=(1,3,5.....)
1-2α
(7.2.6)

and 100% reflection is obtained whenever m =(0,2,4....). Note that the switching
contrast, given by the modulation depth of the function |E02|2 / |Ein|2, increases as α

approaches 0.5, but the intensity required to reach switching (m=1) increases too, so a
trade-off exists between switching power and obtainable contrast. For example, for a
NOLM consisting of a coupler with α=0.4, and a fibre of 100m length with a mode
field area of 38μm2, the transmissivity will increase from 4% to 100%, for an increase
in coupled input power from 0 to 31.9W (accompanied by a complementary change in
the reflectivity fom 96% to 0%). The transmission function of a NOLM versus input
power for α=0.4 is shown in Fig.7.2.2. The axes show normalised units, with n2L/λ=1.

31.9W peak represents a relatively high CW switching power. However this


level is easily obtained in the pulsed regime, and can be substantially reduced by using
longer loop lengths. Instability and sensitivity to environmental fluctuations, such as
temperature, do not effect the Sagnac interferometer configuration1,3,4 since the signal
paths are common. The switching contrast ratio may be compromised through
polarisation modification over long fibre lengths, however this may be improved
through the use of polarisation maintaining fibres in the loop or the incorporation of
polarisation strainers6.

179
Fig.7.2.2. Transmitted power vs input power for NOLM with α=0.4 and n2L/λ=1.The
solid lines represent the minimum and maximum values obtainable.

Since the response of the electronic Kerr effect which induces SPM in silica
optical fibres is very fast (~10fs), this device offers the possibility of extremely high
frequency switching and signal processing. Therefore, the transient response of the
NOLM is important. The response to a pulsed input signal is, however, sensitive to the
instantaneous intensity of the pulse, resulting in different transmission function values
across the pulse profile. Only perfectly square pulses with durations short compared
with the loop transit time will give rise to a transmission function similar to the ideal
CW case considered up to now, and illustrated in Fig.7.2.2. For realistic bell-shaped
pulses output characteristics similar to that shown in Fig.7.2.3. are obtained.

Clearly the high contrast and transmission characteristics have been


compromised. This occurs through incomplete or partial pulse switching. If the peak
of a pulse has an intensity corresponding to 8 units in Fig.7.2.3., portions of the wings
with intensity 5 units will be reflected, as well as portions with intensity 1 unit.
Significant transmitted pulse shaping can occur through this device, and in section 7.3.
an experiment is described which exploits this characteristic.

180
Fig.7.2.3. Transmitted energy (or average power) vs input energy for a typical bell
shaped (sech2) pulse. The CW case of Fig.7.2.2. is shown dotted for reference.

Square shaped pulses experience complete pulse switching through a NOLM


because the accumulated phase shift across the pulse, which is intensity dependent, is
constant. In section 1.4., the concept of soliton propagation was introduced in the
anomalous GVD regime. The nonlinear spectral broadening due to SPM was shown to
support bound state soliton solutions of the form u=n sech(τ), where n is an integer
called the soliton number. The lowest order or n=1 soliton has a stationary envelope
function given by Eq(1.4.26):


u(ξ, τ) = sech(τ) exp
2 (7.2.7)

Higher order solitons are more complex, and have an envelope function which
periodically repeats with propagation in ξ with period π/2 (the soliton period). The
phase factor in the n=1 solution shown in Eq(7.2.7) φ=ξ/2 applies to the phase of the
whole pulse, and this indicates that solitons exhibit the special property of aquiring a
uniform phase as they propagate. The effect of soliton shaping in a NOLM can be
included if dispersion is considered, which imposes a requirement that the loop length

181
be of comparable length to the soliton period Z0 (or dispersion length LD) in order for

the phase shift imposed through SPM to be communicated throughout the pulse. This
length can be reduced by employing short pulse durations, or a high anomalous GVD
parameter. In general, the phase of a soliton evolves linearly with propagation
distance3 in accordance with Eq(7.2.7), and linearly with the amplitude of the soliton,
independently of the evolution of the soliton envelope function. This property makes
soliton pulses ideal signals for switching in a NOLM device, which would operate on
the whole pulse.

Neglecting pulse shaping effects through propagation through the loop and
XPM coupling between the counter-propagating pulses, the total energy output of the
NOLM for an input soliton u=Asech(τ) becomes4:

Eout = 2A2 1- 2α 1-α 1+ COS 2Z 1-2α A2 + α - 1-α A


(7.2.8)

where Z=L/LD is the loop length in soliton units, and LD is the dispersion length which
is related to the soliton period Z0 by Eq(1.4.32). A=1 refers to the single or

fundamental soliton. For a fixed loop length, the soliton phase evolves with amplitude,
hence shorter pulses will achieve switching at lower average power levels, since the
phase evolves faster for a shorter soliton period. The energy of the soliton pulses
determines the reflectivity of the NOLM. Fig.7.2.4. shows the transmission
characteristics of a (α=0.42) NOLM as a function of input pulse energy for CW fields
or square pulses (ideal), bell-shaped non-square pulses (approximate), and solitons.

182
Fig.7.2.4. NOLM transfer function vs input energy for α=0.42, indicating ideal CW
response for square pulses, non-ideal response for bell-shaped pulses, and restored
ideal response for solitons.

The ideal CW or square pulse response is clearly restored in the simulation


shown in Fig.7.2.4. Deep switching contrast is restored for soliton energies
comparable to non-soliton pulse energies. In the soliton NOLM configuration, the loop
length should be at least twice the soliton period, and hence short pulses produce
better performance, since shorter pulses have a shorter soliton period from Eq(1.4.32).

7.3 Pedestal Suppression and Nonlinear Pulse Shaping

employing a Nonlinear Optical Loop Mirror

As shown in Fig.7.2.2. the NOLM has an intensity dependent response which


is periodic with input peak intensity for non-soliton pulses. Since the phase
accumulated at a pulse peak is larger than that in the wings, significant pulse shaping
can occur across a pulse in transmission through a NOLM. Fig.7.3.1. shows computer
simulations of the predicted transmission of a bell-shaped (sech2) pulse profile (top)
and the associated background free autocorrelation functions (bottom) for increasing
peak power, in units consistent with Fig7.2.2. (n2L/λ=1).

183
Fig.7.3.1. Intensity profiles (top) and corresponding background free autocorrelation
functions (bottom) of (a) sech2 input pulses, and transmitted pulses through NOLM for
input peak powers of (b) 3, (c) 4, (d) 5, (e) 6, and (f) 8 normalised units (n2L/λ=1)
(Timescale arbitrary).

Fig.7.3.1.(a) shows the input sech2 pulse profile, and its autocorrelation (the
latter being broader by a factor 1.55; see appendix). In (b), the transmitted pulse

184
through a NOLM with α=0.4 is shown for an input peak power of 3 normalised units.
The peak of this pulse therefore coincides with the first transmission maximum of the
curve shown in Fig.7.2.2., with a resultant transmitted pulse profile which is
temporally shorter, and shows less energy in the pulse wings. The output pulse is also
squared off, because the transfer function changes only slowly near the transmission
peak, resulting in a triangulated autocorrelation trace, shown below. For higher input
peak powers (c)-(f), the pulse undergoes significant splitting and reshaping, whilst the
total energy transfer through the system represented by the area of the pulses shown,
fluctuates only slightly in accordance with the curve shown in Fig. 7.2.3. The
fractional energy transfer corresponding to the pulse shapes in Fig.7.3.1. was (from (b)
to (f)) 0.744, 0.601, 0.396, 0.380 and 0.653. respectively. The inadequacy of the
device in the normal dispersion regime for switching based on pulse energy is clear
from these simulations.

From these simulations, a pulse with 3 normalised units of peak power, as


shown in Fig.7.3.1(b), undergoes slight compression and pedestal suppression. In
Fig.7.3.2, a simulation is presented in which a sech2 pulse with an input peak power of
2.5 normalised units and a significant (10%) noise pedestal (a) is transmitted through a
NOLM. The transmitted output pulse is shown in Fig.7.3.2(b). Autocorrelations of the
respective signals are presented below for later reference.

Fig.7.3.2(a) sech2 intensity profile with a 10% noise component (top) and
corresponding autocorrelation function (bottom). The filtered versions transmitted
through the NOLM are shown in (b) for 2.5 normalised units of peak power, with

185
n2L/λ=1.

A clear suppression of the pedestal noise occurs in the simulation,


accompanied by a slight temporal compression in transmission through the NOLM.
75% of the incident energy is transmitted in this simulation, exhibiting very low loss
to yield clean pulses from the noisy input signal. The behaviour decribed above is
similar to that exhibited by a saturable absorber, whose absorption is linear until
saturation occurs, at which point transmission occurs. In fact the reflected mode of the
NOLM provides the characteristic of an inverse saturable absorber simultaneously; no
energy is actually absorbed or dissipated in this device.

To confirm the predictions of the preceding simulations, the following


experiment was conducted. Signal pulses were provided by a synchronously pumped,
mode locked NaCl:OH Colour Centre Laser (CCL), which produced 10-15ps pulses at
around 1.59μm (the peak of the crystal tuning range) at a 76MHz repetition rate. A
standard background free autocorrelator (non-collinear - see appendix) and a scanning
Fabry-Perot interferometer were used to characterise the signal pulses, which when
optimised had a time-bandwidth product close to the ideal value of ΔνΔτ=0.32 for
sech2 pulses. The experimental configuration is shown in Fig.7.3.3. The signal pulses
were first coupled into a coupler providing a ~25% output tap, in order to measure the
forward coupled power into the NOLM.

Fig.7.3.3. Experimental schematic for investigating the noise-suppressing properties


of a NOLM.

186
The loop comprised a low loss directional coupler with power splitting ratio
α=0.52, and 200m of dispersion-shifted (λ0=1.62μm, |D|=1.6ps nm-1km-1 at 1.59μm)

non-polarisation preserving fibre, which was fusion spliced between the output ports
of the coupler. Splicing losses of ~1dB were incurred through splicing fibres with
different cross-sections. Since the fibre loop did not preserve polarisation, to ensure
efficient interference of the counter-propagating fields in the coupler, a strain-induced
birefringent polarisation rotator was inserted into the low power (α=0.48) arm of the
loop. This introduced a lump power loss of ~3dB, and was therefore the dominant
symmetry breaking element in the NOLM, effectively increasing the intensity
difference in the counter-propagating fields dramatically.

For a lump loss of power transmissivity TL in the fibre loop, the transfer

function of the normally dispersive (non-soliton) NOLM given in Eq(7.2.5) requires


modification thus:

2
E02 Pout
2
= = TL 1 - 2α 1-α 1+COS 1-α TL- α φNL
Ein Pin

Pin 2π n2 L
where φNL =
λ Aeff
(7.2.8)

For no lump loss, i.e. TL=1, this reduces exactly to Eq(7.2.5). This results in an
overall attenuation in transmission by a factor TL, but a proportionate reduction in the

required switching power. This also allows the use of an α=0.5 coupler, with its
associated excellent switching contrast. The power transfer characteristic of a NOLM
with α=0.52 is shown in Fig.7.3.4.for a lump loss of both TL=0 and TL= 0.5 (3dB),
once again normalised by setting n2L/λ=1, and Aeff=1.

187
Fig,7.3.4. Power transfer characteristic of a NOLM for α=0.52, with and without a
lump loss TL=0.5 in the low power (α=0.48) arm. Units are normalised with
n2L/λAeff=1.

The normalised switching power, Ps, required for the transfer function to reach

the first transmission maximum in the absence of loss is given from Eq(7.2.6) by
(TL=0):

Ps 2π n2 L
= π
λ Aeff 1-2α

∴ Ps = 1
2 1-2α (7.2.9)

Hence for α=0.52, Ps=12.5 normalised units in the absence of lump loss, in

accordance with Fig(7.3.4). If the effect of the lump loss is considered, the normalised
switching power Ps is reduced to:

Ps = 1
2 1-α TL- α
(7.2.10)

188
For α=0.52 and TL=0.5, this yields a switching power Ps~1.8 units, as shown in

Fig.7.3.4., which is substantially lower than the lossless case. An interesting


alternative is presented if a lump gain is substituted for a lump loss. In this scenario,
TL is replaced by the gain coefficient TG, which would naturally operate in the high

power output arm of the coupler in the NOLM, so as to enhance any intrinsic intensity
imbalance. The transfer function of a NOLM with α=0.52, and a lump gain of TG=2 in

the high power (α=0.52) arm is shown in Fig.7.3.5.

Fig.7.3.5. Power transfer characteristic of a NOLM for α=0.52, with a lump gain
TG=2 in the high power (α=0.52) arm. Units are normalised with n2L/λAeff=1.

The input power Ps required to achieve the first transmission maximum in the
presence of a lump gain TG is given by:

Ps = 1
2 α TG- 1-α
(7.2.11)

The intensity difference in the counter-propagating beams is dominated by the


gain, since it amplifies the signal in one direction before it propagates and acquires a
much higher nonlinear phase shift. The presence of loss or gain clearly reduces the
required input switching power, whilst it yields a high contrast ratio by use of a

189
coupler with α close to 0.5. The presence of gain enhances the signal in transmission
as well, making the device suitable for applications which require a high signal fanout.
Fig.7.3.6. is a plot of the switching power Ps required to obtain the first transmission

peak, versus the gain in a NOLM for various values of α.

Fig.7.3.6. Switching power required Ps vs gain expressed in dB, for a NOLM


containing a lump gain, for various values of power splitting ratio, α.

For moderate values of gain, e.g.10dB, the power splitting ratio, α, has very
little influence on the switching power Ps, whilst high contrast switching can be

achieved with α close to 0.5. In the example given for α=0.5 and a 10dB lump gain,
the switching power Ps is as low as 0.11. For a typical fibre with a 50μm2 mode field

area this represents a peak power - length product of ~260Wm, requiring only 2.6W
of peak power for switching in a 100m NOLM.

In the experiment described here, a lump loss of 3dB (TL=0.5) was presented

by the polarisation rotator. The fibre mode field area was ~50μm2, and hence the
predicted switching power - length product was 4185Wm. This corresponded over the
200m length of fibre to a 21W peak switching power. The NOLM transmission with
α=0.52 was expected to vary from a maximum of 50% to a minimum of ~0.2%,

190
yielding a very high contrast as illustrated in Fig.7.3.4. Throughout the experiment,
the polarisation rotator was adjusted so as to obtain maximum reflectivity of the
NOLM in the low intensity (linear) regime, by reducing the launched signal power to
an absolute minimum, and monitoring the unused (backward propagating) arm of the
power tap coupler. Fig.7.3.7. shows the experimentally observed autocorrelations of
the NOLM output as the launched signal power from the CCL laser was increased at
the NOLM input. The peak powers indicated were calculated from those measured
including the coupling losses incurred throughout the experiment.

Fig.7.3.7. Autocorrelation traces of (a) input signal pulses from CCL laser, and
transmitted pulses through the NOLM for launched peak powers indicated.
The input signal pulses from the CCL laser are shown in trace (a) for
comparison with the output pulses (b) - (f). For 21W launched peak power (b) a

191
triangular trace was observed, in accordance with that predicted in the simulation
shown in Fig.7.3.1(b). In fact all of the results were consistent with the predicted pulse
shaping effects, up to those shown in Fig.7.3.7(f), which represented the maximum
launched peak power obtainable. The results indicated a measured switching power
Ps~21W peak, which is identical to the predicted value when losses are considered.

To illustrate the pedestal suppression property of the NOLM further, which is


evident in Fig.7.3.7(b), the cavity length of the CCL signal laser was detuned from
optimum. A detuning of less than 10μm produced poorly mode locked pulses with
significant pedestals and inter-pulse radiation, as shown in Fig.7.3.8(a). Provided that
the peak power of the pulses was approximately Ps=21W, the transmitted signal

through the NOLM gave a measured autocorrelation trace shown in Fig.7.3.8(b). The
device transmitted a narrower, relatively pedestal-free pulse, and clearly exhibited
properties suitable for suppression of low intensity noise and inter-pulse radiation.

Fig.7.3.8. Background free (non-collinear) autocorrelation traces of (a) input pulses


and (b) pedestal-free compressed pulses transmitted through the NOLM. (5ps/div
horizontal scale, and different arbitrary vertical scales).

In section 7.2, the response of the NOLM in the soliton (anomalous GVD)
regime was described, where it exhibited complete pulse switching, rather than the
partial pulse switching and reshaping illustrated above. In the soliton regime, the
pedestal suppression property of a NOLM is therefore removed, as any pulse which
develops into a soliton would undergo either complete transmission or reflection.

192
Nevertheless, the usefulness of such a device is not impaired, since the device would
be expected to operate as a soliton filter, allowing transmission of only soliton pulses
and reflecting inter-pulse radiation. This would require the incorporation of
anomalously dispersive fibre in the loop, with a length at least twice that of the soliton
period.

To investigate this possibility, optimised CCL laser pulses were injected into a
NOLM consisting of 200m of fibre with a dispersion minimum around 1.3μm. The
now anomalous GVD parameter |D| was ~18ps nm-1km-1 at 1.59μm, and the predicted
fundamental soliton power for the ~13ps pulses in the fibre was ~200μW peak power.
A typical launched peak power of ~20W represented excitation of a very high order
(n~300) soliton in the fibre, with a soliton period of Z0~3.5km. The loop length

(200m) was only a small fraction of the soliton period, and hence complete switching
of the entire ~13ps soliton was not possible. Since the switching power of this NOLM
was extimated to be Ps~150W peak, pulse shaping or indeed transmission of the ~13ps

was not expected. However since the launched pulse formed an n~300 soliton, high
order soliton formation was expected operate on the pulse envelope, compressing it to
less than 100fs, beyond which, pulse breakup was expected to occur. A femtosecond
soliton component in the loop would hence acquire sufficent peak power to cause
complete pulse switching of itself.

Fig.7.3.9. Autocorrelation of the NOLM output, illustrating transmission of the


compressed multisoliton component, shown below the input pulse from the CCL laser.
(Different vertical scales).
Fig.7.3.9. shows an autocorrelation of the NOLM output when ~13ps pulses

193
were launched in the soliton regime. The input pulse shape is shown for reference.
Note that the only component that switched through the NOLM was the short
component arising from the high order soliton breakup. Dispersive energy which
would normally be present in the process was effectively suppressed, and the device
operated as a soliton filter.

The underlying priciples, advantages and disadvantages of the NOLM device


have been presented in both the normal and anomalously dispersive configuration. The
pulse shaping property in the normal GVD regime has been shown to produce
significant pulse shaping, and has been exploited for suppressing pedestals and inter-
pulse radiation in noisy pulses from a detuned CCL laser. The soliton filtering
property of the device has also been shown to make it an ideal element for use as a
soliton discriminator in a soliton based communications system. Since the device
operates in reflection (transmission) in a similar manner to a (inverse) saturable
absorber, it may be used to provide passive mode locking in a fibre laser. Straining the
polarisation of the fibre in the loop of the NOLM can, in fact, induce a phase shift so
that it operates in either mode, both in transmission or reflection The incorporation of
gain in the loop requires investigation, and may yield a source of pulses in a feedback
arrangement.

7.4 References

(1) K.Otsuka. Opt.Lett. 8,471(1983)


U U

(2) H.Kawaguchi. Opt.Lett. 10,411(1985) U U

(3) N.J.Doran & D.Wood. J.Opt.Soc.Am. B4,1843(1987) U U

(4) N.J.Doran, K.J.Blow & D.Wood. SPIE. V836,238(1987) U U

(5) A.E.Siegman. IEEE J.Quant.Elec. QE-9, 247(1973) U U

(6) H.C.Lefevre. Elec.Lett. 16,778(1980)


U U

194
195
8. CONCLUSIONS

Throughout this thesis, various aspects of soliton generation and amplification


have been considered. The main theme of the investigation has been that of
amplification. This is an essential ingredient in any fibre based transmission system,
operating in the linear or nonlinear intensity regime. However, the ultimate
transmission speed of a system is limited by the decrease in useful bandwidth
occurring through the presence of nonlinearity. By operating with soliton signals, the
nonlinearity can be exploited to produce a larger overall system bandwidth.

Two methods of all optical amplification have been considered: stimulated


Raman amplification, and (Erbium and Neodymium) doped fibre amplification.
Raman scattering has been shown to be of vital importance in any transmission system
operating in the nonlinear regime. It provides a method for producing tunable pulse
sources and for regenerating pulses which have become attenuated below the
fundamental soliton power. However the presence of Raman gain can result in the
growth of soliton pulses of considerable intensity from noise, or lead to cross-talk
between channels in a multichannel system, both of which may lead to system errors.
The soliton self-frequency shift also imposes a limit on the shortest pulses which can
be used in a transmission system. Since this work began, the application of Raman
amplification in long distance transmission systems has been superseded
predominantly by the use of Erbium doped fibre amplification, which shows
substantially higher gain per pump photon, and improved overall performance when
employed as a lumped amplifier. The bandwidth of both gain media has been
demonstrated to be capable of amplifying pulses of less than 200fs duration. However
in practice to avoid distortion due to Raman scattering, modulation rates are generally
restricted to <10GHz per channel (τ>10ps) over transatlantic distances. To achieve
higher transmission capacities over long distances, signals are expected to be
multiplexed, however higher transmission rates should be possible over short
distances, using shorter signal pulses and higher modulation rates.

196
Amplification of spontaneous noise or seed signals in the presence of
anomalous dispersion can also give rise to modulational instability. Pulse trains
generated through modulational instability can have ultra-high repetition rates of up to
~5GHz. Hence the process is extremely useful for generating signals and clock pulses
for use in a time division multiplexed soliton transmission system. Both the
modulation and routing operations may be performed through nonlinear coupling such
as cross phase modulation, which has been shown to induce modulational instability
and induce pulse generation in the presence of anomalous dispersion. Anomalous
dispersion, gain and nonlinearity could be integrated into an Erbium laser, to produce
a CW train of control or signal pulses through intracavity modulational instability. The
feasibility of such a system has been demonstrated in the work presented here.

Ideally, soliton signals will be derived from semiconductor sources or fibre


lasers, providing direct compatibility with the transmission medium. Raman
amplification of semiconductor laser diode signals has been demonstrated, giving rise
to the generation of a train of fundamental solitons at a predetermined wavelength.
Erbium amplification should have a similar effect.

The use of high gain in a transmission system may lead to the evolution of
solitons from noise, which may be suppressed by the insertion of a natural 'noise eater
' or soliton filter based on the nonlinear loop mirror. Both gain and noise suppression
may be included within the device, which has been shown to allow transmission of
soliton pulses whilst reflecting non-soliton components, and clean up pedestals on
non-soliton pulses, suppressing inter-pulse radiation.

Amplification in the highly gain regime can result in stable regeneration of


sub-fundamental power pulses into solitons. Amplification of picosecond dispersive
pulses into solitons has been demonstrated for the first time, with no evidence of
soliton self-frequency shift for the power, pulse duration and fibre lengths used.
Ultrashort pulses at ultra-high repetition rates may yield extremely high bandwidth

197
transmission over short distances over which collisions do not occur. This would find
application for local area network communication rather than transatlantic systems. At
ultra-high repetition rates, the most important limiting factor, apart from high order
nonlinear effects, would be gain saturation.

Up to now using solitons, high repetition rates of up to ~5 Gbits/s per channel


over 10,000km have been achieved with a 75km loop, 50ps signal pulses, and
relatively short (25km) repeater spacing. Also, up to ~10 Gbits/s has been achieved
single pass over 300km with ~20ps pulses and a 50km repeater spacing2,3. The
possibility of wavelength division multiplexing up to 5 similar closely spaced (~1nm)
channels has also been theoretically proposed4, demonstrating that soliton systems will
outperform linear systems in long haul and local area network applications, with
perhaps up to 50Gbits/s total bidirectional capacity. Both systems are based on very
short repeater spacing compared with the soliton period, which ensures that
perturbations over the soliton period average out to a constant level. In this way ,the
propagating pulse, despite large fluctuations in energy in each amplifier, has the
average characteristic of an n=1 soliton over the total transmission line.

References

(1) L.F.Mollenauer, M.J.Neubelt, S.G.Evangelides, J.P.Gordon, J.R.Simpson


& L.G.Cohen. CLEO CPDP17, Anaheim, CA, USA. May 1990
(2) M.Nakazawa, K.Suzuki, H.Hubota, E.Yamada, & Y.Kimura. Submitted to
IEEE. J.Quant. Elec. (1990)
(3) M.Nakazawa. Ultrafast Phenomena 232, Monterey, CA, USA. May 1990
(4) L.F.Mollenauer & S.G.Evangelides. CLEO CFI1, Anaheim, CA, USA.
May 1990

198
199
APPENDIX ULRASHORT PULSE MEASUREMENT:

AUTOCORRELATION.

Throughout this thesis, experimental measurements have been made to deduce


the durations of pulses emerging from a variety of lasers and fibres. Typical durations
have varied from 700ps to as short as ~100fs. Electro-optic streak cameras operating
in synchroscan mode have an extremely large dynamic range, and currently yield a
temporal resolution as short as ~ 600fs single shot (Hamamatsu FESCA). The streak
camera operating principle is well known1, and an optical sampling oscilloscope based
on this technology (Hamamatsu OOS-1/IR) with ~15ps temporal resolution was used
wherever possible for real time optimisation of laser pulses.

To examine pulses with temporal resolution below 100fs, second order


autocorrelation was used. The principle of Second Harmonic Generation (SHG)
autocorrelation2,3, is based on the enhancement of the SHG field produced when two
coherent signals superpose in the SHG crystal. By imposing a varying temporal delay
between two identical pulses derived from the same input pulse, and combining them
in a SHG crystal, the resultant SHG intensity was measured as a function of delay. The
pulse duration was deduced from the delay length over which SHG enhancement
occurred. Fig.A-1 shows the schematic of both a collinear (bottom) and a non-
collinear (top) autocorrelator of the type used here, based on a Michelson
interferometer.

The input signal was divided as close to 50/50 as possible with beamsplitter
(BS) to maximise the transmission through the interferometer. The pulses in one arm
were then delayed by adjusting the axial position of mirror (M2), and both signals
were then combined in the SHG crystal (LiIO3) via filter (F1) and lens (L). Filter (F1)

prevented stray ambient light around the second harmonic frequency (2ω) from
entering the detector.

200
Retro

BS
M SHG
L
ω


PMT
F2 F1
XTAL
M2

Background -free
(Non-collinear) Delay

IN

M1

BS
SHG
2ω+ω ω
PMT
F2 F1
XTAL M2

With Background
(Collinear)
Delay

IN

Fig.A-1. Schematic of (top) non-collinear (background free) and (bottom) collinear


(with background) autocorrelators used throughout this thesis.

Two configurations were used: that of collinear interference of the beams,


which overlap throughout the system, and the non-collinear geometry, in which the
beams take separate paths and only interfere at the crystal centre. In both cases, the
SHG signal (2ω) was separated from the fundamental frequency (ω) by filter (F2),

before being detected by a temporally integrating photomultiplier tube (PMT). In the


non-collinear case, a mask was placed between the SHG crystal and the PMT tube, to
ensure that the only SHG signal detected was that generated when both pulses
overlapped in the crystal. The process of collecting SHG intensity data versus delay
was automated by mounting mirror (M2) on a mechanical shaker, yielding a sinusoidal

variation in delay. The PMT current was then displayed on the Y-channel of an

201
oscilloscope whilst a trigger signal in phase with the shaker oscillation was applied to
the timebase. The use of a digital storage oscilloscope allowed averaging to be made
over many integrations, greatly improving the signal to noise ratio in a number of
cases. In this way the real time autocorrelation was measured, facilitating optimisation
of the measured signal.

The principle behind SHG autocorrelation can be illustrated simply. If an


optical pulse with slowly varying envelope E1(t) and optical field e1(t) such that

e1 t = Re E1 t e iωt
(A-1)

interacts with a nonlinear crystal capable of being phase matched for SHG, it
generates a pulse at the second harmonic frequency 2ω with envelope E2(t). The
optical field of the second harmonic pulse E2(t) can be shown to be of the form:

e2 t = Re E2 t e 2iωt (A-2)
α Re E 2 t e2 iωt
1 (A-3)

If the fundamental pulse is split into two and recombined coherently within the
SHG crystal with a relative delay τ imposed between the pulses, the total field for
equal intensity pulses incident on the crystal at frequency ω is:

202
etot t = Re E1 t eiωt + E 1 t-τ eiω t-τ (A-4)
= Re E1 t + E 1 t-τ e-iωτ eiωt (A-5)

Now, if

E t = E1 t + E1 t-τ e -i ωτ (A-6)

then Eq(A-5) becomes:

etot t = Re E t eiωt (A-7)

The second harmonic field emerging from the crystal is proportional to the
square of the complex amplitude of the fundamental, as shown in Eq(A-3). The second
harmonic envelope function E2(t) is therefore given by:

2 2
E2 t α Et = E1 t + E1 t-τ e -iωτ (A-8)
= E21 t + E21 t-τ e-2i ωτ + 2E1 t E1 t-τ e-i ωτ
(A-9)

If only the SHG signal at 2ω is allowed to pass into the PMT (using filter F2)

the photocurrent generated i(t) is proportional to the incident second harmonic


intensity:

203
2 2
it α E2 t E*2 t = E1 t E*1 t + E1 t-τ E*1 t-τ
+ 4E1 t E*1 t E1 t-τ E*1 t-τ + S τ (A-10)

The term S(τ) contains harmonic functions in ωτ and 2ωτ, which tend to
average to zero by integration in the PMT as the delay is scanned, unless an extremely
slow trace is taken, resulting in an interferometric autocorrelation, which resolves the
interference fringes between the beams in a collinear geometry. Since the PMT cannot
resolve oscillations on a timescale of t, and averages these out to a DC level, terms in τ
can be collected in terms of the intensity I, given by EE*, where E is the field:

iτ α I2 t + I2 t-τ + 4 I t I t-τ (A-11)

where I replaces the complex field amplitude multiplied by its complex conjugate.

After normalisation, the PMT current can be shown to consist of a DC bias,


plus a function with a dependence on τ:

iτ α 1+2G 2 τ (A-12)
where G(2)(τ) is given by:

I t I t-τ 2 2
G2 τ = = E1 t E1 t dt
I2 t -∞
(A-13)

G(2)(τ) represents the well known second order autocorrelation function, and is
non-zero when the pulse overlaps with itself, i.e. a photocurrent is enhanced when the
delay is adjusted such that the two pulses overlap temporally in the SHG crystal. The
pulse duration is related to the magnitude of the delay range Δτ over which an
enhancement in the SHG photocurrent occurs. Δτ was measured throughout the
experiments presented in this thesis by moving mirror (M2) with a motorised

204
translation stage, or with a shaker, and measuring the change in the photocurrent i(τ)
versus τ. For perfect relative overlap, i.e. τ=0, a coherent pulse of duration τ0 will
produce a PMT photocurrent i(0)=3 units, and for τ>>τ0, will produce a current
i(τ>>τ0)=1, since G(2)(τ=0)=1 and G(2)(τ>>τ0)=0 in Eq(A-12). This defines the

autocorrelation intensity versus delay for a collinear autocorrelation (Fig.A-1, bottom),


and produces a trace with a constant background level. If the trace is taken slowly
enough, S(τ) does not average to zero, and fringes occur within the coherence time of
the overlapping pulses. This results in an enhancement of the contrast from 3:1 to 8:1.
Since the coherence length of a Fourier transform limited pulse is equal to the pulse
envelope duration, interferometric autocorrelation can also indicate the presence of a
chirp, with a chirped pulse resulting in a narrower interferometric enhancement of the
contrast4.

If a non-collinear geometry is used (Fig.A-1,top), as was most common


throughout this thesis, it can be shown that the first terms in Eq(A-10) are zero, and
the interference terms in S(τ) disappear.The photocurrent from the PMT is then
simply:

iτ α 2G 2 τ
(A-14)

which is only proportional to the second order autocorrelation function, without DC


bias background. This function is more sensitive to pulse shape, and is more useful for
measurement of pulse/pedestal ratio or detection of satellite pulses.

For a linear delay function Δτ , the collinear and non-collinear autocorrelation


functions for a variety of signal pulses are shown in Fig.A-2.

205
Continuous Noise Noise Burst Coherent Pulse

Background free
(Noncollinear)
2
1 1
Δυ Δυ

Second Harmonic Intensity (a.u.)


1 Δτ
Δτ
0
With Background

3
1
(Collinear)

1
Δυ Δυ
2 Δτ
Δτ
1

0
Autocorrelation Delay (arbitrary units)

Fig.A-2. Second Harmonic Autocorrelation Functions for a variety of input signals.

The first trace (left) is that of a continuous noise signal, with signal bandwidth
Δν, resulting in the generation of a coherence spike of duration Δν-1 on an infinitely
long background level. A semi-infinite noise burst gives a trace similar to that shown
in the middle. Note that the ratio of spike, pedestal and background intensities give a
good indication as to the nature of the signal when both collinear and non-collinear
autocorrelation are used. A fully coherent pulse yields a trace similar to that shown on
the right. The non-collinear geometry has no background level in this case, and is
hence a sensitive measure of proper mode locking in a laser, since inter-pulse radiation
is easily detected.

The autocorrelation function is insensitive to specific pulse shape, since it


represents the convolution of the pulse envelope function with itself. The measured
autocorrelation pulse duration Δτ must hence be deconvolved according to the
assumed pulse shape present. Pulse shapes may be inferred by fitting, thus yielding the
pulsewidth once a simple factor is taken into account. Generally, the measured
convolved pulsewidth is longer than that of the actual pulse.

206
Table A-1. Deconvolution factors Δτ/ΔT and time bandwidth products ΔΤΔν for a
variety of common pulse shapes. (Top to bottom- Square, Gaussian, Sech2 and single-
sided exponential.)

The the factor Δτ/ΔΤ for determining the pulse FWHM duration ΔΤ from the
autocorrelation FWHM width Δτ is given in Table A-1 for a variety of commonly
encountered pulse shapes. The time bandwidth product ΔΤΔν of a Fourier transform
limited pulse of duration ΔΤ and bandwidth Δν is also shown for reference. The

I(t) Δ τ/Δ T ΔTΔ ν maximum


resolution of
1 (0≤ t≤Δ T) 1 .886 the
autocorrelator
-(4ln2) t2 .441
exp 2
ΔT used

2 1.76 t 1.55 .315


sech ΔT

-(ln2) t
exp 2 .110 throughout this
ΔT (t>0)
thesis was
directly shown to be ~<18fs5, limited by the bandpass of the SHG crystal, and any
spectrally narrowing or dispersive elements placed before it.

References.

(1) D.J.Bradley & G.H.C.New. Proc. IEEE. 3,314(1974)


U U

(2) J.A.Armstrong. Appl.Phys.Lett. 10,16(1967)


U U

(3) K.L.Sala, G.A.Kenney-Wallace, G.E.Hall.


IEEE J.Quant.Elec. QE-16, 990(1980)
U U

(4) J.C.Diels, J.J.Fontaine, I.McMichael & F.Simoni. Appl. Opt. 24,1270(1985)


U U

(5) A.S.Gouveia-Neto, A.S.L.Gomes & J.R.Taylor. J.Mod.Opt. 35,7(1988).


U U

207
JOURNAL PUBLICATIONS.

(1) Subpicosecond Pulse Generation Through Cross-Phase Modulation Induced


Modulational Instability In Optical Fibres. A.S.Gouveia-Neto, M.E.Faldon,
A.S.B.Sombra, P.G.J.Wigley and J.R.Taylor. Opt Lett 13 (1988) U U

(2) A CW Neodymium Doped Fibre Ring Oscillator. P.G.J.Wigley,


A.S.B.Sombra and A.S.Gouveia-Neto. Revista de Fisica
Aplicada e Instrumentacao 3(1988) U U

(3) Amplification Of Picosecond Pulses in Nd-Doped Single Mode Optical Fibre.


A.S.Gouveia-Neto, A.S.B.Sombra, P.G.J.Wigley and J.R.Taylor.
J of Mod Opt. 36(1989)
U U

(4) A Synchronously Pumped Dispersion Compensated Fibre Raman Ring Laser


Around 1.4µm. A.S.Gouveia-Neto, P.G.J.Wigley and J.R.Taylor.
Opt. Comm. 70(1989)
U U

(5) Soliton Generation Through Raman Amplification Of Pulses With


Sub-Fundamental Soliton Powers. A.S.Gouveia-Neto, P.G.J.Wigley
and J.R.Taylor. Opt. Comm. 72(1989) U U

(6) Soliton Generation Through Raman Amplification Of Noise Bursts.


A.S.Gouveia-Neto, P.G.J.Wigley and J.R.Taylor. Opt. Lett. 14(1989) U U

(7) Generation Of 2THz Repetition Rate Pulse Trains Through Induced


Modulational Instability. E.J.Greer, D.M.Patrick, P.G.J. Wigley
and J.R.Taylor. Elect. Lett. 25(1989) U U

(8) Tunable Femtosecond Soliton Generation From Amplified CW Diode Laser


Signals. E.J.Greer, D.M.Patrick, P.G.J.Wigley, J.I.Vukusic and
J.R.Taylor. Opt. Lett. 15(1989)
U U

(9) Femtosecond Soliton Amplification In Erbium Doped Silica Fibre.


B.J.Ainslie, K.J.Blow, A.S.Gouveia-Neto, P.G.J.Wigley, A.S.B.Sombra
and J.R.Taylor. Elec. Lett. 26(1990)U U

(10) Tunable CW Fibre Raman Ring Laser Centred At 1.42µm. E.J.Greer,


P.G.J.Wigley, and J.R.Taylor. Electron. Lett.26(1990) U U

(11) Picosecond Pulse Generation From A Continuous Wave Diode Laser Through

208
Cross Phase Modulation In An Optical Fibre. E.J.Greer, D.M.Patrick,
P.G.J.Wigley and J.R.Taylor. Opt. Lett. 15(1990)
U U

(12) Pulse Shaping, Compression and Pedestal Suppression Employing A


Nonlinear Optical Loop Mirror. K.Smith, N.J.Doran and P.G.J.Wigley.
Accepted for publication in Opt. Lett.(1990)

(13) Mode Locking of a Continuous Wave Neodymium Doped Fibre Laser with a
Linear External Cavity. P.G.J.Wigley, P.M.W.French and J.R.Taylor.
Elec. Lett. 26(1990)
U U

(14) Active Mode Locking of an Erbium Doped Fibre Laser using an Intra-Cavity
Diode Laser Device. P.G.J.Wigley, A.V.Babushkin, J.I.Vukusic
and J.R.Taylor. Phot. Techn. Lett. 2(1990)
U U

CONFERENCE PUBLICATIONS.

(1) Modulational Instability Induced Through Cross-Phase Modulation From


Pulses In The Normal Dispersion Regime. A.S.Gouveia-Neto, M.E.Faldon,
A.S.B.Sombra, P.G.J.Wigley and J.R.Taylor. Technical Digest Of IQEC'88,
Tokyo, Japan,ThE2, 664(1988).
U U

(2) Picosecond Pulse Amplification In A Single Mode Neodymium Doped Fibre.


A.S.Gouveia-Neto, A.S.B.Sombra, P.G.J.Wigley and J.RTaylor.
Proc. Of EQEC'88, Hanover, FRG (1988).

(3) Soliton Generation Through Raman Amplification Of Modulational Instability


And Fundamental Soliton Power Pulses. A.S.Gouveia-Neto, D.M.Patrick,
P.G.J.Wigley and J.R.Taylor. Proc. of Topical Meeting On 'Nonlinear Guided
Wave Phenomena: Physics And Applications'. Houston, Texas, USA (1989)

(4) Soliton Generation Through Gain Modulation. E.J.Greer, D.M.Patrick,


P.G.J.Wigley and J.R.Taylor. Proc. of 'Applications Of Ultrashort Pulses
For Optoelectronics'. London, UK (1989)

(5) Soliton Generation Through Raman Amplification Of Diode Laser Signals.


E.J.Greer, D.M.Patrick, P.G.J.Wigley and J.R.Taylor. Proc. of
'Ultrafast Phenomena In Spectroscopy'. Neubrandenburg, E. Germany (1989)

209
(6) Optical Solitons. A.S.Gouveia-Neto, E.J.Greer, D.M.Patrick, P.G.J.Wigley
and J.R.Taylor. Proc. of the 2nd European Quantum Electronics Conference.
Dresden, E. Germany (1989)

(7) The Dispersion Compensated Fibre Raman Ring Laser - Applications To


Soliton
Noise Investigations. P.G.J.Wigley. Technical Digest of the 'Ninth National
Quantum Electronics Conference', Oxford, UK (1989)

(8) Raman Solitons. P.G.J.Wigley. Meeting of the IOP Quantum Electronics


Group on 'Femtosecond Optics', London, UK (1989)

(9) Tunable Femtosecond Soliton Generation From Continuous Diode Laser


Signals
Using Pulsed Stimulated Raman Amplification. E.J.Greer, D.M.Patrick,
J.I.Vukusic, P.G.J.Wigley, and J.R.Taylor. Technical Digest of
'Integrated Photonics Research' , Hilton Head, South Carolina, USA(1990)

(10) Picosecond Pulse Generation From A Continuous Wave Diode Laser Using
Cross Phase Modulation. E.J.Greer, D.M.Patrick, P.G.J.Wigley and
J.R.Taylor. Technical Digest of CLEO'90, Anaheim, California ,USA (1990)

(11) Mode Locking Of Solid State Lasers Using A Linear External Cavity.
P.M.W.French, S.M.J.Kelly, P.G.J.Wigley and J.R.Taylor.
Technical Digest of CLEO'90, Anaheim, California, USA (1990)

(12) Pulse Shaping, Compression, And Pedestal Suppression Employing A


Nonlinear Optical Loop Mirror. K.Smith, N.J.Doran and P.G.J.Wigley.
Technical Digest of CLEO'90, Anaheim, California, USA (1990)

(13) Continuous Wave Diode Laser Seeding Of Tunable Femtosecond Soliton


Plulses. E.J.Greer, D.M.Patrick, J.I.Vukusic, P.G.J.Wigley, and
J.R.Taylor. Technical Digest of Topical Meeting on Ultrafast Phenomena V11,
Monterey, California, USA. Springer (1990)

210
ACKNOWLEDGEMENTS

I would like to acknowledge the excellent supervision and guidance of Dr. Roy
Taylor throughout my Doctoral research, and for the critical reading of this
manuscript. For their encouragement and advice, I would like to thank Drs.Paul
French, and John Vukusic. I would also like to thank Dr.Stephen Kelly for many
useful discussions.

I would also like to thank Drs.Nick Doran, Keith Blow and Kevin Smith for
their assistance, and for allowing me to collaborate in their research at BTRL.

For their help in the lab, I would like to thank Dr. Arthur Gouveia-Neto, Elaine
Greer, David Patrick, and Nilay Pandit. I would also like to thank Elaine, Nilay and
Glenn Atkins for correcting my spelling.

The expert technical assistance of John Bean, Bob Wilkins, Dave Mountain,
Ted Bates, Charlie Cooper and Roy Morrison is also gratefully acknowledged.

The financial support for the work presented here was provided jointly by the
Science and Engineering Research Council and British Telecom Research
Laboratories.

211

S-ar putea să vă placă și