+ All Categories
Home > Documents > spectral hole-burning and slow light in emerald and ruby

spectral hole-burning and slow light in emerald and ruby

Date post: 11-Sep-2021
Category:
Upload: others
View: 2 times
Download: 0 times
Share this document with a friend
122
SPECTRAL HOLE-BURNING AND SLOW LIGHT IN EMERALD AND RUBY IVANA CARCELLER A THESIS SUBMITTED IN FULFILMENT OF THE REQUIREMENTS FOR THE DEGREE OF MASTERS BY RESEARCH School of Physical, Environmental and Mathematical Sciences University of New South Wales Canberra March 2012
Transcript
Page 1: spectral hole-burning and slow light in emerald and ruby

SPECTRAL HOLE-BURNING AND SLOW

LIGHT IN EMERALD AND RUBY

IVANA CARCELLER

A THESIS SUBMITTED IN FULFILMENT OF THE

REQUIREMENTS FOR THE DEGREE OF MASTERS BY

RESEARCH

School of Physical, Environmental and Mathematical Sciences

University of New South Wales Canberra

March 2012

Page 2: spectral hole-burning and slow light in emerald and ruby

COPYRIGHT STATEMENT

‘I hereby grant the University of New South Wales or its agents the right to archive and to make available my thesis or dissertation in whole or part in the University libraries in all forms of media, now or here after known, subject to the provisions of the Copyright Act 1968. I retain all proprietary rights, such as patent rights. I also retain the right to use in future works (such as articles or books) all or part of this thesis or dissertation. I also authorise University Microfilms to use the 350 word abstract of my thesis in Dissertation Abstract International (this is applicable to doctoral theses only). I have either used no substantial portions of copyright material in my thesis or I have obtained permission to use copyright material; where permission has not been granted I have applied/will apply for a partial restriction of the digital copy of my thesis or dissertation.' Signed ……………………………………………........................... Date 31-03-2012

AUTHENTICITY STATEMENT

‘I certify that the Library deposit digital copy is a direct equivalent of the final officially approved version of my thesis. No emendation of content has occurred and if there are any minor variations in formatting, they are the result of the conversion to digital format.’ Signed ……………………………………………........................... Date 31-03-2012

Page 3: spectral hole-burning and slow light in emerald and ruby

Originality Statement

I hereby declare that this submission is my own work and to the best of my

knowledge it contains no materials previously published or written by another

person, or substantial proportions of material which have been accepted for the

award of any other degree or diploma at UNSW or any other educational

institution, except where due acknowledgement is made in this thesis. Any

contribution made to the research by others, with whom I have worked at

UNSW or elsewhere, is explicitly acknowledged in the thesis. I also declare that

the intellectual content of the thesis is the product of my own work, except to the

extent that assistance from others in the project's design and conception or in

style, presentation and linguistic expression is acknowledged.

Ivana Carceller

Signed ..........................

Date 31/03/2012

Page 4: spectral hole-burning and slow light in emerald and ruby

Acknowledgements

Firstly, I would like to thank my supervisor Prof. Hans Riesen for his guidance,

patience, support and kindness in the development of this thesis. Without his

knowledge, encouragement and willingness, the completion of this thesis would

have been impossible. I am very grateful to him for his generous help,

tolerance, sense of humour and respect even during the hard times.

I also want to thank Dr. Wayne Hutchison for his proofreading despite the short

notice we gave him.

I want to express my gratitude to the University of New South Wales Canberra

for giving me the opportunity and offering me a Research Training Scholarship.

I wish to thank the other students and staff at the Australian Defence Force

Academy for all the good times and company in the last two years.

I would like to thank my Mum, Dad and the rest of family because without their

love and support, I would not have finished this thesis.

Special thanks go to my beloved friends Renata for her unconditional support

during hard times in Canberra, her wise advices and maturity and to Karina for

her sense of humour and for making me laugh even in the worst moments.

I would like to thank Greg for his patience, invaluable support and

understanding throughout the final stage of this project.

I want to mention my good friends Sarah, Lucia, Bea, Maria, Miriam and Alex

for their invaluable friendship throughout so many years regardless the

distance. I want to thank David and Tom for their help upon my settling into

Canberra. Also thanks to Federica, Bridget, Rhys, Michael and Rob for her

friendship and encouragement.

Also I would like to thank all the staff and students in the School of Physical,

Environmental and Mathematical Sciences at the University of New South

Page 5: spectral hole-burning and slow light in emerald and ruby

ii

Wales. In particular, I want to thank Emeritus Prof. Brian Lees, Prof. Warrick

Lawson, Ms Tessa Hodson, Dr Steve James, Dr Barry Gary, Mrs Annabelle

Boag and Mrs Nadia Seselja. My sincere thanks are also due to Ms Julie Kesby

who helped me when I was sick.

Also, thanks to Tracy Massil for her wise advices in my first months in Canberra

and her funny stories. I would like to thank Baran and Xianglei for their

proofreading of my thesis.

Page 6: spectral hole-burning and slow light in emerald and ruby

iii

Abstract

The temperature dependence of the R1 line of Chatham laboratory grown

emeralds (Be3Al2(SiO3)6) containing low concentrations of chromium (III) (0.04%

and 0.0017% per weight) is investigated in the temperature range of 2.5 - 260 K

by spectral hole-burning, luminescence line narrowing experiments and

luminescence spectroscopy. The contribution to the homogeneous linewidth at

6 K is 70 kHz and over 400 GHz at 260 K. The data is well described by the

two-phonon Raman process above 80 K. Below 80 K, the direct process

between the two 2E levels is dominant. However, at temperatures below 10 K a

low energy phonon is required to explain the temperature dependence.

The generation of slow light by transient hole-burning is reported for the first

time. A Gaussian probe pulse is propagated in an optically dense medium

provided by the R1(±3/2) line of 130 ppm Bagdasarov ruby. A delay of 10.8 ns is

observed, corresponding to a reduction of the group velocity of 213000 m/s

which is a reduction by a factor of 1400 compared to air.

Page 7: spectral hole-burning and slow light in emerald and ruby

iv

List of Publications

Publications that will result from this thesis:

1. Ivana Carceller, Wayne D. Hutchison, Hans Riesen; Temperature

dependence of the R1 linewidth in emerald; manuscript in preparation for

publication.

2. Hans Riesen, Aleksander Rebane, Ivana Carceller, Alex Szabo; Slow light in

pink ruby by transient spectral hole-burning; manuscript in preparation for

publication.

Page 8: spectral hole-burning and slow light in emerald and ruby

v

Table of contents

Chapter 1 Introduction.........................................................................................1

1.1. Theoretical background.....................................................................3

1.1.1. Electronic transition.............................................................3

1.1.2. Homogeneous linewidth......................................................4

1.1.3. Inhomogeneous broadening................................................8

1.2. Chromium (III) systems....................................................................10

1.2.1. General properties..............................................................10

1.3. Temperature dependence of the homogeneous linewidth...............14

1.4. Laser spectroscopy..........................................................................19

1.4.1. Spectral hole-burning (SHB)..............................................19

1.4.1.1. transient spectral hole-burning.............................20

1.4.1.2. Persistent hole-burning.........................................21

1.4.1.3. Potential applications of SHB...............................22

1.4.2. Fluorescence line narrowing (FLN)....................................23

1.5. Slow light..........................................................................................24

1.5.1. Kramers-Kronig relations....................................................28

1.5.2. Methods to create slow light...............................................29

1.5.2.1. Electromagnetically induced transparency (EIT)..29

1.5.2.2. Coherent population oscillation (CPO).................32

Page 9: spectral hole-burning and slow light in emerald and ruby

vi

1.5.2.3. Hole-burning .......................................................36

1.6. References......................................................................................38

Chapter 2 Experimental....................................................................................44

2.1. Sample preparation.........................................................................44

2.2. Room temperature spectroscopy....................................................44

2.3. Spectroscopy at Liquid temperatures..............................................44

2.3.1. Mounting of samples.........................................................45

2.3.2. Cooling method and temperature control..........................46

2.3.3. Absorption and Transmission spectroscopy......................46

2.3.4. Non-selective luminescence spectroscopy........................47

2.3.5. Spectral hole-burning.........................................................48

2.3.6. Fluorescence line narrowing..............................................49

2.3.7. Slow light experiments.......................................................50

2.4. Application of an external magnetic field.........................................57

2.5. Laser sources..................................................................................58

2.5.1. Nd:YAG laser ....................................................................59

2.5.2. He-Ne laser........................................................................60

2.5.3. Diode laser.........................................................................61

2.5.4. External cavity diode laser.................................................63

2.5.4.1. Frequency lock.....................................................64

2.6. Monochromator................................................................................65

2.7. Fabry-Pérot interferometer...............................................................67

Page 10: spectral hole-burning and slow light in emerald and ruby

vii

2.8. The Laue method.............................................................................69

2.9. Data acquisition and analysis...........................................................70

2.9.1. Data acquisition..................................................................70

2.9.2. Data analysis......................................................................71

2.10. References.....................................................................................73

Chapter 3...........................................................................................................75

3.1. Emerald............................................................................................75

3.1.1. Crystal structure and background......................................75

3.1.2.Results and discussion........................................................77

3.2. Slow light in ruby by transient spectral hole-burning........................90

3.2.1. Crystal structure and orientation........................................90

3.2.2. Results and discussion.......................................................92

3.2.2.1. Experiment...........................................................92

3.2.2.2. Simulations..........................................................97

3.3. References.....................................................................................100

Chapter 4................................................................................................. ........102

Appendix 1.......................................................................................................103

Appendix 2.......................................................................................................104

Appendix 3.......................................................................................................105

Appendix 4................................................................................................. ......107

Appendix 5.......................................................................................................109

Page 11: spectral hole-burning and slow light in emerald and ruby

Chapter 1

Introduction

The optical spectroscopy of transition metal ions in the solid state

has been of interest for decades. Electronic transitions in the solid

state are subject to inhomogeneous broadening as a consequence

of the presence of impurities or imperfections in the lattice. Many

laser-based techniques have been developed to overcome this

broadening; these techniques include hole-burning, fluorescence

line narrowing and photon echo measurements which can give us

information at very high resolution enabling a better understanding

of the most fundamental properties of electronic structure.

Therefore, understanding the dynamics and structure of crystalline

materials (as well as amorphous solids) is an important aim and a

significant effort has been made over the years in order to achieve

this goal.

Many potential applications of laser-based techniques are also

receiving significant attention such as frequency and time domain

optical data storage which are based on spectral hole-burning. This

implies that in the future new technologies may be applied for

advancement in the areas of data storage and quantum computing.

Parallel to this, researchers have made a great effort in order to slow

down the speed of light. In the telecommunications industry,

Page 12: spectral hole-burning and slow light in emerald and ruby

enormous amounts of data are transmitted through optical fibres at

very high transfer rates. But, unfortunately, information cannot be

processed at those rates. Due to limiting factors of existing

technology, light signals cannot be stored or processed without

being transformed into electrical signals, which only work on a much

slower timescale. The idea would be to control the light signals by

light in order to avoid the necessity to convert them into electrical

signals.

Page 13: spectral hole-burning and slow light in emerald and ruby

3

1.1. Theoretical background

Some theoretical concepts are presented in this section which are used

throughout this thesis in order to provide the reader with a better understanding

of the work undertaken. A rigorous quantum treatment of electronic excitations

and their linewidths can be found in [1, 2].

1.1.1. Electronic transition

The simplest electronic transition takes place in a two-level system where an

electron jumps from the ground state a to the excited state b (see Figure 1.1).

When this transition occurs, the electron gains energy which is subsequently

lost through radiative or non-radiative processes when it relaxes back to the

ground state.

In the radiative relaxation process the system is deactivated to the ground state

by emitting a photon. The emitted photon has an energy equivalent to the

energy difference between the ground and the excited state. On the other hand,

in non-radiative relaxation the electronic excitation energy is dissipated into

phonons (vibrations) to the rest of the crystal lattice.

The three basic processes of absorption, spontaneous and stimulated emission

are not independent of each other in a system. At equilibrium with an external

electromagnetic radiation field the overall transition rate from level a to level b

must be the same as that from level b to level a [3, 4].

Page 14: spectral hole-burning and slow light in emerald and ruby

4

Figure 1.1. The three possible transitions between a ground and an excited state.

Two parameters determine the width of an electronic transition: T1, which is the

lifetime of the excited state and T2*, which is the pure dephasing time [1].

1.1.2. Homogeneous linewidth

A significant effort has been made in order to rigorously understand electronic

structures in crystalline and amorphous solids and many sophisticated

experimental techniques have been applied so far to achieve this goal.

As even the most perfect crystals contain either imperfections or impurities, an

important experimental technique can be the optical spectroscopy of these

impurities (or optical centres) in the host structure. Hence, optical spectroscopy

of impurities embedded in a host matrix is an important tool to unravel its

electronic structure and associated dynamics [5].

An electronic transition between the levels of a two-level system can be

represented by a one-dimensional damped harmonic oscillator as is illustrated

in Figure 1.2.

relaxation

excitation

Incident photon

ground state a

nonradiative relaxation

OR

radiative relaxation

excited state b

emitted photon

Page 15: spectral hole-burning and slow light in emerald and ruby

5

Figure 1.2. Dampened harmonic oscillator. Adapted from [6].

If monochromatic electromagnetic radiation disturbs a two-level system, then

after a time t, an electron can be excited. A harmonic oscillation then happens

in this system and it is damped by dissipation of energy (radiative and non-

radiative relaxation processes as explained above). The response of such a

system can be represented with the following equation:

Eq. (1.1)

The damping term can be expressed as the inverse of the excited state lifetime,

that is, γ=1/T1 which is determined by the radiative and non-radiative relaxation

processes. The angular frequency is given by:

Eq. (1.2)

x(t)

x0

Time t

Page 16: spectral hole-burning and slow light in emerald and ruby

6

where k is the restoring force constant and is the mass of the electron.

The solution of the Eq. (1.1) is

Eq. (1.3)

where is the amplitude of the undamped oscillator [8].

To understand this better, let us consider a system with just one impurity at very

low temperatures. This impurity has an absorption associated with the two

electronic levels responsible for the transition. When that transition is weakly

coupled to lattice vibrations of the host, the spectrum has a Lorentzian line

shape which is the Fourier transform of Eq. (1.3). The Fourier transform of Eq.

(1.3) has a Lorentzian profile that can be expressed in terms of frequency by

[8]

Eq. (1.4)

where is the resonance frequency and is the full width at half maximum in

Hz expressed by

Eq. (1.5)

If, now, we consider many impurities distributed randomly in a perfect crystal

with equivalent local geometry and environment, the resulting spectrum will still

have a Lorentzian lineshape as in the case of just one impurity.

Page 17: spectral hole-burning and slow light in emerald and ruby

7

At room temperature, the interactions between electrons and phonons and, to a

lesser extent, electron and nuclear fluctuations are the main contributions to the

homogeneous linewidth; this means that the pure dephasing time T2*

determines the linewidth at high temperatures. In contrast, at lowest

temperatures, the excited state lifetime, T1, can become the dominant

contributor. The homogeneous spectral linewidth is given by the effective optical

dephasing time, T2 [9, 10]. The full width at half maximum of the Lorentzian line

shape can be expressed as [11, 12]

Гhom=

=

+

Eq. (1.6)

where the relationship between the excited state lifetime T1, the pure dephasing

time T2* and the effective dephasing time T2 is given by

Eq. (1.7)

Figure 1.3. Perfect crystal and the Lorentzian lineshape of identical optical centres.

w0

γ=1/T1

Angular frequency w

Page 18: spectral hole-burning and slow light in emerald and ruby

8

1.1.3. Inhomogeneous broadening

Unfortunately, all crystals contain imperfections, impurities, dislocations, isotope

distribution, etc. As a consequence every single optical centre is in a different

local environment. Because the local field of an impurity does affect its

transition frequency, each single impurity will have a slightly shifted transition

frequency with respect to a central frequency. As a result, the lineshape will

most likely be Gaussian since it is the sum of shifted individual Lorentzian line

shapes [7]. This effect is called inhomogeneous broadening and its width is

usually referred to as Гinh.

The theory of inhomogeneous broadening in crystals is founded on the

statistical model of Stoneham [13]. He assumes that in crystals some point

defects are distributed randomly and that those defects interact with the impurity

transition [13]. If we consider r as the distance between the optical centre and

the defect then the perturbation of the optical centres transition frequency varies

as 1/r3 and a Lorentzian lineshape is predicted. When the defects are more

concentrated the lineshape changes from Lorentzian to Gaussian [14, 15].

Figure 1.4. Inhomogeneous broadening. The spectrum is the sum of a large number of much narrower individual homogeneously broadened lines that are each shifted in frequency with respect to each other. Gaussian line

shapes are often observed. Adapted from [5].

Page 19: spectral hole-burning and slow light in emerald and ruby

9

At liquid helium temperatures, that is at temperatures < 5 K, only the

inhomogeneous linewidth can be observed in non-selective spectroscopy as the

phonon-induced fluctuations are frozen out and hence, are negligible. In other

words, the homogeneous linewidth cannot be observed in conventional

spectroscopy at low temperatures [5]. On the other hand, at high temperatures,

the inhomogeneous broadening becomes less important and the homogeneous

linewidth may be directly observed and the lineshape becomes Lorentzian [5].

When MacFarlane and his collaborators [16] started using single frequency dye

lasers in the 1970s, many spectral lines of rare-earth impurities were shown to

have inhomogeneous linewidths of 1-10 GHz. In the late 1980s, the

development of reliable single frequency lasers enabled a significant increase of

the number of materials and linewidths that could be studied [17, 18].

The inhomogeneous linewidth of an electronic origin is mainly determined by its

nature. So far rare-earth doped crystals have provided the narrowest

inhomogeneous linewidths. To a great extent this is because their f-f transitions

are not strongly perturbed by the surrounding atoms. Also pure spin-flip

transitions in chromium (III) are not very sensitive to inhomogeneous

broadening and present narrower inhomogeneous linewidths than most other d-

d transitions [11].

An interesting example for rare earth impurity ions is the case of Nd3+ ions at

low concentration in Y7LiF4 showing inhomogeneous broadening as narrow as

10 MHz [19].

Another example of spectral lines with very low inhomogeneous broadening,

are the transitions of Pr3+ in CaF2, where six hyperfine lines of 600 MHz were

resolved [20].

On the other side, the largest inhomogeneous linewidths observed are for

charge-transfer transitions and for ligand-field transitions. These transitions are

also affected by strong coupling to vibrations.

Page 20: spectral hole-burning and slow light in emerald and ruby

10

Table 1.1 shows a comparison of d-d and f-f transitions in crystalline and

amorphous structures is presented summarizing the very different values for the

inhomogeneous linewidths [11].

Transition Crystalline host

(cm-1)

Amorphous host

(cm-1)

d – d 0.1 - 100 10 - 1000

f - f 0.01 – 10 1 - 100

Table 1.1. Typical inhomogeneous linewidths of d-d and f-f transitions in crystalline and amorphous hosts [11].

1.2. Chromium (III) systems

The 4A2 – 2E transition in chromium (III) systems is discussed in this section. A

brief overview of the structure and bonding characteristics of these systems are

presented to allow a better understanding of the spectroscopic observations

explained later.

1.2.1. General properties

The chromium (III) cation is a 3d3 system having 3d electrons on the outside of

the ion. These outer electrons are very sensitive to the environment. The

dominant ligand field component in chromium (III) complexes is generally of

octahedral symmetry. Apart from this contribution, lower symmetry fields can

exist, such as trigonal, rhombic or tetragonal components, and they have to be

taken into account in a rigorous description [1, 21].

Page 21: spectral hole-burning and slow light in emerald and ruby

11

Figure 1.5. Octahedral geometry. This is a representation of the central chromium ion surrounded by the six ligands in an octahedral environment. Adapted from [22].

As is shown in the Figure 1.5, due to the octahedral ligand field symmetry, the

five degenerate d orbitals are divided into two sets of levels, eg (twofold

degenerate) and t2g (threefold degenerate). This splitting is due to the

interaction between electron orbitals of the metal and electron configuration of

the ligand. The dxy, dxz and dyz orbitals are stabilised and their energy is

reduced. The dz2 and dx

2-y

2 orbitals are destabilised due to the fact that these

orbitals point in the same direction as the ligands. The energy difference

between these two orbitals eg and t2g is characterized as 10Dq, where Dq is a

ligand field parameter that measures the octahedral field strength [21, 23]. The

splitting is shown in Figure 1.6.

Figure 1.6. Octahedral ligand field splitting. Both eg and t2g levels exhibit degeneracy.

Energy

dyz

dx2-y

2

dxz

dz2

dxy

10 Dq

t2g

eg

Page 22: spectral hole-burning and slow light in emerald and ruby

12

In a perfect octahedral field, numerous energy levels are present in a chromium

(III) system. Two factors are important when it comes to determine the energy

of these levels: the octahedral field strength Dq and the Racah parameter B

which is a parameter for the description of electron-electron repulsion.

In the Tanabe-Sugano [1] diagram shown in Figure 1.7, the energy levels for

3d3 systems are represented.

Figure 1.7. Tanabe-Sugano energy level diagram for 3d3 electronic configuration in an octahedral field. The

octahedral field increases from left to right. The energy of the 4A2 ground state is used as a reference to measure

the energy of each electronic state. Adapted from [1].

Emerald Ruby

Page 23: spectral hole-burning and slow light in emerald and ruby

13

In addition to the lower symmetry fields that add more complexity to the

structure of chromium (III) systems, the effects of spin-orbit coupling must be

considered. In the Figure 1.8 a diagram for the 2E-4A2 transition is shown.

Figure 1.8. Schematic energy level diagram for ruby and emerald. a) Energy level diagram of Cr3+

in a an octahedral field. b) Fine structure of

4A2 and

2E. c) Luminescence spectrum of ruby at low temperature. d) Absorption spectrum

of ruby at low temperature. Two intense and broad spin-allowed transitions are shown.

At very low temperatures only the transitions from the lowest excited states can

be observed in luminescence of chromium (III) ions. This is because higher

energy states are not populated at these temperatures as they relax non-

radiatively to the lowest excited state. The 4A2 – 2E transition is a spin-forbidden

transition (because ∆S≠0). Despite being forbidden, it still happens due to lower

symmetry fields and spin-orbit coupling. In chromium (III) systems with strong

ligands, the 2E spin-flip level is the lowest excited state. Luminescence is

possible from the two levels of 2E as seen in Figure 1.8. As a result the four R-

lines due to ground and excited state splitting can be observed [24, 25].

4A2

2E

d) Absorption c) Luminescence b) Octahedral +

trigonal ligand

field + spin-orbit

coupling

R1 R2

a) Octahedral field

No

n-r

adia

tive

rel

axat

ion

Ab

sorp

tio

ns

4T1

4T2

Emis

sio

ns

(flu

ore

sce

nce

)

Page 24: spectral hole-burning and slow light in emerald and ruby

14

Transition metal complexes are very sensitive to the metal-ligand distances,

which affect the ligand field and can cause large shifts in the electronic

potentials of the energy levels. For instance, the displacement between ground

and excited state potentials in d-d spin-allowed transitions can be very

significant and lead to very pronounced vibrational sidebands. Also the

electronic origin accounts for a fraction of the total intensity of the spectrum

only. Hence, the intensity and energy of the electronic origin are very

susceptible to changes of the metal-ligand distances for spin-allowed d-d

transitions.

When transitions occur in the same electronic configuration, they are to a great

extent independent of the ligand field. As is seen in the Tanabe-Sugano energy

diagram of Figure 1.7, the 2E energy remains almost constant when increasing

the octahedral field strength. However, the energy separation of other levels

such as 4T1 and 4T2 grow linearly as the octahedral field is increased (these last

energy levels are interconfigurational d-d spin-allowed transitions). Hence,

spin-flip transitions such as 4A2 → 2E present weak vibrational sidebands and

most intensity is in the narrow electronic origin.

Additional broadening of an optical transition is observed when other factors

affect the symmetry of the system. Spin allowed transitions usually appear

without a specific structure due to a high degree of inhomogeneous broadening.

4A2 – 2E spin-flip transitions which happen in the same electronic configuration

are less sensitive to these changes owing to their first order independence from

the ligand field strength [8, 26].

1.3. Temperature dependence of the homogeneous linewidth

The temperature dependence of the homogeneous linewidth in crystalline

structures is due to different mechanisms which are based on electron and

nuclear spin-spin interactions and electron-phonon interactions [8, 11].

Page 25: spectral hole-burning and slow light in emerald and ruby

15

Figure 1.9. Schematic representation of important mechanisms of electron-phonon interaction. From left to right: direct process, Orbach process and Raman process.

The first case is the direct process. This process occurs very rapidly when the

energy gap is close to the maximum density of phonon states. Due to the high

density of phonon states in coordination compounds, it is a very important

source of homogeneous line broadening. In coordination compounds the energy

gaps are typically between 10 and 3000 cm-1. As a consequence the direct

process is very likely to happen because of the high density of phonon states is

in the same range [1].

The Raman and Orbach processes are non-radiative transitions mechanisms

between close-lying levels through two phonons as depicted in Figure 1.9. In

the Orbach process one phonon is absorbed from the lower to the upper state

and then emitted back to another sublevel of the lower state. The Raman

process is basically the same as the Orbach mechanism but the difference is

that the upper state is virtual only. There is rapid broadening with temperature

of the homogeneous linewidth when one of these two processes takes place

according to the Debye approximation for the density of phonon states [1, 8,

25].

A highly influential paper in the field of thermally induced broadening and shift of

optical transitions of impurity centres in solids was published by McCumber and

Sturge in 1963 [27]. They calculated the contributions of one-phonon (“direct”)

a

b

c

Page 26: spectral hole-burning and slow light in emerald and ruby

16

and two-phonon (“Raman”) processes to the broadening and shift for optical

transitions for systems with level splitting of the order of kT. By means of non-

selective spectroscopy they analysed the R-lines of ruby [27]. They used a

perturbation approach in the weak coupling limit resulting in the Eq. (1.8.a):

Eq. (1.8.a)

where is the Debye temperature. The linewidth contribution expressed in Eq.

(1.8.a) is due to two phonon Raman scattering of Debye phonons and it

describes the temperature dependence of the homogeneous linewidths of

impurity centres in solids.

In the 1980s, Hsu and Skinner [28] published a series of papers using a non-

perturbative approach for the temperature dependence of the homogeneous

linewidth within the weak-coupling regime and the Debye approximation for the

density of phonon states. They obtained the following temperature dependent

contribution to the homogeneous linewidth for the Debye model of acoustic

phonons:

Eq. (1.8.b)

In this equation W is the quadratic electron-phonon coupling constant,

where is the Boltzmann constant. In the weak coupling limit (W<<1) Eq.

(1.8.b) simplifies to Eq. (1.8.a) (original equation obtained by McCumber and

Sturge) [27]. They suggested that the direct one-phonon process (a phonon is

absorbed or emitted between the 2Ā and Ē levels of the 2E state separated by

∆) is not relevant for temperatures > 90 K.

The direct one-phonon process is expressed by the well known Eq. (1.9):

Page 27: spectral hole-burning and slow light in emerald and ruby

17

Eq. (1.9)

where is the 2E splitting (29 cm

-1 in ruby) and is the R2 linewidth

at 2.5 K ( expresses the zero-point lifetime of the 2Ā level).

Blume et al. [29] calculated the relaxation lifetime of 2Ā(2E) to Ē(2E) level due to

spontaneous phonon emission in ruby and it was found to be 0.3 ns. This

allowed calculating the direct one-phonon process contribution. More recently a

study [30, 31] was undertaken to measure the temperature dependence of the

R lines in ruby. The temperature dependence of the R2 linewidth was found to

be dominated by a direct one-phonon process up to 50 K and by a two-phonon

Raman scattering process above that temperature. In Figure 1.10 a re-

examination of the R1(±3/2) linewidth in ruby by doing transient hole-burning

measurements in a low magnetic field is shown. One can observe the direct

one-phonon process between the split levels of the 2E excited state up to 50 K

and the two-phonon Raman scattering process at higher temperatures.

Page 28: spectral hole-burning and slow light in emerald and ruby

18

Figure 1.10. Temperature dependent contribution to the homogeneous linewidth of the R1( ) in ruby. Adapted from [30].

Some data on the temperature dependent contribution of the linewidths at low

temperature has also been measured by Muramoto et al. [32]. They obtained

constant values below 10 K.

Extensive measurements of the temperature dependence of the linewidth and

lineshift of the electronic origins of many materials have been made [33-38].

Raman

Direct

Direct + Raman

Temperature (K)

Co

ntr

ibu

tio

n t

o h

om

oge

neo

us

linew

idth

(M

Hz)

Page 29: spectral hole-burning and slow light in emerald and ruby

19

1.4. Laser spectroscopy

With the development of tunable lasers over the last decades there have been a

variety of techniques for increasing spectral resolution and extracting

information which is obscured by inhomogeneous broadening. The existing

techniques can be classified into two categories:

- Time-domain techniques, such as photon echoes and optical free

induction decay.

- Frequency-domain techniques, such as fluorescence line-narrowing

and spectral hole-burning.

Spectral hole-burning and fluorescence line-narrowing, which have been used

in the experiments of the present thesis, are explained in the following

paragraphs.

1.4.1. Spectral hole-burning

There are two types of hole-burning: persistent and transient hole-burning.

Within persistent hole-burning, one must distinguish between photochemical

and non-photochemical hole-burning.

- Photochemical HB: the so-called “photoproduct” absorption band is

well separated from the original absorption band. This type of hole

can persist infinitely at low temperatures.

- Non-photochemical HB (NPHB): this kind of hole is typical for

amorphous systems. In this case the “photoproduct” is within the

inhomogeneous absorption line.

- Transient HB (THB): in this case, a fraction of the population is stored

in any excited level. A narrow-band laser of width Гl and frequency fl

excites molecules within the inhomogeneous broadening that absorb

Page 30: spectral hole-burning and slow light in emerald and ruby

20

resonantly with the laser. A hole is created in the original absorption

band. The hole width can approach the homogeneous width Гhom in

ideal cases [10].

1.4.1.1. Transient spectral hole-burning

Some of the optical centres lying in the inhomogeneously broadened transition

are excited. The lifetime of the transient hole is given by the lifetime of the state

in which the population is stored.

When high powered lasers are needed, care must be taken in order avoid

excessive heating of the sample. A larger broadening of the transition is

obtained otherwise. Obviously, stimulated emission takes part in the readout

process if the population is stored in the excited state [8, 11].

Transient hole-burning experiments can be performed by using one or two

single-frequency lasers. If only one laser is used, this has to be scanned by fast

modulation, e.g. current. When two lasers are used, the first burns the hole at a

particular frequency and the second laser scans the spectrum of the hole [10,

39, 40]. In this project, transient hole-burning has been conducted by using a

single laser.

The first transient hole-burning experiments in the solid state were carried out

by Szabo in the 1970’s [39]. However, hole-burning was observed for the first

time in condensed matter in the 1960’s [41].

The value of Гhom can be estimated as Гhom~ 0.5 Гhole– Гl, where Гl is the laser

linewidth.

Page 31: spectral hole-burning and slow light in emerald and ruby

21

Figure 1.11. Representation of mechanisms for transient hole-burning. From left to right: direct depletion of ground state, relaxation to metastable excited state and relaxation to metastable sublevel of the ground state. Adapted

from [11].

1.4.1.2. Persistent hole-burning

There are two possible mechanisms to create persistent hole-burning. Their

names are photochemical hole-burning (PHB) and non-photochemical hole-

burning (NPHB).

In PHB optical centres in the excited state can be subject to a photochemical

reaction and a shift in their transitional frequency results leading to a decrease

of the absorbance at that particular frequency. As a result the spectrum

changes and there is a photoproduct in the absorption spectra outside of the

inhomogeneous linewidth [5, 8, 11].

Ground state

Read

laser

Bu

rn laser

Excited state

Page 32: spectral hole-burning and slow light in emerald and ruby

22

Figure 1.12. Persistent hole-burning. Appearance of photoproduct and difference between spectra before and after burning. Adapted from [11].

On the other hand, NPHB is more likely to happen in amorphous structures

rather than in crystalline systems. The idea is that the appearance of anti-holes

occurs in proximity to the burned hole due to rearrangements of host-guest

interactions [42-44].

1.4.1.3. Potential applications of spectral hole-burning

Hole-burning is not only important for being a tool to unravel details of the

electronic structures of ions. It has potential applications in quantum computing,

laser stabilisation and in frequency and time domain optical storage (particularly

persistent spectral hole-burning).

Chemical modification

Photoproduct

Page 33: spectral hole-burning and slow light in emerald and ruby

23

When a narrow bandwidth laser excites those centres resonant with the laser

frequency, multiple spectral holes can be burned within the absorption line. The

absence or presence of spectral holes at given frequencies can be used to

represent binary data. The ratio of inhomogeneous to homogeneous linewidth,

determines the storage capacity in the frequency domain and is the figure

of merit (FOM) for this application [45, 46].

One of the major challenges is to find materials with a high storage capacity at

room temperature, i.e. a high FOM.

1.4.2 Fluorescence line narrowing (FLN)

In the 60’s Denisov et al. [47] were able to narrow the luminescence spectrum

of Eu3+ in a glass by means of mercury lines. FLN was also demonstrated in a

crystal by Szabo in 1970 [48, 49]. He observed a narrowing of the R1 line in

ruby when using resonant ruby-laser excitation. By means of a narrow band

laser, a subset of optical centres within an inhomogeneously broadened

absorption line is excited and the fluorescence is dispersed with a

monochromator or a Fabry-Pérot interferometer. The subset of optical centres

in the inhomogeneous broadened transition has the same transition energy. As

the energy levels in crystalline systems are very correlated, this is a site

selective technique in these systems [8]. It is another method of approaching

the homogeneous width of an optical transition.

The ions or molecules that are in resonance with the laser are excited and only

these molecules or ions will emit light. As a consequence much narrower lines

will appear in the fluorescence spectrum in comparison with non-selective

excitation [3, 5]. When the luminescence is observed resonantly with the laser

frequency, this luminescence must be separated from the laser light by means

of gating techniques [8]. Mechanical choppers can be used when the excited

Page 34: spectral hole-burning and slow light in emerald and ruby

24

state lifetime is longer than 10 µs. However for shorter excited state lifetimes

acoustic-optic or electro-optic modulators are recommended [1].

FLN processes in resonant experiments involve two photons so the linewidth

observed is twice the homogeneous linewidth. The optical centre is excited by

one photon and another photon is observed in emission. In non-resonant

experiments the lifetime and correlation of the third energy level affect the

measured linewidth [1, 8].

Figure 1.13. FLN technique, illustrating both resonant and non-resonant experiments. Adapted from [8].

1.5. Slow light

The concept of dispersive wave propagation and the group velocity has been

interrelated historically. In 1839, Sir William R. Hamilton [50] introduced the

concepts of phase and group velocities when he explained dispersive wave

Reso

nan

t FLN Laser excitation

No

n-reso

nan

t FLN

Page 35: spectral hole-burning and slow light in emerald and ruby

25

propagation as a coherent superposition of monochromatic wave disturbances.

Rayleigh [51] attributed the original definition of group velocity to Stokes [52]

claiming that: 'when a group of waves advances into still water, the velocity of

the group is less than that of the individual waves of which it is composed; the

waves appear to advance through the group, dying away as they approach its

anterior limit. This phenomenon was, I believe, first explained by Stokes, who

regarded the group as formed by the superposition of two infinite trains of

waves, of equal amplitudes and of nearly equal wavelengths, advancing in the

same direction'. Rayleigh [53] explained the difference between phase and

group velocity by using those results.

The understanding of both phase and group velocity is crucial to interpret the

results described in chapter 3. Let us start explaining some concepts related to

wave propagation [54]. The total electric field E(t) of a non-monochromatic wave

formed by two waves with amplitude A(t), angular frequencies and , wave

numbers and propagating along the z axis, is [55]:

Eq. (1.10)

By applying a trigonometric transformation, we obtain:

Eq. (1.11)

where , , and . If

both frequencies and are of the same order, Eq. (1.11) can be considered

as a non-monochromatic wave expressed by the product of a carrier wave (high

wave number and frequency ) and a modulation wave (low wave number

and frequency

). The carrier wave propagates at phase velocity and

the modulation wave propagates at [56].

The intensity of the field can be considered as

Eq. (1.12)

Page 36: spectral hole-burning and slow light in emerald and ruby

26

Eq. 1.12 the first term represents the carrier term oscillating at a frequency

and the second term represents the envelope travelling at a low frequency

, which is the frequency of the intensity modulation and modulates the

first term.

The group velocity of a wave is the velocity with which the intensity modulation

of the wave propagates though space whilst the phase velocity is the velocity at

which the carrier term displaces.

Figure 1.14. Phase and group velocity of a wave. Adapted from [54].

From the phase velocity equation

, it can be observed that each

monochromatic wave propagates at different phase velocities. In the case that

the monochromatic waves have close frequencies, the group velocity can be

expressed as follows [57]:

Eq. (1.13)

where is the real part of the refractive index of a medium and is frequency

dependent (because each wave comprises many frequencies travelling at

Page 37: spectral hole-burning and slow light in emerald and ruby

27

different phase velocities). From Eq. (1.13) it follows that the group velocity can

be smaller (slow light) or larger (superluminal) depending on the value of

.

Slow light can be defined as a reduction in the group velocity of the light (vg<<c)

and fast light as an increase in the group velocity (vg>c or vg is negative) [58].

When it comes to an absorbing medium, the group velocity can be influenced

by the refractive index and the real part of the refractive index can be

frequency dependent in a narrow optical resonance of the medium. It is in that

narrow optical region is where slow light takes place [57].

Figure 1.15. Horizontal axis represents optical frequency. Rapid change of the index of refraction (blue line) in a region of rapid change of absorption (gray line). The steep and linear region of the refractive index in the center

gives rise to slow light. Adapted form [59].

There are several methods to create slow light. The common factor is to create

a narrow spectral region with high dispersion. The methods are usually

classified in two groups [60]:

- Material dispersion, such as Electromagnetically Induced

Transparency (EIT), Coherent Population Oscillation (CPO), hole-

burning (HB) and Four Wave Mixing (FWM). They all produce a rapid

change in refractive index as a function of the frequency. A nonlinear

effect is used to modify the dipole response of a medium to a “probe”

field,

Rapid change of

absorption

Index of

refraction

Page 38: spectral hole-burning and slow light in emerald and ruby

28

- and waveguide dispersion, such as photonic crystals or other

resonator structures which modify the k-component of a wave

(structural dispersion).

1.5.1. Kramers- Kronig relations

In order to produce slow light in a system it is necessary that the refractive

index changes rapidly as a function of frequency. This is possible when the

material is in resonance (or very close) with the applied optical field. The

Kramers- Kronig relations are presented below. These equations relate the real

part of the refractive index and the absorption in a material [61]:

Eq. (1.14.a)

Eq. (1.14.b)

By analysing Eqs. (1.14.a) and (1.14.b) it can be observed that a narrow dip in

an absorption spectrum will lead to strong normal dispersion (dn/dw>>0) and

that a gain will lead to anomalous dispersion (dn/dw<<0). Figure 1.16 illustrates

the first case.

Figure 1.16. Absorption and refractive index are related though the Kramers-Kronig relations. A narrow spectral dip leads to strong normal dispersion. Adapted from [62].

Page 39: spectral hole-burning and slow light in emerald and ruby

29

The group index of a pulse can be expressed by Eq. (1.15).

Eq. (1.15)

It is obvious that if the dispersion is large, the group index can be made large as

well. Hence, the conclusion is that in order to produce slow light in a material, a

process that leads to narrow and strong features must be undertaken. By

Kramers-Kronig relations, this characteristic will lead to the necessary large

dispersion for slow light.

1.5.2 Methods to create slow light

The most relevant methods to create slow light are discussed in the following

sections.

1.5.2.1. Electromagnetically induced transparency (EIT)

It is well-known that a rapid change in the refractive index near the resonance of

a material can lead to large values of the group index. Low group velocities are

accompanied by strong absorption so the observation of this effect is very

difficult. In order to reduce this strong absorption the technique of

electromagnetically induced transparency has been used to render the material

medium very transparent and retain the strong absorption necessary to observe

slow light propagation [63-66]. In the 1990s Harris et al. observed low group

velocities for the first time in strontium [67-68].

Using this technique Kasapi et al. observed a group velocity of vg=c/165 in a 10

cm long Pb vapour cell [69]. The lowest group velocity observed so far in an EIT

medium was by Budher et al in 1999. Group velocities as low as 8 m/s were

observed in a warm thermal rubidium vapour [70]. In the same year, Hau and

Page 40: spectral hole-burning and slow light in emerald and ruby

30

his collaborators [71], used this method in order to produce slow light. The

group velocity of light was as slow as 17 m/s in an ultracold atomic gas at 450

nK. In the first experiments of ultra-slow light experiments in a solid, Turukhin et

al. observed a velocity of 45 m/s corresponding to a group delay of 66 μs in a

praseodymium doped Y2SiO5 crystal [72].

EIT is a coherent optical nonlinearity which presents a transparent medium over

a narrow spectral range within an absorption line. Dispersion is created within

the transparency window leading to slow light [73].

In order to observe slow light by means of this procedure, two optical fields

must be used (for example, two lasers). These optical fields must interact with

three quantum states of a material. The probe field (wp) and a much stronger

coupling field (also called pump or control, wc) are tuned near resonance at

different transitions. Hence, EIT can be achieved in a three-level atom with a

coupling and a probe field to produce a two-quantum coherence with a long

lifetime. There is population trapping in a dark, non-absorbing state and this

produces the required coherence. The atoms are trapped in this state and

cannot be excited due to destructive interference between transition paths to

the excited state. As a consequence of the existence of the dark state

coherence, there is a narrow "hole" in the absorption line for the probe. As the

Rabi frequency induced by the coupling field is larger than the holewidth, a

narrow EIT hole will be produced [74].

Figure 1.17. in order to produce EIT a control beam, wc, ia applied between I1> and I2>. This splits level I2> and a probe beam sees less absorption over a narrow spectral range.

strong

control

beam probe beam

I3>

I1>

I2>

Page 41: spectral hole-burning and slow light in emerald and ruby

31

In a configuration of three states at least two out of the three possible transitions

between the states must be dipole allowed and the other dipole forbidden. One

of these three states is connected to the other states by means of the probe and

coupling field.

As is described in the literature coherent preparation of the medium produces

remarkable changes in the optical properties of a molecular medium or an

atomic gas [75, 76]. The laser-induced coherence of atomic states is the cause

of the modification of the optical response of an atomic medium. This will lead to

quantum interference between the excitation paths which control the optical

response. When quantum interference between the excitation pathways

happens, linear susceptibility can be removed (absorption and refraction) at the

resonant frequency of the transition [77, 78].

A relationship exists between the optical properties of molecular and atomic

gases and their energy-level structures. The first order susceptibility χ(1) can

describe an atom that responds linearly to resonant light. This susceptibility can

be divided in two factors: Re(χ(1)) and Im(χ(1)). The real part determines the

refractive index and the imaginary part represents the absorption. As we can

see in Figure 1.18 the imaginary part as a function of frequency can be

represented by a Lorentzian function while the real part can be approximated by

a dispersion profile.

Figure 1.18. The picture to the left is the imaginary part of χ(1)

and to the right is the real part of χ(1)

.

(wp-w31)/γ31 (wp-w31)/γ31

Im(χ

(1) )

Re(χ

(1) )

Page 42: spectral hole-burning and slow light in emerald and ruby

32

Figure 1.18 shows the susceptibility as a function of the probe field wp with

respect to the atomic resonance frequency w31. The dashed line is for a

radiatively broadened two-level system with width γ31 while the solid line is an

EIT system with resonant coupling field.

In conclusion, EIT creates a narrow transparency window in the absorption line

and when a probe pulse propagates through that window it may be delayed.

This results in slow light [73, 79]. As dictated by the Kramers-Krӧnig (see

section 1.5.1.) relations a change in absorption in a narrow spectral range yields

a rapidly changing refractive index in that region. This rapid change produces a

very low group velocity.

1.5.2.2. Coherent population oscillation (CPO).

Years after EIT was developed, several groups investigated different ways to

apply slow light to more realistic fields and to overcome the hole bandwidth

problem in EIT experiments. In this section some aspects of slow light by

means of coherent population oscillations are analysed [58].

CPO involves the creation of a "spectral hole" due to population oscillations.

The "hole" is due to the fact that the ground state population is modulated at the

beat frequency between the probe and pump field applied to the material.

"Spectral holes" as narrow as 36 Hz lead to group velocities of 57 m/s [58].

In 1967 Schwartz predicted the appearance of these "holes" [80]. In 1983,

Hillman observed for the first time a CPO "hole" in ruby [81]. An argon-ion laser

was used in these experiments to pump population from the ground state to the

4T2 absorption band. The decay from this level to the metastable 2Ā and Ē

levels of 2E is fast (on the order of picoseconds) and from these levels to the

ground state is on the order of a few miliseconds (T’1). A "hole" at the laser

frequency is created as a consequence of that long lifetime. The width of that

hole was 37 Hz which is the inverse of the population relaxation time.

Page 43: spectral hole-burning and slow light in emerald and ruby

33

More recently Bigelow [82] carried similar experiments for the observation of

slow light in ruby at room temperature. In Figure 1.19 the 4A2 ground state is

denoted as a, the levels 2Ā and Ē are denoted as c and the 4T2 absorption band

is b. As said above level b decays very rapidly so the system is reduced to a

two-level system. In the picture T1 is the ground state recovery time and it is

twice the lifetime of level c T’1. T2 is the dipole moment dephasing time.

Figure 1.19. Schematic diagram of energy levels in ruby. Adapted from [82].

Bigelow et al. claim that when light with a frequency w1 is in resonance with the

transition a-b (with the intensity modulated at interacting with a system

as shown in Figure 1.19 there will be a decrease in the ground state population

Ng because of absorption. The ground state re-populates by transition c-a. Its

state population will be modulated at the modulation frequency and will be

delayed with respect to the incident intensity. The metastable level lifetime 1/T1’

and the recovery time of the pumping cycle are of the same order. The

modulation depth of the ground state population Ng is larger when the pumping

intensity has a modulation period similar to that of the metastable level

lifetime. When that resonance happens a narrow overall resonant behaviour

occurs and a decrease in absorption is observed. By Kramers-Kronig relations,

the real part of the refractive index nr depends on the ground state population

Ng. As Ng is frequency dependent, that dependence will give large values of

near resonance leading to small values of group velocities.

Bigelow [82] suggested that a beam with amplitude modulation experiences

large group index due to the spectral dip observed. In Figure 1.20 the delay as

a

b

a

b

c

W1

Rapid decay

Page 44: spectral hole-burning and slow light in emerald and ruby

34

a function of the modulation frequency for different pump powers is shown. It

was found that the largest delay corresponded to the deepest and narrowest

hole. Bigelow et al. reported a group velocity of 57 m/s and they claimed that

slow light can be observed only by applying a single intense pulse of light to

induce the saturation needed for the slow light, i.e. no separation of pump and

probe waves is required.

Figure 1.20. Time delay versus modulation frequency for different imput powers. The inset shows the normalized 60 Hz input and output signal at 0.25 W. There is a delay of 612 µs which corresponds to 118 m/s. Adapted from [82].

Different Gaussian pulses were applied to observe the delay when propagating

through the ruby. The longer pulses experienced the longer delays.

Page 45: spectral hole-burning and slow light in emerald and ruby

35

Figure 1.21. Normalized input and output intensities of different pulse lengths with the corresponding group velocities. The inset shows a close-up of the 20 ms pulse. Adapted from [82].

However, other researchers countered that the above CPO investigations are

the consequence of lack of physical understanding of these effects. Zapasskii et

al. claim that the discussed results presented above can be regarded as a

consequence of the model of a saturable absorber [83].

A saturable absorber is a layer of an optical medium whose absorption shows

saturation with increasing light intensity. Various groups [83-86] show how the

pulse delay is the result of any medium with nonlinear intensity and there is no

evidence of any change in the light group velocity. They say that when the light

interacts with a nonlinear medium a change in the intensity spectrum is

produced. The interpretation of these effects does not need the introduction of

concepts such as slow light or group velocity reduction [84-86]. In particular,

Selden claims that there is no demonstration of group velocity reduction since

the results are compatible with saturable absorption theory [86].

Separation of the probe and pump beams seems to be a minimum requirement

for making the difference between hole-burning and saturable absorption in

slow light experiments. In this way, the depth of the hole would be reduced,

Page 46: spectral hole-burning and slow light in emerald and ruby

36

which should produce a reduction in the slow light. Only when a very narrow

source is used to scan the absorption spectrum in the homogeneous linewidth a

hole can be created in the absorption line [87].

CPO experiments on slow light are not well interpreted because the observed

phase shift is treated as a signal transit time. This is used to calculate a group

velocity that can be done very small by reducing the length of the sample

(vg=L/Td≥L/Ts) where Td is the observed delay and Ts is the metastable lifetime.

A correct interpretation is that the transit time depends on sample length and it

should be larger than the relaxation time of the absorber in a long sample [88].

As a conclusion, the results of slow light by means of CPO do not demonstrate

unequivocally group velocity reduction.

1.5.2.3. Hole-burning.

Rebane et al. proposed to produce slow light by means of persistent spectral

hole-burning [89, 90]. A burn and a probe laser are applied to the system. The

burn laser creates a transparency window for the probe, creating narrow

spectral filters with very high contrast and, as a consequence, with high

dispersion.

Let us consider a weak probe pulse of amplitude launched into a medium at

the input end. In the case of a weak probe, the molecules behave as classical

damped harmonic oscillators and the interaction between the spectral hole-

burning medium and the probe pulse can be interpreted within classical pulse

propagation in a linear absorbing medium. The change in the phase and

amplitude can be expressed as

Eq. (1.16)

where is the complex frequency domain amplitude response function

expressed as

Page 47: spectral hole-burning and slow light in emerald and ruby

37

Eq. (1.17)

is the intensity transmission through the sample in the longitudinal

direction at frequency expressed as where is the

absorption coefficient and the length of the medium. is the phase of the

response function calculated by employing the Hilbert transform:

Eq. (1.18)

In Eq. (1.19) represents the frequency domain amplitude of the probe,

Eq. (1.19)

Eq. (1.20) expresses the time domain intensity at the output of the medium,

Eq. (1.20)

which is calculated for a Gaussian probe pulse.

Slow light takes place when all frequency components experience the same

attenuation and when the phase changes linearly with the frequency.

In 2004, spectral hole-burning was used to create slow light. Fan et al. [91]

slowed down light in a solid to approximately 43 m/s through the change in the

refractive index of the medium at room temperature.

In this thesis, this theory is applied to transient hole-burning in the R1-line of

chromium (III) in ruby.

Page 48: spectral hole-burning and slow light in emerald and ruby

38

1.6. References

1. Henderson, B.; Imbusch, G. F.; Optical spectroscopy of inorganic solids.

Oxford Science Publications (1989).

2. Demhelt, H.; Paul, W.; Ramsey, M. F.; Rev. Mod. Phys. 62, 595 (1990).

3. Riesen, H.; Coord. Chem. Rev. 250, 1737-1754 (2006).

4. Einstein, A.; Ann. Phys. 17, 132 (1905).

5. Skinner, J. L.; Moerner, W. E.; J. Phys. Chem. 100, 13251-13262 (1996).

6. http://farside.ph.utexas.edu/teaching/315/Waves/node9.html.

7. Skinner, J. L.; Moerner, W. E.; J. Phys. Chem. 100, 13251-13262 (1996).

8. Krausz, E.; Riesen, H.; Inorganic electronic structure and spectroscopy. Vol

I; Methodology; Solomon, E. I.; Lever, A. B. P.; Wiley: New York (1999).

9. Hayward, B. F.; BSc (Hons) Thesis: UNSW (2004).

10. Volker, S.; Annu. Rev. Phys. Chem. 499-530 (1989).

11. Riesen, H.; Structure and bonding, 107, 179-205 (2004).

12. Schenzle, A.; Brewer, R. G.; Phys. Rev. A 14, 1756 (1976).

13. Stoneham, A.; Rev. Mod. Phys. 41, 82 (1969).

14. Jaaniso, R.; Hagemann, H.; Bill, H.; J. Chem. Phys. 101, 10323 (1994).

15. Skinner, J. L.; Mashl, R. J.; Orth, D. L.; J. Phys.: Condens. Matter 5, 2533

(1993).

16. MacFarlane, R. M.; Springer Series in Optical Sciences 54 (Lasers,

Spectrosc. New Ideas), 205-23 (1987).

17. MacFarlane, R. M.; J. of Lum. 100, 1-20 (2002).

Page 49: spectral hole-burning and slow light in emerald and ruby

39

18. Weber, M. J.; Phys. Rev. 156 (1967).

19. MacFarlane, R. M.; Meltzer, R. S.; Malkin, B. Z.; Phys. Rev. B 58, 5692

(1998).

20. MacFarlane, R. M.; Shelby, R. M.; Burum, D.; Opt. Lett. 6, 593 (1981).

21. Sugano, S.; Tanabe, Y.; Kamimura, H.; Multiplets of Transition Metal ions;

Academy Press: New York (1970).

22. Greenwood, N. N.; Earnshaw, A.; Chemistry of the Elements 2nd Edition;

Butterworth-Heinemann (1997).

23. Mabbs, F. E.; Magnetism and transition Metal Complexes; Chapman and

Hall: London (1973).

24. Milos, M.; Kairouani, S.; Rabaste, S.; Hauser, A.; Coord. Chem. Rev. 252,

2540-2551 (2008).

25. Imbusch, G. F.; Kopelman, R.; Laser spectroscopy of solids. Topics in

Applied Physics 49; Imbusch, P. M.; Springer-Verlag: Berlin (1981).

26. Riesen, H.; Krausz, E.; Comments Inorg. Chem. 14, 323 (1993).

27. McCumber, D. E.; Sturge, M. D.; J. Appl. Phys. 13, 1682 (1963).

28. Hsu, D.; Skinner, J. L.; J. Chem. Phys. 83, 2107 (1985).

29. Blume, M.; Orbach, R.; Kiel, A.; Geschwind, S.; Phys. Rev. 139, A314

(1995).

30. Riesen, H.; Szabo, A.; Chem. Phys. Lett, 484, 181-184 (2010).

31. Rives, J. E.; Meltzer, R. S.; Phys. Rev. B 16, 1808 (1977).

32. Muramoto, T.; Fukuda, Y.; Hashi, T.; Phys. Lett. A 48, 181 (1974).

33. Imbusch, G. F.; Yen, W. M.; Schawlow, A. L.; McCumber, D. E.; Sturge, M.

D.; Phys. Rev. 133, A1029 (1964).

34. Bartolo, B. Di.; Peccei, R.; Phys. Rev. 137, A1770 (1965).

Page 50: spectral hole-burning and slow light in emerald and ruby

40

35. Yen, W. M.; Selzer, P. M.; Laser spectroscopy of solids 2nd Edition; Springer

series on Topics in Appl. Physics p. 49 (1985).

36. Kushida, T.; Kikuchi, M.; J. Phys. Soc. Japan 23, 21333 (1967).

37. Gourley, J. T.; Phys. Rev. B 5, 22 (1972).

38. Babbitt, W. R.; Lezama, A.; Mossberg, T. W.; Phys. Rev. B 39, 1987

(1989).

39. Szabo, A.; Phys. Rev. B 11-4512 (1975).

40. MacFarlane, R. M.; Shelby, R. M.; J. of Lum. 36, 179 - 207 (1987).

41. Spaeth, M. L.; Soy, W. R.; J. Chem. Phys. 23, 113 (1984).

42. Hayes, J. M.; Jankowiak, R.; Small, G. J.; Moerner, W.; Persistent hole-

burning science and applications. Topics in current Physics 44 (1988).

43. Holliday, K.; Manson, N. B.; J. Phys. Condensed Matter 1, 1339 (1989).

44. Riesen, H.; Bursian, V. E.; Manson, N. B.; J. Lumin. 85, 107 (1999).

45. Moerner, W. E.; Gehrtz, M.; Huston, A. L.; J. Phys. Chem. 88, 25, 6459-

6460 (1984).

46. Lenth, W.; Moerner, W. E.; Opt. Commun. 58, 4 (1986).

47. Denisov, Y. V.; Kizel, V. A.; Opt. Spectr. 23, 251 (1967).

48. Szabo, A.; Phys. Rev. Lett. 25, 14 (1970).

49. Selzer, P. M.; Huber, D. L.; Barnett, B. B.; Yen, W. M.; Phys. Rev. B 17, 12

(1978).

50. Hamilton, W. R.; Proc. R. Ir. Acad. I 341 (1839).

51. Rayleigh; Proc. Lond. Math. Soc. IX 21 (1877).

52. Stokes, G. G.; Mathematical and Physical papers 5 (Cambridge University

Press, Cambridge) p. 362 (1905).

Page 51: spectral hole-burning and slow light in emerald and ruby

41

53. Rayleigh; Nature XXIV 52 (1881).

54. http://mathpages.com/home/kmath210/kmath210.htm. Phase, group and

signal velocity.

55. http://www-eaps.mit.edu/~rap/courses/12333_notes/dispersion.pdf.

56. Jenkins, F. A.; White, H. E.; Fundamentals of optics; New York: McGraw-

Hill. p. 223 (1957).

57. Zhao, Y.; Zhao, H-W.; Zhang, Z-Y.; Yuan, B.; Zhang, S.; Optics and Laser

Technology 41, 517-525 (2009).

58. Bigelow, M. S.; Lepeshkin, N.; Boyd, R. W.; Phys. Rew. Lett. 90, 11

(2003).

59. http://benabb.wordpress.com/2009/06/09/sci-fi-reality-the-basis-of-reality-in-

all-astrophysics/

60. Hau, L. V.; Harris, S.E.; Dutton, Z.; Behroozi, C. H.; Nature 397 (1999).

61. Lucarini, V.; Saarinen, J.; Peiponen, K. E.; Vartiainen, E. M.; Kramers-

Krönig relations in Optical Materials Research, Springer, Heidelberg (2005).

62. Saleh, B.E.A.; Fundamentals of Photonics 2nd Edition; John Wiley & Sons

Inc., New Jersey (2007).

63. Brillouin, L.; Wave propagation and group velocity, Academic Press, New

York (1960).

64. Boyd, R. W.; Gauthier, D. J.; Progress in Optics XLIII, Wolf, E.; Elsevier,

Amsterdam p. 497-530 (2002).

65. Tewari, S. P.; Agarwal, G. S.; Phys. Rev. Lett. 56, 1811 (1986).

66. Bennink, R. S.; Phys. Rev. A 63, 033804 (2001).

67. Harris, S. E.; Phys. Rev. Lett. 62, 1033 (1989).

68. Harris, S. E.; Phys. Rev. Lett. 70, 552 (1993).

Page 52: spectral hole-burning and slow light in emerald and ruby

42

69. Kasapi, A.; Phys. Rev. Lett. 74, 2447 (1995).

70. Budker, D.; Kimball, D. F.; Rochester, S. M.; Yashchuk, V.V.; Phys. Rev.

Lett. 83, 1767-1770 (1999).

71. Hau, L.V.; Harris, S.E.; Dutton, Z.; Behroozi, C.H.; Nature 397, 594-8

(1999).

72. Turukhin, A. V.; Phys. Rev. Lett. 88, 023602 (2002).

73. Goldfarb, F.; Ghosh, J.; David, M.; Ruggiero, J.; Chaneliere, T.; Le Gouet, J.

L.; Gilles, H.; Ghosh, R.; Bretenaker, F.; Lett. Journal Expl. 82, 54002

(2008).

74. Yannopapas, V.; Paspalakis, E.; Vitanov, N. V.; Phys. Rev. B 80, 035104

(2009).

75. Fleischhauer, M.; Imamoglu, A.; Marangos, J. P.; Rev. Of Modern Phys. 77

(2005).

76. Klein, M.; Hohensee, M.; Xiao, Y.; Kalra, R.; Phillips, D. F.; Walsworth, R.

L.; Phys. Rev. A 79, 053833 (2009).

77. Hansch, T. W.; Toschek, P.; Z. Phys. 236, 213 (1970).

78. Gray, H. R.; Whitley, R. M.; Stroud, C. R.; Opt. Lett. 3, 218 (1978).

79. Arimondo, E.; Prog. Opt. 35, 259 (1996).

80. Schwartz, S. E.; Tan, T. Y.; Appl. Phys. Lett. 10, 4 (1967).

81. Hillman, L. W.; Opt. Commun. 45, 416 (1983).

82. Bigelow, M. S.; Ultra-slow and superluminal light propagation in solids at

room temperature; Thesis: University of Rochester (2004).

83. Zapasskii, V. S.; Kozlov, G. G.; Non linear and quantum optics, optics and

spectroscopy 104, 1, 95-98 (2007).

84. Zapasskii, V. S.; Kozlov, G. G.; Optics Express 17, 24, 22154 (2009).

Page 53: spectral hole-burning and slow light in emerald and ruby

43

85. Macke, B.; Segard, B.; Phys. Rev. A 78, 013817 (2008).

86. Selden, A. C.; Optics and spectroscopy 106, 6, 881-888 (2009).

87. Ku, P. C.; Sedgwick, F.; Chang-Hasnain, C. J.; Opt. Lett. 29, 2291 (2004).

88. Boyd, R. W.; Gauthier, D. J.; Gaeta, A. L.; Willner, A. E.; Phys. Rev. A 71

023801 (2005).

89. Rebane, A.; Shakhmuratov, R. N.; Megret, P.; Odeurs, J.; J. of Lum. 127,

22-27 (2007).

90. Shakhmuratov, R. N.; Rebane, A.; Megret, P.; Odeurs, J.; Phys. Rev. A 71,

053811 (2005).

91. Fan, B.H., Zhang, Y.D., Yan, P.; Phys. Rev.Lett.; 54, 10, 4692-4695

(2005).

Page 54: spectral hole-burning and slow light in emerald and ruby

44

Chapter 2

Experimental

2.1. Sample preparation

Ruby and emerald crystals were cut with a water cooled diamond blade in a

Shell-Lap Gemmasta saw. The samples were polished using 1 µm, 0.25 µm

and 0.1 µm diamond paste on a Pedemin wheel. All the crystals were carefully

cleaned before they were used in optical experiments.

2.2. Room temperature spectroscopy

Non-polarised absorption spectra at room temperature were measured by a

Varian Cary 50 UV-Vis spectrometer. PC software facilitated the calibration and

baseline corrections.

2.3. Spectroscopy at Liquid Helium Temperatures

Very low temperatures (< 10 K) are necessary to observe details of electronic

structure that are obscured at higher temperatures. In this section an

explanation is provided about the steps required for correct mounting and

cooling of the samples. Also a brief description of the instrumentation used in

this project is given.

Page 55: spectral hole-burning and slow light in emerald and ruby

45

2.3.1. Mounting of samples

The sample was cleaned with hexane and was mounted over a hole in a brass

disc. The sample was imbedded in heat conducting cry-con copper grease. The

grease is used to ensure maximum thermal conductivity between the cold finger

of the cryostat and the sample. The sample holder was then attached to the

cold finger of the closed-cycle refrigerator (CCR).

Figure 2.1. Schematic of sample holder.

Crystal

sample

Brass plate

Cry-con

grease

Page 56: spectral hole-burning and slow light in emerald and ruby

46

2.3.2. Cooling method and temperature control

A Janis/Sumitomo SHI-4.5 closed-cycle refrigerator (CCR) was used in order to

cool down the samples. Temperatures as low as 2.5 K can be reached with the

use of this CCR without the need of liquid nitrogen and liquid helium. The CCR

uses a two-stage Gifford-McMahon cycle to cool down the cold finger. The cold

head is separated from the compressor unit by means of two high pressure

hoses. These hoses carry the expanded and compressed helium out and into

the cold head. The cold finger of the CCR is cooled to low temperature when

the helium gas is expanded [1].

In order to isolate the cold finger from the external environment, it is enclosed in

a radiation shield and a vacuum jacket. In this way convective and radiative

energy transfer between the environment and the CCR is minimized.

The vacuum jacket was evacuated using a turbomolecular pump to pressures

<8x10-4 mBar before cooling the system. The cooling process to 2.5 K takes at

least two hours. When the lowest temperature was reached a pressure of

around 10-6 mBar was measured. To measure and control the temperature at

the sample a Lakeshore Model 330 temperature controller and a 50 Ohm heater

was employed.

2.3.3. Absorption and Transmission spectroscopy

Low temperature transmission spectra were obtained using an Osram HLX

Xenophot 64840 bulb powered by a GW GPC-1850 dual-tracking power supply

to provide white light. The white light was passed through a polarising filter to

measure the polarised spectrum. In order to reduce luminescence effects in the

sample an OD filter was used to control the light power reaching the sample. A

200 mm lens was used to focus the light onto the sample and another 75 mm

Page 57: spectral hole-burning and slow light in emerald and ruby

47

lens was used to collimate the transmitted light after the cryostat. A third 200

mm lens was used to re-focus the light through a Thorlabs MC1000 opto-

mechanical chopper onto the entrance slit of a monochromator. The light was

then dispersed by a 1 m single monochromator (Spex 1404 equipped with a

1200 grooves/mm holographic grating). The light was detected using a side-

window photomultiplier tube (Hamamatsu R928). A current-to-voltage

preamplifier (Femto DLPCA-200) and digital lock-in amplifier (Stanford

Research Systems Model SR810 DSP) processed the signal from the

photomultiplier. Finally the signal was collected by a PC (see Figure 2.4). The

PC processed the transmission spectra in order to obtain absorption spectra

(dividing the baseline by the spectra obtained).

Eq. (2.1)

Io is the intensity of the light before the sample and I is the intensity after the

sample. The absorption spectra was obtained using the transmission spectra

by means of the relationship:

Eq. (2.2)

Similar experiments were conducted using a laser diode to obtain spectra with

much higher resolution.

2.3.4. Non-selective luminescence spectroscopy

For non-selectively excited luminescence spectra a Nd:YAG was used with a

wavelength of 532 nm to excite the sample. A polarising filter was used to find

out the polarisation dependence of the luminescence and also an OD filter to

reduce sample heating. 75 nm and 200 mm lenses were used to collimate the

emitted light. The luminescent light passed through a chopper, monochromator

Page 58: spectral hole-burning and slow light in emerald and ruby

48

and detected by a photomultiplier. A pre-amplifier and lock-in amplifier

processed the signal from the photomultiplier. Finally signals were collected on

a PC (see Figure 2.5).

2.3.5. Spectral hole-burning

The crystals were cut perpendicular and parallel to the c axis. Spectral holes

were burnt and read out by a stabilised laser diode (Thorlabs TEC2000

temperature controller with a ultra-low noise current source ILX LightWave LDX-

3620). After a short burn period at constant wavelength, the laser frequency

was tuned by modulating the injection current with a triangular waveform

generated by a waveform synthesizer (Stanford Research Systems SRS

DS345).

Figure 2.2 shows the ramp used to modulate the injection current using a diode

laser.

Figure 2.2. Waveform used to modulate the laser current. Typical frequency of ramp and burst are 2500 Hz and 830 Hz, respectively.

2500 Hz

Burst 830 Hz

Page 59: spectral hole-burning and slow light in emerald and ruby

49

Scans were calibrated by passing the modulated laser diode light through a

Fabry-Pérot etalon (1 mA increase in current corresponds to a frequency shift of

approximately 2.5 GHz).

The laser light was focused onto the sample after attenuation by using

polarising films and density filters. The laser light was measured in transmission

by a Si photodiode (Thorlabs PDA55). The signal was averaged on a digital

storage oscilloscope (Tektronik TDS210) and subsequently on a PC. Typically

1280 waveforms were averaged (see Figure 2.6).

Very narrow holes were measured at very low temperatures (< 14 K). For these

experiments an external cavity diode laser (Toptica DL110) was used. The

external cavity diode laser was locked to a 1.5 GHz Fabry-Pérot etalon and the

latter was scanned by a triangular ramp (see Figure 2.7).

2.3.6. Fluorescence line narrowing

Measurements were carried in two different ways. A Fabry-Pérot interferometer

(Burleigh RC-110) was used for intermediate temperatures (see Figure 2.9) up

to 80 K and the Spex 1704 monochromator for temperatures > 80 K (see Figure

2.8).

A Thorlabs TEC2000 temperature controller and a Thorlabs LDC500 current

controller controlled a Thorlabs TCLDM9 thermoelectric mount. Two opto-

mechanical choppers (Thorlabs MC 1000) operated with a phase shift of 1800

and a duty cycle of 30% were utilized to eliminate the laser light from reaching

the detector for measurements with the monochromator. Only one chopper was

necessary for FLN with Fabry-Pérot interferometer as the laser light was passed

though the chopper wheel as the FLN signal. The signal was processed as is

discussed in 2.3.3 and 2.3.4.

Page 60: spectral hole-burning and slow light in emerald and ruby

50

2.3.7. Slow light experiments

We have investigated the creation of slow light by transient spectral hole-

burning in a ruby crystal.

An external cavity diode laser (ECDL) was used. The amplitude modulation was

undertaken by an acoustico-optic modulator (Isomet 1205C-1; modulator driver

222A-1). The laser light was measured in transmission by a Si photodiode

(Thorlabs fast PDA10A). The signal was averaged on a digital storage

oscilloscope (LeCroy Wavesurfer 422) (see Figure 2.10).

Burn pulses with a period from 0.25 to 2 ms and a 50 ns Gaussian probe pulses

were separated by an interval of 0.01 to 0.25 ms as shown in Figure 2.3.

Figure 2.3. Burn and probe pulse.

0.01-0.25 ms

Burn pulse

0.25 - 2 ms

Probe pulse

50 ns

Page 61: spectral hole-burning and slow light in emerald and ruby

51

Figure 2.4. Schematic diagram for transmission measurement at liquid helium temperatures.

Photomultiplier

tube

CCR

sample Temperature

controller

24 V halogen

lamp

Power supply

Polarising

filter

Chopper

OD filter

200 mm lens

75 mm lens

200 mm lens

Chopper driver

Lock-in Amplifier

Computer

Pre-amplifier PMT

Monochromator

Page 62: spectral hole-burning and slow light in emerald and ruby

52

Temperature

controller

Figure 2.5. Schematic diagram for non-selective luminescence measurement at liquid helium temperatures.

Chopper driver

Lock-in Amplifier

Computer

Polarising

filter

Pre-amplifier

Chopper

OD filter

PMT

Monochromator

Nd:YAG

CCR

sample

200 mm lens

75 mm lens

200 mm lens

Page 63: spectral hole-burning and slow light in emerald and ruby

53

Figure 2.6. Schematic diagram for spectral hole-burning measurement at liquid helium temperatures.

Waveform generator

200 mm lens

200 mm lens

75 mm lens

Temperature

controller

Polarising

filter

OD filter

Current

Controller

Temperature

controller

Laser diode

CCR

sample

Mirror

Computer

Photo diode Digital oscilloscope

Page 64: spectral hole-burning and slow light in emerald and ruby

54

Figure 2.7. Schematic diagram for high resolution spectral hole-burning measurement at liquid helium temperatures.

Computer

Photo diode

Temperature

controller

Polarising

filter

OD filter

Current

Controller

Temperature

controller

external

cavity diode

laser

CCR

sample

Mirror

Shutter Waveform generator 2

optical

isolator

Fabry-

Perot

Photo-

detector

Waveform generator 1

Digital oscilloscope

DigiLock

110

Page 65: spectral hole-burning and slow light in emerald and ruby

55

Figure 2.8. Schematic diagram for fluorescence line narrowing measurement with a monochromator at liquid helium temperatures.

luminescence

Chopper driver 2

Lock-in Amplifier

Computer

Pre-amplifier PMT

Monochromator

Chopper 2

CCR

sample Chopper driver 1

Chopper 1

Temperature

controller

Current controller

Laser diode

Excitation laser

200 mm lens

200 mm lens

75 mm lens

sync

Page 66: spectral hole-burning and slow light in emerald and ruby

56

Figure 2.9. Schematic diagram for fluorescence line narrowing measurement with Fabry-Pérot interferometer at liquid helium temperatures.

prism

CCR

sample

Temperature

controller

Current controller

Fabry- Pérot

interferometer

Chopper driver Chopper

200 mm lens

75 mm lens

200 mm lens

Laser diode

Page 67: spectral hole-burning and slow light in emerald and ruby

57

Figure 2.10. Schematic diagram for slow light experiments.

2.4. Application of an external magnetic field

Uniform low strength magnetic fields were applied to the sample via external

Helmholtz coils. The coils are formed by two water-cooled copper wire coils

(250 mm diameter) mounted around the CCR and separated by a distance of

125 mm by plastic spacers (125 mm corresponds to the radius of the coil).

Currents in the range of 0 to 9.5 A were supplied by two power supplies

(Hewlett-Packard 6269B DC power supply). The magnetic field is a

Oscilloscope PC

ECDL Toptica

DL110 AOM

2 mm aperture

CCR

sample

150 mm lens

750 mm lens

Waveform generator

Photo

diode

Page 68: spectral hole-burning and slow light in emerald and ruby

58

superposition of the fields created by two sets of current loops. These external

Helmholtz coils behave according to the Biort-Savart law given by the following

Eq. (2.3) [2]:

Eq. (2.3)

In Eq. (2.3) B is the magnetic field strength, N is the number of turns in a coil, I

is the current, µ0 is the permittivity of vacuum (4π∙10-7 T∙m/A) and R is the

radius of the coil. The separation between the two coils must be equal to the

radius of the coils in order to have an optimally uniform magnetic field at the

centre position.

2.5. Laser sources

In this section the laser sources used in this project are presented. All the laser

sources used are potentially dangerous and without any precaution can cause

permanent eye injure. Keeping this in mind, all the OHS measures were taken

into account in order to reduce this risk. Moreover, the power of these lasers

was high enough to damage the photomultiplier used in this project, so care

was taken when working with these lasers.

Page 69: spectral hole-burning and slow light in emerald and ruby

59

2.5.1. Nd:YAG laser

Nd:YAG is a solid state laser that uses a neodymium doped yttrium aluminium

garnet crystal (Nd:Y3Al5O12). These lasers are optically pumped by a flashtube

or a diode laser. They lase at a wavelength of 1064 nm in the infrared [3].

Nd:YAG lasers can operate in both continuous and pulsed mode. In pulsed

mode an optical switch is incorporated into the laser cavity until a maximum

population inversion occurs in the neodymium ions before it opens. After that,

light can go through the cavity provoking a depopulation of the excited laser

medium. The high-intensity pulses can be frequency doubled to generate laser

light at 532 nm within the green range of the spectrum [4].

As shown in Figure 2.11 below, an external laser diode excites neodymium ions

to a higher lying excited state. There is a decay from 4F3/2 to 4I11/2 (a photon is

emitted) after a non-radiative decay from the higher excited level. Finally, non-

radiative decay processes take place to go back to the ground state.

Figure 2.11. The four level laser system Nd:YAG.

Non-Radiative decay

Non-Radiative decay

Photon emission

Excitation

1064 nm

Page 70: spectral hole-burning and slow light in emerald and ruby

60

This laser was used to perform non-selective luminescence experiments. It is a

class 3b laser product.

2.5.2. He-Ne laser

This type of laser was used for alignment of optical axis. The mechanism to

produce population inversion and light amplification in a He-Ne laser comes

from many collisions of electrons with ground state helium atoms in the gas

mixture.

As shown in Figure 2.12 helium atoms are excited from the ground state to the

excited state (23S1 and 21S0 metastable states) by electron collisions. As a

consequence of a coincidence between these two metastable states and the 3s

and 2s levels of neon, collisions between atoms lying in these next levels

provoke a transfer of energy from helium to neon [5]. The population in 3s and

2s levels of neon is increased. When the population of these two levels is larger

than that of the ground state in neon, then there is a population inversion. The

light can be amplified by the medium in the 2s to 2p transition and the 3s to 2p

transition. Very fast radiative decay from 2p level to the 1s state is produced. A

final decay makes the population reach the ground state.

Page 71: spectral hole-burning and slow light in emerald and ruby

61

Figure 2.12. Energy transfer between Helium and Neon ions. Adapted from [6].

An optical oscillator is created by placing highly reflecting mirrors at each side of

the amplifying media so that the wave is reflected back and forth gaining more

power. When this happens, there is a stable laser beam output.

2.5.3. Diode laser

A range of current and temperature stabilised single-frequency laser diodes

were used in this project in high-resolution laser experiments. Conventionally,

they are used in devices such as DVD players, laser pointers, optical fiber

communications, etc. The most important advantage is the small size, high

power efficiency and low maintenance requirements [7].

Energy

He-Ne collision

Fast radiative transitions

Ground

state

Diffusion to walls

Excitation by

electron collision

Page 72: spectral hole-burning and slow light in emerald and ruby

62

The active medium in laser diodes is a semiconductor. The simplest kind of a

laser consists of a p-n junction (homojunction laser) and is powered by an

injection of electric current. When it comes to p-i-n structure (heterostructure)

electrons and holes can recombine and photons are released as energy

quantas. This process can be either spontaneous or stimulated by incident

photons leading to optical amplification, and with optical feedback in a laser

resonator to laser oscillation. When an electric potential is applied through the

p-n junction, electrons relax from the conduction band to the valence band of

the material. In this process there is emission of radiation. The semiconductor

material determines the output wavelength. One of the most common

semiconductors used in laser diodes is GaAlAs, commonly found in CD players

[8-10].

Laser diode sources were the laser source in experiments such as spectral

hole-burning or fluorescence line narrowing. They were installed in a Thorlabs

TCLDM9 temperature controller mount which was controlled by a Thorlabs

TEC2000 thermoelectric temperature controller and a ILX ultra low noise

current supply. By changing the supplied current and the temperature of the

laser diode, the frequency of the emitted light can be controlled.

The wavelength was measured using a wavemeter (Coherent Scientific

Wavemate). A calibration of this instrument can be found in [11].

In our experiments the laser light from the laser diode was attenuated with a

density filter and focused onto the sample by means of a lens. The injection

current of the laser diode was modulated using a waveform generator (Stanford

Research Systems DS345) yielding the desired frequency modulation.

Page 73: spectral hole-burning and slow light in emerald and ruby

63

2.5.4. External Cavity Diode Laser

In order to perform high resolution laser spectroscopy of very dilute samples,

where homogeneous linewidths are very narrow, a laser with a very small

linewidth must be used. In many diluted systems the spectral hole widths

observed are mainly due to the linewidth of the laser. A Toptica DL100 External

Cavity Diode Laser was used.

In an External Cavity Diode Laser a lens and a diffraction grating are placed at

the front of a standard laser diode as shown in the picture below allowing a

decrease in the linewidth of the laser output.

Figure 2.13. External cavity diode laser diagram. Adapted from [10].

The two most common configurations are Littrow and Littman-Metcalf

geometries. In both configurations a collimating lens and a diffraction grating

are placed in front of the laser diode. In the first configuration, the diffraction

grating is rotated so that the wavelength output is tuned. The inconvenience is

that when tuning the direction of the grating, the direction of the laser beam is

changed. The second configuration resolves this problem. In this geometry, a

mirror reflects the first-order beam back to the laser diode. Then the tuning of

the frequency is made by rotating that mirror. A high degree of selectivity can be

done, so narrower linewidths can be obtained [10].

Output

LD Lens Diffraction grating

Page 74: spectral hole-burning and slow light in emerald and ruby

64

Figure 2.14. External cavity diode laser configurations: Littrow to the left and Littman-Metcalf to the right. Adapted from [10].

.

The ECDL laser used in this thesis was based on the Littrow geometry with

optics that compensate for the beam walk.

2.5.4.1. Frequency lock

The output frequency of a diode laser depends on the injection current and the

temperature. Lasers with well-defined frequency are needed in many

spectroscopy experiments. But unfortunately after a time, the central frequency

of a diode laser with grating feedback will drift. This drift is the consequence of

fluctuations in temperature and injection current and mechanical

fluctuations. Hence, it is very important to stabilize the laser by locking it to an

external reference in order to reduce that drift.

The external cavity diode laser used in this project is a Toptica DL110 controlled

by a DigiLock 110 Feedback controlyzer.

The DigiLock PID controller was used to control the laser current in order to lock

the frequency of the laser. The laser was locked employing the locking

electronics to one of the fringes of a Thorlabs 1.5 GHz Fabry-Pérot cavity,

which has higher frequency stability. An error signal was created to measure

the deviation in frequency to the reference. The error signal is filtered (with the

locking electronics) and fed back to the laser. This feedback makes the laser

lens lens

output

diffraction

grating

mirror

Page 75: spectral hole-burning and slow light in emerald and ruby

65

frequency follow the reference point. In our case, one side of the fringe was

selected by scanning the laser. The laser was locked to a single optical

frequency when excursions of the laser frequency < 200 MHz were required.

The parameters of the locking electronics (like gain) were optimised for

scanning and locking the laser.

2.6. Monochromator

A 1m Spex 1404 Czerny-Turner monochromator has been used for

luminescence experiments. The main goal of a monochromator is to separate

and transmit a narrow portion of an optical signal from a broader range of

wavelengths available at the input.

In the simplest case a monochromator has two slits (entrance and exit) and a

dispersion element (prism or diffraction grating). In these two elements, the

dependence of the refraction angle (prism) or the reflection angle (grating) on

the radiation wavelength is very important. The dispersion or difraction in the

monochromator can be controlled only if the light is collimated, that is if all the

rays of light are parallel [12].

Monochromators use photoelectric recording of a selected small spectral

interval. An exit slit allows only limited intervals of wavelengths to go through to

the photoelectric detector [7]. The spectrum is scanned by rotating the grating.

This moves the grating normal with respect to the incident and diffracted

beams. The diffracted wavelength is reflected towards the second mirror. The

spectrum is still focused at the exit slit for each wavelength because the light

incident in the grating is collimated.

Page 76: spectral hole-burning and slow light in emerald and ruby

66

Figure 2.15. Czerny-Turner monochromator configuration. Light goes through a slit and is reflected by a collimating mirror to a diffraction grating. A focusing mirror reflects the diffracted light to the exit slit. Adapted from [13].

The spectral resolution of a monochromator is closely related to its spectral

dispersion. The dispersion governs how far apart two wavelengths are. The

resolution determines whether the separation of those two wavelengths can be

resolved. The Rayleigh criterion says that two wavelengths λ1 and λ2 are

resolved when the central maximum of one line falls on a diffraction minimum of

the other (see Figure 2.16). Hence, the spectral resolution can be defined by

Eq. (2.4):

Δλ =

Eq. (2.4)

where is the average wavelength between the two lines (λ1 and λ2), is the

angular dispersion of the system and is the slit width [7, 12].

Entrance slit

Exit slit

Collimating mirror

Focusing mirror

Focal

plane

Page 77: spectral hole-burning and slow light in emerald and ruby

67

Figure 2.16. Monochromator resolution.

2.7. Fabry- Pérot interferometer

A scanning plane-parallel Fabry-Pérot interferometer (Burleigh Instruments RC-

110) was used for fluorescence line narrowing experiments. It consists of two

parallel mirrors. There are multiple reflections between the two spaced mirrored

surfaces.

The condition for producing a maximum in transmission is given by the

condition

Nλ = 2nd cos θ Eq. (2.5)

where λ is the wavelength of the radiation in air, N is an integer, n is the index of

refraction of the air within the interferometer, d is the separation of the reflecting

surfaces and θ is the angle of incidence.

λ1 λ2

Page 78: spectral hole-burning and slow light in emerald and ruby

68

Figure 2.17 shows the transmission as a function of wavelength for a given

angle of incidence. If a particular N0 and λ0 satisfy the above equation, the

interference between the transmitted rays will be constructive and a maximum

of transmission will be produced. Another maximum of transmission will be

observed when other combinations such as N0+1 or λ0 - Δλ satisfy the

mentioned equation. The spacing Δλ between maxima of transmission is the

free spectral range (FSR). The equation for the FSR is given by [7]

Eq. (2.6)

Figure 2.17 shows how the reflectivity of the mirrors affects the transmission. It

follows that the maximum in transmission becomes broader when the reflectivity

is low and when the reflectivity is high the maxima of transmission observed will

be narrower.

Figure 2.17. Transmission as a function of wavelength through a Fabry-Perot interferometer. High reflectivity surfaces show sharper peaks and lower transmission minima than lower reflectivity surfaces . Adapted from [8].

High reflectivity surfaces

Wavelength

Low reflectivity surfaces

Transmission

Page 79: spectral hole-burning and slow light in emerald and ruby

69

The concept of finesse of the interferometer is very important as it measure the

ability of the interferometer to resolve separated spectral lines.

The total finesse of an interferometer can be expresses as the ratio of the free

spectral range to Δ, where Δ is the FWHM (full width half maximum) of the

response function of the system, also known as minimum resolvable bandwidth.

Eq. (2.7)

Two lines separated by Δ are resolvable when the sum of the two individual

lines at the midway point is at most equal to the intensity of one of the original

lines.

The reflectivity finesse Fr can be expresses as:

Eq. (2.8)

where R is the reflectivity of the surfaces. For high reflectivity, the transmission

of maxima of slightly different wavelengths can be resolved because the

transmission maxima are narrow.

2.8. The Laue method

White radiation is transmitted through or reflected from a fixed crystal. Arrays of

spots are formed by the diffracted beams lying on curves on a film. The Bragg

angle is fixed for each set of planes in the crystal. Every single wavelength from

the white radiation that accomplishes the Bragg law is diffracted by each set of

planes. The reflection from planes which belong to one zone are the spots on

one curve. There are two kinds of the Laue method, the transmission and the

back-reflection Laue method.

Page 80: spectral hole-burning and slow light in emerald and ruby

70

Figure 2.18. Transmission to the left and back-reflection Laue method to the right. Adapted

from [14].

In the transmission Laue method, the crystal is placed before a film to record

the beams transmitted through the crystal. In the back-reflection method, the

film is located between the crystal and the x-ray source. The beams recorded

are diffracted in a backward direction.

2.9. Data acquisition and analysis

2.9.1. Data acquisition

The acquisition of the data were undertaken by a PC (500 MHz Intel Pentium III

processor, Intel MMX Technology computer with 128 MB RAM, 13 GB HDD and

Windows 98 OS). Motion controllers, software and National Instruments NI-

488.2 General Purpose Interface Bus (GPIB) hardware devices were used.

Several programs written by Riesen in Microsoft Visual Basic 6.0 for 32-bit

windows were used to enable remote operation of the instruments and

acquisition of processed data. The text format (.txt) in ASCII format was used to

save all the collected data so that standard software could be used to analyse

the data.

Page 81: spectral hole-burning and slow light in emerald and ruby

71

2.9.2. Data analysis

All the data were analysed using a PC (EliteBook 8530p, Intel® Core™ 2 Duo

Processor T9600, 120 GB Hard Disk Drive, OS Windows Vista). Both Microsoft

Office 2007 and Wavemetrics IGOR Pro 6.02A were used to analyse the data.

The sophistication of IGOR enabled us to graph, analyse and work up the data

presented in this thesis. Some fit functions such as Lorentzian or Gaussian

were used in order to measure very narrow linewidths or spectral holes in the

collected spectra. In addition to built-in functions some user-defined functions

were created in order to study temperature dependences and other data

presented in the chapter 3. This characteristic of IGOR to allow the creation of

user-defined functions makes this program very useful in comparison to other

analysis software [15].

In order to analyse lineshapes obtained in this project it was necessary to use

user-defined functions. For instance, the Voigt profile was used to analyse hole-

burning measurements and luminescence data. The Voigt profile is the

convolution of a Lorentzian with a Gaussian.

A Gaussian profile is represented by the equation:

Eq. (2.9)

and a Lorentzian profile is expressed by:

Eq. (2.10)

The convolution of both profiles gives us the equation:

Eq. (2.11)

By using Igor it was possible to obtain a mathematical model to fit our data. The

curves were fitted by choosing the optimal coefficients which make the function

match the data as closely as possible. Voigt fits were used in order to obtain the

Lorentzian contribution at higher temperatures in the presence of the fixed

Gaussian low temperature contribution.

Page 82: spectral hole-burning and slow light in emerald and ruby

72

In order to analyse the data obtained from FLN (intermediate range of

temperatures) the inhomogeneous distribution was taken into account. The

equation used is as follows

Eq. (2.12)

where is the Gaussian distribution of optical centres (inhomogeneous

distribution). For resonant FLN is equal to , being the Lorentzian

distribution of the emission and absorption functions respectively. In this case,

the inhomogeneous linewidth was fixed at 2.5 K and the Lorentzian contribution

at higher temperatures was obtained.

For the analysis of the final data a non-perturbative function was built in order to

obtain a good fit for the resulting graph (see Appendix 1, 2, 3 and 4).

Page 83: spectral hole-burning and slow light in emerald and ruby

73

2.10. References

1. Jirmanus, M.N.; Introduction to laboratory cryogenics; Janis Research

Company, Wilmington (1990).

2. Young, H.D.; Freedman, R.A.; University Physics 9th Edition; Adison-

Wesley Reading (1996).

3. Yariv, A.; Quantum Electronics 3rd Edition; Wiley p. 208–211 (1989).

4. Koechner, W.; Solid-state laser engineering, Springer-Verlag p. 507

(1965).

5. Javan, A., Bennett, W. R.; Herriott, D. R.; Phys. Rev. Lett. 6 (3), 106–110

(1961).

6. http://laser.physics.sunysb.edu/~dli/hnwork.html

7. Demtröder, W.; Laser Spectroscopy, Basic concepts and instrumentation

2nd Edition; Springer (1995).

8. Atkins, P.W.; Phys. Chem. 8th Edition; Oxford University Press (2000).

9. Hall, R. N.; Phys. Rev. Lett. 9 (9), 366 (1962).

10. http://www.rp-photonics.com/ Encyclopaedia of Laser Physics and

Technology.

11. Hayward, B. F.; BSc (Hons) Thesis: UNSW (2004).

12. Domanchin, J. L., Gilchrist, J. R.; Size and spectrum, Spectrometers;

Photonics spectra (2011).

13. http://gratings.newport.com/library/handbook/toc.asp 'Plane gratings and

their mounts'.

Page 84: spectral hole-burning and slow light in emerald and ruby

74

14. http://webpages.iust.ac.ir/panahib/Cryst_files/frame.htm#slide0013.htm. X-

ray application in crystal.

15. www.wavemetrics.com Igor Pro 6.02A manual, Wavemetrics, Inc.

Page 85: spectral hole-burning and slow light in emerald and ruby

75

Chapter 3

Results and discussion

The results obtained for the temperature dependence of the homogeneous

linewidth in emerald and the creation of slow light in ruby by transient spectral

hole-burning are presented and discussed in this chapter.

3. 1. Emerald

3.1.1. Crystal structure and background

Emerald is the green variety of the mineral beryl and its idealized formulae is

Be3Al2(SiO3)6. The bright clear green variety is emerald while a bluish green

type is called aquamarine.

In emerald a small percentage of the aluminium (III) sites are replaced by

chromium(III). Beryl comprises of hexagonal rings formed of six Si-O irregular

tetrahedra. These Si-O distances may vary from 1.54 to 1.68 Å. Each Al is

octahedrally coordinated by group of six oxygen atoms and each Be atom is

coordinated by four oxygen atoms [1, 2].

Alkalis, such as Na, Li, K, Cs and Rb can be found in many samples of natural

beryl. The pink colour in beryl is associated with alkali traces, especially Li.

Also, other impurities that can be found are H2O+ [3].

Beryl crystallises in the hexagonal space group P6/mcc. The hexagonal Si6O18

RRrings form channels parallel to the c axis. As it is seen in the picture below

Page 86: spectral hole-burning and slow light in emerald and ruby

76

tetrahedrally coordinated beryllium and octahedrally coordinated aluminium are

responsible to keep these channels together [3].

Several optical studies show that iron(II) and iron(III) can also substitute for

aluminium ions [4, 5].

Figure 3.1. The crystal structure of beryl. Adapted from [3].

Emerald has been the subject of many spectroscopic studies because of the

presence and interactions of many impurities in its structure [4-9, 10]. Rigby et

al. [9] observed, for the first time, persistent hole-burning in 1992. In those

experiments it was reported that the presence of other impurities such as

titanium may be responsible for persistent holes [9]. Recently, it was found that

persistent spectral hole-burning is present only in natural emerald but not in

laboratory created samples [6]. In addition, Riesen demonstrated that natural

emerald exhibits much broader inhomogeneous broadening in comparison with

laboratory created emerald [11].

Page 87: spectral hole-burning and slow light in emerald and ruby

77

There are two spin-allowed transitions from the 4A2 ground state to 4T2 and 4T1

[8]. The band with a better defined structure is the one from the ground state to

2E leading to very sharp zero-phonon lines at 679.5 and 682.6 nm.

The 4A2 ground state and 2E excited state are split as a consequence of the

trigonal crystal field plus spin-orbit coupling. As a result splittings of

approximately 1.7 cm-1 in the ground state and of 63 cm-1 in the excited state

occur [8].

Emerald created Chatham laboratory with Cr3+ 0.04% (pale green) and

0.0017% (nominally chromium free) per weight were employed in order to

obtain the temperature dependent contribution to the homogeneous linewidth of

the R1(±3/2) line in emerald.

3.1.2. Results and discussion

Polarised transient spectral hole-burning experiments in the R1-line of a

laboratory created emerald (0.0017% Cr(III) per weight) from 2.5 K to 14 K were

obtained. The sample was excited with a diode laser operating at 682.47 nm.

The experiments are reported as a function of low magnetic fields BIIc and

temperature.

A contribution to the homogeneous broadening of less than 1 MHz at 60 K was

measured in this sample. The hole width at 13 K is 46 MHz as is illustrated in

Figure 3.2.

Page 88: spectral hole-burning and slow light in emerald and ruby

78

Figure 3.2. High resolution transient spectral hole-burning in a laboratory created emerald (0.0017% Cr(III) per weight)at different temperatures. The blue lines represent Voigt fits.

When a magnetic field B is applied parallel to the c axis the 2E lower excited

state level ±1/2 (2Ā) and the 4A2 lower ground state level ±3/2 (2Ā) are split by

gexµBBIIc and 3ggsµBBIIc (where gex PPandPP

Pggs

PPare the g-factors of the excited state

and ground state, respectively, B is the magnetic field strength and µB RRis the

Bohr magneton). The trigonal field and spin-orbit coupling splits the ground and

the excited states into two Kramers doublets each. The splitting of the 4A2

ground state and 2PE excited state is 1.79 cm-1 and 63 cm-1, respectively, as

depicted in Figure 3.3.

Page 89: spectral hole-burning and slow light in emerald and ruby

79

Figure 3.3. Schematic energy level diagram for spectral hole-burning experiments in the R1(±3/2)-line in an external magnetic field, BIIc. The laser is in resonance with the -1/2 -3/2 and +1/2 +3/2 transitions yielding side holes at

+3/2 +1/2 and -3/2 -1/2, respectively.

Transient spectral holes of the R1-line in π polarization (E||c) are illustrated as a

function of low magnetic fields in Figure 3.4. The separation between the two

side holes and the main hole is ±(3ggs

-gex

)µBBIIc. Taking the gex

factor as 1.04

and ggs=1.97 as it was established in Ref. [6] we conclude that the scan is

approximately 300 MHz.

+1/2

-1/2

+3/2

-3/2

+1/2

-1/2

+3/2

-3/2

Laser

1.7

9 c

m-1

63

cm

-1

2E Ē

4A2

2Ā (±3/2)

Ē

+(1/2)gexμBB

-(1/2)gexμBB

-(3/2)ggsμBB

+(3/2)ggsμBB

2Ā (±1/2)

Page 90: spectral hole-burning and slow light in emerald and ruby

80

Figure 3. 4. Transient spectral hole spectra in π polarization of the R1(±3/2)-line in a laboratory created emerald (Cr(III) 0.0017% per weight) in low magnetic fields B||c. The inset shows the relative shifts of the side holes as a

function of the magnetic field strength.

Figure 3.5 shows transient spectral hole-burning measurements of the R1(±3/2)-

line between 10 and 30 K in the same sample (0.0017% Cr(III) per weight). As

is illustrated, the resonant hole gets rapidly broader and is hard to measure

above 30 K. For instance, at 27 K a hole width of ~ 500 MHz is observed while

at 2.5 K the observed width is dominated by the instrumental resolution given by

the frequency jitter of the free running diode laser, which is ~ 20 MHz as

measured by a 300 MHz FSR confocal Fabry-Pérot interferometer (Coherent

model 240). From a deconvolution it is found that the homogeneous linewidth at

2.5 K is 4 MHz. This linewidth is much narrower than the 30 MHz linewidth of

Page 91: spectral hole-burning and slow light in emerald and ruby

81

20 ppm ruby in zero field [12]. The linewidth in ruby is dominated by Cr3+

electron spin flip-flops in the environment of a Cr3+ centre [13].

The inset shows three different hole-burning experiments together with their

respective fits. At 2.5 K, the linewidth is defined by a Gaussian fit reflecting the

laser jitter limitation and at higher temperatures Voigt profiles i.e. the

convolution of the Gaussian laser line shape with a Lorentzian were used.

Figure 3.5. Transient spectral hole-burning in the R1(±3/2) line of a laboratory created emerald (0.0017% Cr(III) per weight) at 2.5, 10, 16, 20 and 27 K. The laser wavelength was at 682.4 nm. The inset shows transient spectral holes

(solid lines) with their respective Voigt profiles (dashed lines).

Page 92: spectral hole-burning and slow light in emerald and ruby

82

Luminescence line narrowing (FLN) experiments were measured in two ways. A

Fabry-Pérot interferometer and a monochromator were used for temperatures

up to 80 and 80 to 170 K, respectively.

High resolution FLN experiments conducted with a Fabry-Pérot interferometer

are displayed in Figure 3.6. Conducting the experiments with a Fabry-Pérot

provides us with better results in comparison with a monochromator since the

resolution of the former is of the order of tens of MHz whereas for the

monochromator it is about 20 GHz.

The laser, operating single mode, resonantly excites Cr3+ ions to the 2E(2Ā)

level in the same way as is illustrated in Figure 3.3 (see diagram to the left)

which then luminesce to the doublet 4A2 levels to give the observed spectrum.

Figure 3.6 shows a representative FLN spectrum for three different free spectral

ranges (FSRs) (20, 60 and 120 GHz). These experiments were conducted from

30 up to 80 K. For the spectra with a 20 GHz FSR, good results were obtained

by fitting the linewidths with the convolution of two Lorentzian profiles

(instrumental linewidth and 2 homogeneous linewidth). The instrumental

contribution was found to be approximately 1 GHz. For the spectra measured at

60 and 120 GHz FSR, the linewidth was expected to be determined by a triple

function of the inhomogeneous broadening, emission and absorption lineshape

since these measurements were done at higher temperatures. The

inhomogeneous width was found to be 27 GHz as measured in luminescence at

liquid helium temperature.

Page 93: spectral hole-burning and slow light in emerald and ruby

83

Figure 3.6. Fluorescence line narrowing in a laboratory created emerald (0.04% Cr(III) per weight) for 20, 60 and 120 GHz FSR at 30, 42.5 and 56.5 K, respectively. The dashed lines represent the Voigt fits.

Resonant FLN experiments with a monochromator were undertaken in the

temperature range of 80 to 180 K for the same emerald (0.04% Cr(III) per

weight). A representative spectrum is illustrated in Figure 3.7. Above 90 K the

homogeneous linewidth is much larger than the inhomogeneous width (27 GHz)

and consequently the linewidths can be measured by conventional

luminescence experiments. The splitting of the 4A2 ground state (1.79 cm-1) is

resolved in the R1 line of diluted emerald as is seen in Figure 3.7.

Page 94: spectral hole-burning and slow light in emerald and ruby

84

Figure 3.7. Fluorescence line narrowing (FLN) with a monochromator in a laboratory created emerald (Cr(III) 0.04% per weight) for 2.5, 61, 121 and 161 K. At 2.5 K (red line) the

4A2 ground state splitting can be observed and the R2

line is not populated. The inset to the left represents two spectra at 2.5 and 101 K (solid lines) with their respective Voigt fits (dashed lines). The inset to the right illustrates the spectrum of the single mode laser working at 682.46

nm.

Luminescence experiments were obtained for both a very dilute lab created

emerald (0.0017% Cr(III) per weight) and the more concentrated emerald

sample (0.04% Cr(III) per weight) from the lowest temperature up to 260 K. The

sample was excited with the 532 nm light of a Nd:YAG laser.

Page 95: spectral hole-burning and slow light in emerald and ruby

85

Figure 3.8. Temperature dependence of the non-selectively excited luminescence spectrum of a laboratory created emerald (0.0017% Cr(III) per weight)in the R-lines at 2.5, 120, 180, 220 and 260 K. The inset displays the spectrum at

2.5 and 220 K (solid lines) together with their Gaussian and Voigt fit (dashed lines), respectively.

Figure 3.8 illustrates the temperature dependence of the non-selectively excited

luminescence spectrum of very dilute emerald in the region of the R lines (2E →

4A2 transitions). At low temperature the R lines linewidths are obscured by

inhomogeneous broadening whereas at higher temperatures the homogeneous

linewidth dominates. In particular, the R1 line broadens from an inhomogeneous

linewidth of 27 GHz (2.66 cm-1) to a homogeneous width of 526 GHz (17.5 cm-1)

in the range of 2.5 to 260 K. The shift to the red is ~ 600 GHz (20 cm-1) for the

same temperatures range. In comparison, the ruby R1 linewidth is 360 GHz (12

cm-1) at room temperature and the total red shift between helium and room

temperature is ~90 GHz (3 cm-1). Gaussian profiles are observed at 2.5 K while

Voigt profiles are observed for higher temperatures.

Page 96: spectral hole-burning and slow light in emerald and ruby

86

The linewidth data of the four experiments are summarized in Figure 3.9 where

the temperature dependent contribution to the R1(±3/2) linewidth between 5.6 K

and 260 K is illustrated. The different data sets are represented with different

colours in the graph depending on the experiment conducted. The data are well

described by the non-perturbative expression for developed by Hsu

and Skinner [14] in Eq. (3.1)

Eq. (3.1)

where and are the direct and Raman processes respectively

expressed by Eqs. (1.8.b) and (1.9).

Figure 3.9. Temperature dependent contribution ΔГhom to the homogeneous linewidth Гhom of the R1(±3/2) line (circles) for emerald. The solid circles show the data for a very diluted lab created emerald (0.0017% Cr(III) per weight) and the open circles represent the data for a more concentrated emerald. Trace 1: non-perturbative

approach using Eq. (3.1). Trace 2: non-perturbative approach using Eq. (3.1) plus a phonon sideband at the lower temperature range. Trace 3: two-phonon Raman process using Eq. (1.8.b). Trace 4: direct process using Eq. (1.9). No

side phonon is considered in this fit. Trace 5: two-phonon Raman process but with 5 modes only using Eq. (3.7).

ΔГ h

om

/ M

Hz

Page 97: spectral hole-burning and slow light in emerald and ruby

87

A good fit at high temperature is obtained by the parameters W=-0.43, Г0(R2)=

12500 MHz and TD= 845 K. The Debye temperature (TD) is the temperature of a

crystal's highest mode of vibration. The Debye temperature is given by

Eq. (3.2)

where is Planck's constant, is Boltzmann's constant and is the Debye

frequency. According to Figure 3.10 our highest mode of vibration is at 711 nm

(589 cm-1 from the R1). This results in a Debye temperature of 845 K. We note

here that our Debye temperature is at variance from the 580 K reported in [15].

The value for Г0(R2) is calculated by assuming that the two-phonon Raman

process is the same for R1 and R2 at 80 K and also the inhomogeneous

broadening. The luminescence linewidth observed for R1 is

Eq. (3.3)

and the linewidth observed for R2 is

Eq. (3.4)

The difference between R1 and R2 linewidths yields

Eq. (3.5)

From values of and in luminescence at 80 K we calculate that is

15500 MHz, which differs from 12500 MHz used in the fit illustrated in Figure

3.9. We believe that this difference might be due to the Voigt fit at this particular

temperature, which is not very good since our R1 line in luminescence starts to

be asymmetrical at high temperatures (see Figure 3.8).

However, in contrast to ruby, the data was not well described for the lowest

temperature range. This mismatch comes possibly from the simplicity of the

Debye approximation. Spin-lattice relaxation cannot be taken into account for

the variation of measured and calculated linewidths below 8 K because it is very

Page 98: spectral hole-burning and slow light in emerald and ruby

88

slow [16] and according to Eq. (3.6) has little influence in the contribution to the

homogeneous broadening.

Eq. (3.6)

It is possible that there is a phonon sideband at low energy. After observing the

asymmetrical lineshape of the spectra in luminescence in Figure 3.8 and the

spectra of Figure 3.10 it appears that a low energy phonon is located at 683.05

nm (15 cm-1 separated from the peak of the R1 line). A good fit is obtained when

this is taken into consideration (see Trace 2 in Figure 3.9). It is also possible

that spectral diffusion causes extra broadening at this temperature.

Inspecting the summary of linewidth data provided in Figure 3.9 another

discrepancy was found at higher temperatures. It can be observed that the

empty circles (FLN with monochromator and luminescence in more

concentrated emerald) are not fitted by the non-perturbation function. Attempts

were made to fit these points by using pseudo-local phonons. The spectra in

Figure 3.10 exhibits peaks at 691, 697, 702, 707 and 711 nm (184, 324, 408,

509 and 589 cm-1) in emerald. The two-Raman process with 5 modes described

the data as represented by the trace 5 in Figure 3.9. The formulae used in this

case was

Eq. (3.7)

where are the coupling constants, are the centre frequencies of each

phonon and i are all the local modes. We believe that these points do not match

with the rest of the data because these experiments were conducted in a more

concentrated sample where energy transfer occurs. The solid circles in Figure

3.9 represent the same experiments in a more dilute emerald (0.0017% Cr(III)

per weight). Good agreement with the data was obtained in this case. In the

diluted emerald no energy transfer happens and as a consequence narrower

linewidths were obtained.

The 2E → 4T2 energy gap in emerald is about 400 cm-1 [16] compared to 2300

cm-1 in ruby. Excitation into any higher state relaxes quickly to 2E (4T1 and 4T2

Page 99: spectral hole-burning and slow light in emerald and ruby

89

states relax by non-radiative transitions to the 2E excited state) and 2E → 4A2

line emission is observed. As 2E → 4A2 transition is spin-forbidden, the 2E level

has a long lifetime (~1.7 ms) [15, 17, 18]. When increasing the temperature, the

4T2 population grows from 2E and the 4T2 → 4A2 broad band emission line is

observed. The separation between 4T2 and 2E states in these two crystals

results in a large difference between the lifetimes of the metastable levels at

room temperature (~6 10-5 s for emerald and ~3 10-3 s for ruby) [17]. The

lifetime of 4T2 is found to be 10-12 s. We assume that the 2T1 level between 2E

and 4T2 has little contribution to the homogeneous broadening [17, 18].

Figure 3.10. Vibrational side band of the R1 line in diluted emerald. A side phonon is located at 683.05 nm. Other 5 phonons can be observed next to the R1 line.

In conclusion, the low temperature range up to 40 K seems to be dominated by

the direct one-phonon process between Ē and 2Ā of the excited state 2E plus a

two-phonon Raman process by a low energy phonon whereas at higher

temperatures the two-phonon Raman process plays the most important role.

683.05 nm

Page 100: spectral hole-burning and slow light in emerald and ruby

90

3. 2. Slow light in ruby by transient spectral hole-burning

3.2.1. Crystal structure and orientation

Ruby is a very important material in the development of the optical

spectroscopy of impurity ions in insulators and wide bandgap semiconductors.

Ruby was also the first material to display laser action in the visible spectrum as

demonstrated in 1960 by Maiman [19]. The laser action within a ruby rod was

induced by a flash tube. Coherent laser emission took place after the utilisation

of a short pulse of light to pump the ruby rod. The ruby laser is of significant

importance since is the basis of current solid state technology, such as the

Nd:YAG laser.

Ruby is the pink-to-red variety of the mineral corundum and its chemical formula

is Al2O3:Cr(III). It has a trigonal crystalline structure crystallising in the space

group D63d - R3c. In ruby Cr3+ replaces Al3+, which are octahedrally coordinated

by oxygen atoms [20]. Ruby is the second hardest material that occurs

naturally.

There are several crystal growing techniques to grow ruby in the laboratory,

such as the Czochralski [21], floating zone [22], flame fusion [23] and

Bagdasarov methods [24]. The crystal we used was grown by Bagdasarov

method. In the Bagdasarov method the material is placed in a crucible and is

melted by moving the crucible through the heating zone. This method allows

repeated crystallization when other chemical impurities of the raw material are

required.

Page 101: spectral hole-burning and slow light in emerald and ruby

91

Figure 3.11. Crystal structure of corundum.

In order to cut the crystals perpendicular to the c-axis the orientation of the

crystal was determined by the Laue method. Figure 3.12 shows Laue

photographs in 180° backscattering geometry of the crystallographic plane

perpendicular to the c axis for 20 ppm Verneuil ruby and 130 ppm Bagdasarov

ruby.

Figure 3.12. Laue photograph of 20 ppm ruby to the left and 130 ppm (Bagdasarov) to the right of the plane perpendicular to c.

70

06

50

60

05

50

50

04

50

40

03

50

30

02

50

20

01

50

10

05

00

110010501000950900850800750700650600550500450400350300250200150100500

70

06

50

60

05

50

50

04

50

40

03

50

30

02

50

20

01

50

10

05

00

110010501000950900850800750700650600550500450400350300250200150100500

Page 102: spectral hole-burning and slow light in emerald and ruby

92

3.2.2. Results and discussion

3.2.2.1. Experiment

Minimal inhomogeneous broadening is advantageous for the generation of slow

light by spectral hole-burning in ruby as it allows high optical densities with low

chromium (III) concentrations. Therefore, the inhomogeneous broadening of

ruby crystals was measured with different chromium (III) concentrations grown

by a range of methods (Verneuil, Czochralski, Bagdasarov, hydrothermal). It

appeared that gross macroscopic strain broadening affects some crystals,

especially the flame fusion (Verneuil) grown crystals. The best results were

obtained for Bagdasarov ruby. An inhomogeneous linewidth of 2 GHz (FWHM)

was obtained as is illustrated in Figure 3.13. The spectrum was measured in

fluorescence-excitation mode by observing R1 - luminescence whilst scanning

the laser over 30 GHz. The separation between the R1 splitting corresponds to

11.49 GHz.

Figure 3.13. Resonance luminescence excitation of the R1 line in Bagdasarov ruby. The various isotopic lines are also labelled.

Page 103: spectral hole-burning and slow light in emerald and ruby

93

Different linewidths ranging from 2 to 10 GHz were obtained depending on the

position in the crystal. For the slow light experiments, a 130 ppm Bagdasarov

ruby of 2.3 mm thickness was used. The set up for the slow light experiment is

schematically illustrated in Figure 2.10. In particular, a 750 μs burn pulse was

applied followed by a weak 50 ns probe pulse with 10 μs delay (i.e. at 760 μs).

Figure 3.14 shows transmission spectra of a 2.3 mm Bagdasarov ruby (130

ppm) along with its change in absorbance ΔA upon hole-burning. A magnetic

field of 9 mT parallel to the c axis was applied. When an external magnetic field

is applied, the ground state doublets ±1/2(Ē) and ±3/2(2Ā) and the 2E excited

state level Ē are split. Therefore, upon hole-burning in the 2Ā(2E) Ē(4A2) line

two side holes are expected, separated from the resonant hole by ±(3ggs-

gex)µBBIIcR R. Considering that ggs=1.98 and gex=2.44, we obtain side holes

separated by ±445 MHz from the resonant hole. The transitions R1(±3/2) and

R1(±1/2) are resolved for the various stable isotopes of chromium (III).

Figure 3.14. a) Transmission (red line) spectrum in α-polarisation of a 2.3 mm Bagdasarov ruby (130 ppm Cr(III)) in BIIc=9 mT and change in absorbance ΔA (blue line). The black line shows the spectrum without hole-burning. b)

Same as in a) but over an extended frequency range. Chromium (III) isotopes that are responsible for the various R1(±3/2) and R1(±1/2) lines are denoted.

a) b)

Page 104: spectral hole-burning and slow light in emerald and ruby

94

Naturally, there is a change in absorbance (A) after a burn pulse. It is obvious

that cross-relaxation happens between the 4A2 levels of all the isotopes i.e.

cross-relaxation between resonant and non-resonant ions [25].

The decay of the spectral hole burnt by the 750 μs burn pulse was studied by

the application of a 50 kHz burst of 1 μs probe pulses as is illustrated in Figure

3.15.

Figure 3.15. 50 kHz burst of 1 μs probe pulses after a 750 μs burn pulse.

It appears that the initial hole decay is faster in the case when an external

magnetic field BIIc is applied. This is caused by a change in spin-lattice

relaxation in the excited state levels +1/2 and -1/2.

Figure 3.16 summarizes the decay data. It follows that the hole in zero field

decays approximately with the lifetime of the excited state. The hole in BIIc=9

mT initially decays on a timescale of 100 μs due to spin-lattice relaxation

between the +1/2 and -1/2 split levels.

Page 105: spectral hole-burning and slow light in emerald and ruby

95

Figure 3.16. Hole depth decay when a magnetic field BIIc=9 mT is present (blue markers) and in zero field (red markers). For the hole depth decay in a magnetic field the fit parameters are 1/τ1=0.29 ms

-1 and 1/τ1=10.05 ms

-1

whilst for zero field 1/τ=0.29 ms-1

.

To quantify the slow light experiments we have also measured the hole shape

after the 750 μs burn pulse. This is shown in Figure 3.17. The observed hole

width of 50 MHz is limited by laser jitter and spectral diffusion.

Figure 3.17. Transient hole-burning in the presence of a magnetic field of BIIc=9 mT in 130 ppm Bagdasarov ruby. The blue line is a Gaussian fit. The hole width is 50 MHz. The hole is measured with a 1.5 ms delay after a 750 μs

burn pulse.

Page 106: spectral hole-burning and slow light in emerald and ruby

96

Figure 3.18 presents the time delay obtained for the probe pulse after hole-

burning with respect to the probe pulse without hole-burning. A delay of 10.8 ns

is observed for the probe that follows a burn pulse with minimal distortion.

Figure 3.18. Delay of the probe pulse 10 μs after a 750 μs burn pulse (red line) in comparison to the normalized (green line) and non-normalized probe pulse without hole-burning (blue line). Experimental parameters: Гinh = 1.9

GHz, FWHM (pulse duration of input pulse) = 59 ns, A (absorbance before burning) = 1.53, Spectral hole width = 44 MHz, ΔA (change in absorbance) = 1.09.

Page 107: spectral hole-burning and slow light in emerald and ruby

97

3.2.2.2. Simulations

The linear spectral filter theory as outlined in chapter 1 (see section 1.5.2.3)

was applied and a Mathematical code of Professor Rebane and translated by

Riesen to Matlab (see Appendix 5) was used.

As seen in Figure 3.19.a larger delays are obtained when the absorbance is

higher. Also, when the absorbance is increased the pulse width after hole

burning gets broader with respect to the pulse width without burn pulse as seen

in Figure 3.19.b. For these simulations, we used the following input values: Гinh

= 2 GHz, pulse width = 50 ns and hole width = 40 MHz.

Figure 3.19. a) Delay of a Gaussian probe pulse versus ΔA/A for different absorbances A=0.5, 1, 1.5 and 2. Hole width = 40, pulse width = 50 ns, Гinh = 2 GHz. b) Ratio of the widths of delayed and non-delayed pulse versus ΔA/A

for A=0.5 and 1. Hole width = 40, pulse width = 50 ns, Гinh = 2 GHz.

Another simulation was run to calculate the delay time of the probe pulse as a

function of the hole width as is depicted in Figure 3.20. As expected the delay of

the probe pulse decreases with increasing hole width. Calculation were

undertaken for 50 and 100 ns pulse duration of the probe pulse and different

hole depths ΔA. Larger delays are obtained for narrower hole widths and deep

a) b)

Page 108: spectral hole-burning and slow light in emerald and ruby

98

holes ΔA. For broader hole widths there is no difference in delay between

longer and shorter pulses.

Figure 3.20. Delay versus hole width for different changes in absorbance and pulse duration. Pulse width = 50 and 100 ns, A=1.5, Гinh = 2 GHz.

When the hole gets very narrow, probe pulses of 50 ns start to be distorted, i.e.

they lose their Gaussian lineshape as is illustrated in Figure 3.21.

Page 109: spectral hole-burning and slow light in emerald and ruby

99

Figure 3.21. Delay of the probe pulse (red line) after hole-burning in comparison with the normalized (blue line) and non-normalized non-delayed pulse (green line). The dashed lines are the Gaussian fits for each pulse. Гinh = 2 GHz, A

= 1.5, ΔA=1, Pulse width = 50 ns, hole width = 10 MHz.

A calculation to approximate the hole width to the delay of 10.8 ns observed in

the experiment of Figure 3.18 is presented in Figure 3.22. A value of 45 MHz

corresponds to 10.8 ns. By comparing the hole width (50 MHz) experimentally

observed with the hole width calculated theoretically, we deduce that the

broadening observed in experiments is due to laser jitter and/or spectral

diffusion.

Page 110: spectral hole-burning and slow light in emerald and ruby

100

Figure 3.22.Hole width versus delay of the probe pulse. A = 1.53, Гinh = 1.9 GHz, ΔA = 1.09, pulse width = 59 ns.

3.3. References

1. Antsiferov, V. V.; Free running emerald laser; in Technical Physics, 45,

1085-1087 (2000).

2. Morosin, B.; Structure and thermal expansion of Beryl; in Acta Cryst., 28,

1899-1903 (1972).

3. Deer Howie Zussman, Rock forming minerals, Volume 1 (Ortho- and ring

silicates); Longmans (1962).

4. Khaibullin, R.I.; Lopatin, O.N.; Vagizov, F. G.; Bazarov, V. V.; Bakhtin, A.I.;

Aktas, B.; Nucl. Instr. And Meth. In Phys. Res. B, 206, 207 (2003).

5. Viana, R. R.; Jordt-Evangelista, H.; Magela de Costa, G; Phys. Chem.

Minerals, 29, 668 (2002).

6. Riesen, H.; Chem. Phys. Lett., 382, 578 (2003).

7. Wood, D. L.; Nassau, K.; Am. Mineral, 53, 777 (1968).

Page 111: spectral hole-burning and slow light in emerald and ruby

101

8. Wood, D. L.; Ferguson, J.; Knox, K.; Dillon, J. F.; J. Chem. Phys. 39, 890

(1963).

9. Rigby, N. E.; Manson, N. B.; Dubicki, L.; Troup, G. J.; Hutton, D. R.; J. Opt.

Soc. Am. B, 9, 775 (1992).

10. Wood, D. L.; J. Chem. Phys. 42, 3404, 10 (1965).

11. Riesen, H.; Journal of Phys. Chem., 115, 5364-5370 (2011).

12. Riesen, H.; Szabo, A.; Chem. Phys. Lett. 484, 181 (2010).

13. Riesen, H.; Hayward, B. F.; Szabo, A.; J. Lumin. 127, 655 (2007).

14. Hsu, D.; Skinner, J. L.; J. Chem. Phys. 83, 2107 (1985).

15. Hasan, Z.; Keany, S. T.; Manson, N. B.; J. Phys. C: Solid State Phys. 19,

6381-6387 (1986).

16. Hayward, B.; Riesen, H.; Phys. Chem. 7, 2579 (2005).

17. Kisliuk, P.; Moore, C. A.; Phys. Rev. 160, 307 (1967).

18. Quarles, G. J.; Suchocki, A.; Powell, R. C; Phys. Rev. B 38, 14 (1988).

19. Maiman, T. H.; Nature 187, 493 (1960).

20. Van der Ziel, J. P.; Phys. Rev. 9, 2846 (1974).

21. Bobert, C.; Linares, A.; J. Phys. Chem. 26, 1817 (1965).

22. Saito, M.; J. Cryst. Growth 74, 385 (1986).

23. Duker, G.W.; Kellington, C.M.; Katzmann, M.; Atwood, J.G.; Appl. Opt. 4,

109 (1965).

24. http://www.bagdasarovcrystals.com/v1/index.php?id=4&sid=14

25. Szabo, A.; Phys. Rev. B 11, 4512 (1975).

Page 112: spectral hole-burning and slow light in emerald and ruby

102

Chapter 4

Conclusions

Spectral hole-burning, fluorescence line narrowing and luminescence

experiments were conducted to establish the temperature dependence between

2.5 and 260 K of the R1 linewidth in emerald (Be3Al2(SiO3)6:Cr(III)). The

contribution to the homogeneous broadening is only ~70 kHz at 6 K but a

massive ~400 GHz at 260 K. Up to 80 K the direct one-phonon process

between the split levels of the 2E excited state and the presence of a phonon

sideband at low energy seems to dominate the linewidth. Above this

temperature, the two-phonon Raman process becomes important. The non-

perturbative theory explains the two-phonon Raman process very well but at

low temperature, in addition to the direct process, a low energy phonon is

necessary to explain the data.

In this thesis, slow light has been generated by transient spectral hole-burning

for the first time. A delay of 10.8 ns was obtained for a 130 ppm Bagdasarov

ruby of 2.3 mm thickness. The experimental investigation has been extended to

simulations by a mathematical code written in Mathematica by Aleksander

Rebane and translated to Matlab by Riesen. While the delay that was observed

in ruby using transient hole-burning is impressive, the hole width cannot be too

narrow. When this happens, short probe pulses are distorted and they are not

Gaussian. Matching the probe pulse width and the hole width is necessary to

observe slow light. Due to the wide linewidths at room temperature in emerald,

it seems that it is not a suitable candidate for the creation of slow light in this

crystal at high temperatures. However, f-f transitions are usually much less

susceptible to temperature dependent homogeneous broadening and hence

these systems would result in better candidates for slow light generation by

transient spectral hole-burning at higher temperatures. The present

demonstration of slow light by transient hole-burning has opened a wide field of

interesting slow light experiments in d-d and f-f transitions in solidds.

Page 113: spectral hole-burning and slow light in emerald and ruby

103

Appendix 1

Voigt function used for hole-burning

Function voigt(w,x) : FitFunc

Wave w

Variable x

//CurveFitDialog/ w[0] = IGORwidth

//CurveFitDialog/ w[1] = gamma

//CurveFitDialog/ w[2] = scale

variable m, product, xp, integral

m=0

do

xp=-100+m*0.08

product=(exp(-xp^2/(w[0]^2))*(w[1]/((x-xp)^2+(w[1]/2)^2)))

integral=integral+product*0.08

m+=1

while(m<2500)

return integral*w[2]

End

Page 114: spectral hole-burning and slow light in emerald and ruby

104

Appendix 2

Function used for FLN monochromator and 60 and 120 Ghz scan lengths with

Fabry-Perot interferometer

Function FLN(w,x) : FitFunc

Wave w

Variable x

//CurveFitDialog/ w[0] = scale

//CurveFitDialog/ w[1] = INHwidth

//CurveFitDialog/ w[2] = gamma

Variable m, product,xp,integral

m=0

integral=0

do

xp=-0.5+m*0.0005

product=1/((xp-x)^2+(w[2]/2)^2)*exp(-xp^2/w[1]^2)*1/((xp)^2+(w[2]/2)^2)

integral=integral+product*0.0005

m+=1

while (m<2000)

return integral*w[0]

End

Page 115: spectral hole-burning and slow light in emerald and ruby

105

Appendix 3

Voigt function used for luminescence (RRR1RR and RRR2RR lines).

Function voigt(w,x) : FitFunc

Wave w

Variable x

//CurveFitDialog/ w[0] = IGORwidth

//CurveFitDialog/ w[1] = gamma

//CurveFitDialog/ w[2] = x0

//CurveFitDialog/ w[3] = gamma2

//CurveFitDialog/ w[4] = x02

//CurveFitDialog/ w[5] = scale

//CurveFitDialog/ w[6] = ratio

Variable m, product, xp, integral

m=0

integral=0

do

m+=1

xp=-10+m*0.005

product=exp(-(xp)^2/w[0]^2)*(1/(((x-w[2])-xp)^2+(w[1]/2)^2)+w[6]/(((x-w[4])-

xp)^2+(w[3]/2)^2))

integral=integral+product*0.005

while (m<4001)

Page 116: spectral hole-burning and slow light in emerald and ruby

106

return integral*w[5]

end

Page 117: spectral hole-burning and slow light in emerald and ruby

107

Appendix 4

Non-perturbation function used for the final data analysis.

Function nonperturbation(w,T) : FitFunc

Wave w

Variable T

//CurveFitDialog/ w[0] = W

//CurveFitDialog/ w[1] = beta

//CurveFitDialog/ w[2] = gamma0

//CurveFitDialog/ w[3] = beta1

//CurveFitDialog/ w[4] = DT

//CurveFitDialog/ w[5] = a1

//CurveFitDialog/ w[6] = w1

variable f1,f2,f3,f4,f5,f6,f7,x, f8, f9, f10,f11

variable i, lim

variable pi

i=1

lim=1000

pi=3.1415

f1=1.38e-23*w[4]*2*pi/(6.62e-34*(4*pi^2))

f2=w[4]/T

f7=0

do

Page 118: spectral hole-burning and slow light in emerald and ruby

108

x=i/lim+0

f3=ln((1-x)/(1+x))

f4=1/((1+w[0]*(1+3*x^2+3/2*x^3*f3))^2+(w[0]^2)*9*((pi^2)/4)*(x^6))

f5=1+(9*(pi^2)*(w[0]^2)*(x^6)*exp(x*f2)/(exp(x*f2)-1)^2)*f4

f6=ln(f5)

f7=f7+f1*f6/lim

i+=1

while (i<lim)

f8=(w[6])/(T*0.695)

f9=exp(f8)-1

f10=1/f9

f11=w[5]*f10*(f10+1)

return f7/1e6+w[1]*(1/(exp(63/(0.695*T))-1))+w[3]*(1/(exp(6/(0.695*T))-

1))+w[2]+f11

End

Page 119: spectral hole-burning and slow light in emerald and ruby

109

Appendix 5

Matlab code for slow light experiments. % pulse reshaping/slow light calculation following Aleks Rebane's code in

% Mahematica; translated to Matlab by Hans Riesen 2012

%Note: Matlab was written by an engineer and thus has reverse sign

%convention in Fourier transformation i.e. FFT in Matlab is DFFT in

%Mathematica nd vice versa; note that normalisation factors have to b

%adjusted

%define parametres

clear

va=4.32*10^5; % centre of inhomogeneous band/GHz

a1=2; %width of inhomogeneous distribution/GHz

a21=1.5; %absorbance before burning in log10 base

deltaA=0.25; %change in absorbance at spectral hole in log10 base

tsp=10^(-(a21-deltaA)); %transmission at hole

a2=log(10)*a21; %absorbance at max of band before burning in ln base

Dt1=100.0; %pulse duration of input pulse FWHM/ns

a22=exp(-a2);

%following line is from Aleks; I think the peak transmission needs to be

%corrected to: 1. absorbance on log10 base -log10(a22); 2 corrected for

%hole: -0.4; convert back to transmission 10^diff

a3=tsp-a22; %peak transmission of spectral hole (10^-0.4=0.38, exp(-2)=0.135;

%mine: a222=a2+log(10^-0.4)

%mine: a3=exp(-a222)

a4=0.10; %Width of spectral hole

%normalized inhomogeneous absorption line shape

v=4.31980*10^5:0.001:4.32020*10^5;

func01=exp(-log(2)*((v-va)./(a1./2)).^2);

figure (1)

plot(v,func01);

%Intensity transmission without spectral hole

Page 120: spectral hole-burning and slow light in emerald and ruby

110

figure(2)

func02=exp(-a2.*func01);

plot(v,func02);

%temporal pulse shape * Gaussian amplitude profile

t=-100:0.001:100; %nanoseconds

%%%t=t*10^(-9);

func03=exp(-(log(2)./2)*(t./(Dt1./2)).^2);

figure(3)

plot(t,func03.^2);

%evaluate Fourier transform spectrum of the incident pulse

%func03a=func03.*exp(-i*2*t*va*3.1415);

%F=fourier(func03a);

figure(4)

v1=-0.1:0.001:0.1;

denom=sqrt(1/Dt1.^2)*sqrt(2*log(4));

func04a=((1/denom.*exp(-(3.1415)^2*Dt1.^2*((v1).^2)/log(4))).^2);

plot(v1,func04a)

denom=sqrt(1/Dt1^2)*sqrt(2*log(4));

func04=1/denom.*exp(-(3.1415)^2.*Dt1.^2*((v-va).^2)/log(4));

func05=a3.*exp(-log(2).*((v-va)./(a4./2)).^2);

func06=(func02+func05);

figure(5)

plot(v,func06);

%2.Calculate causal complex response function

%2.1. Numerical list representation of the amplitude transmission frequency

%range 40 GHz

list02=func02;

list06=func06;

list02a=1/sqrt(40000)*fft(log(list02));

list02b=list02a;

%cut off negative time part

for k=20000:40001;

Page 121: spectral hole-burning and slow light in emerald and ruby

111

list02b(k)=0;

end

%transform back to frequency

list02c=1/sqrt(40000)*fft(list02b);

list02d=(sqrt(list02).*exp(j*2*imag(list02c)));

%combines ampl;itude with causal phase

for k99=1:1000

list02c(k99);

end

%2.4 with spectral hole

list06a=1/sqrt(40000)*fft(log(list06));

list06b=list06a;

%cut off negative time part

for k=20000:40001;

list06b(k)=0;

end

list06c=1/sqrt(40000)*fft(list06b);

list06d=sqrt(list06).*exp(j*2*imag(list06c));

%combine amplitude with causal phase

%calculate transmitted phase

list04=func04;

list07a=sqrt(40000)*ifft(list04.*list02d);

list07aa=list07a; %make copy

for k=1:20000;

list08a(k)=list07a(20000+k);

end

for k=20001:40001;

list08a(k)=list07aa(k-20000);

end

list07b=sqrt(40000)*ifft(list04.*list06d);

list07bb=list07b; %make copy

Page 122: spectral hole-burning and slow light in emerald and ruby

112

for k=1:20000;

list08b(k)=list07b(20000+k);

end

for k=20001:40001;

list08b(k)=list07bb(k-20000);

end

%

for i1=16000:24000;

list09ax(i1)=-500+i1/40;

list09ay(i1)=abs(list08a(i1).^2);

end

for i1=16000:24000;

list09bx(i1)=-500+i1/40;

list09by(i1)=abs((list08b(i1).^2));

end

figure(6)

plot(list09ax,list09ay,list09bx,list09by);

i1 = 16000:1:24000;

y = [list09ax(i1); list09ay(i1); list09bx(i1); list09by(i1)];

fid = fopen('list09.txt', 'wt');

fprintf(fid, '%12.8f %12.8f %12.8f %12.8f\n', y);

fclose(fid)

%figure(7)

%plot(list09bx,list09by);


Recommended