Survey
* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project
* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project
Quantum control of laser induced dynamics of diatomic molecular ions using shaped intense ultrafast pulses A thesis presented upon application for admission to the degree of Doctor of Philosophy in the Faculty of Engineering and Physical Sciences by Leigh Graham MSci (Hons) 2008 School of Mathematics and Physics Queen’s University Belfast Northern Ireland 17/09/2013 Abstract The beauty of ultrafast science lies inherently in the ability to induce and image dynamics on a timescale comparable to the fastest nuclear motion. In recent years, a plethora of rich and fascinating phenomena involving the interaction of diatomic molecules with intense femtosecond laser pulse has been unveiled. Such research is motivated by the ambition to understand and optically drive chemical reactions to the highest degree of specificity. In this work, the strategy employed toward achieving this goal relies on the interaction of Hydrogenic ions and analytically shaped and well characterized pulses. The ability to manipulate photodissociation dynamics using the instantaneous 00 000 frequency and temporal profile of pulses shaped with quadratic (ϕ ) and cubic (ϕ ) spectral phase functions was studied. A three-dimensional (3D) momentum imaging technique was used to measure the kinetic energy release (KER) and angular distribution of the dissociation fragments. A significant enhancement in the dissociation probability of non-resonant transitions from the low lying vibrational levels using the sign and magnitude of the applied phase function as a control tool was demonstrated. Furthermore, the tractability of Hydrogenic ions means a mechanistic explanation for these observations can be theoretically determined. Investigating the behavior of ions more complex than H+ 2 in strong laser fields can present many theoretical and experimental challenges. Laser-induced fragmentation of CD+ was explored using the 3D momentum imaging technique in the longitudinal field imaging mode. The high mass ratio (12:2) hinders the simultaneous measurement of the two constituents, at all angles and kinetic energies. Alternatively, the recently developed longitudinal and transverse field imaging technique was used to perform a piecewise dissociation measurement. This allowed the branching ratio of the dissociation channels to be obtained. Furthermore, the fragmentation channels of CD+ were identified and studied as a function of laser intensity and wavelength. Contents Abstract 1 1 Diatomic Molecular Structure and its Interaction with Light 2 1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3 1.2 Atomic Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3 1.3 Diatomic Molecules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 1.3.1 Molecular Orbital Theory . . . . . . . . . . . . . . . . . . . . . . 5 1.3.2 Molecular Kinematics and Potential Energy Curves . . . . . . . . 7 1.3.3 Molecular Term Symbols . . . . . . . . . . . . . . . . . . . . . . 10 1.3.4 Transition Dipole Moment and Selection Rules . . . . . . . . . . 12 1.4 1.5 Photodissociation Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . 14 1.4.1 Dissociation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16 1.4.2 Floquet Formalism . . . . . . . . . . . . . . . . . . . . . . . . . . 17 1.4.3 Bond Softening . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18 1.4.4 Above Threshold Dissociation . . . . . . . . . . . . . . . . . . . . 19 1.4.5 Vibrational Trapping . . . . . . . . . . . . . . . . . . . . . . . . . 21 1.4.6 Below Threshold Dissociation . . . . . . . . . . . . . . . . . . . . 22 1.4.7 Angular distribution of fragment ions . . . . . . . . . . . . . . . 23 Ionization Processes in Intense Laser Fields . . . . . . . . . . . . . . . . 25 1.5.1 Non Sequential Ionization . . . . . . . . . . . . . . . . . . . . . . 29 1.5.2 Charge Asymmetric Dissociation . . . . . . . . . . . . . . . . . . 32 CONTENTS 3 1.5.3 Coulomb Explosion and Multielectron Dissociative Ionization . . 34 1.6 Molecular Alignment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40 1.6.1 Adiabatic Alignment . . . . . . . . . . . . . . . . . . . . . . . . . 41 1.6.2 Non-adiabatic Alignment . . . . . . . . . . . . . . . . . . . . . . 41 2 Ultrafast Measurement Techniques 2.1 2.2 2.3 2.4 42 Experimental Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43 2.1.1 Neutral vs. Ionic Targets . . . . . . . . . . . . . . . . . . . . . . 43 2.1.2 Pulse Intensity Effects . . . . . . . . . . . . . . . . . . . . . . . . 45 2.1.3 Temporal Duration . . . . . . . . . . . . . . . . . . . . . . . . . . 47 Experimental Imaging Techniques . . . . . . . . . . . . . . . . . . . . . 48 2.2.1 Time of Flight . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48 2.2.2 Covariance Mapping . . . . . . . . . . . . . . . . . . . . . . . . . 50 2.2.3 COLTRIMS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51 2.2.4 Velocity Map Imaging . . . . . . . . . . . . . . . . . . . . . . . . 52 2.2.5 Pump-Probe Schemes . . . . . . . . . . . . . . . . . . . . . . . . 53 Laser Pulse Shaping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54 2.3.1 Time and Frequency Domains . . . . . . . . . . . . . . . . . . . . 54 2.3.2 Passive Optical devices; Materials, Prisms and Gratings . . . . . 57 2.3.3 Acousto-Optic programmable Dispersive Filters (AOPDF) . . . . 58 2.3.4 Masks in the Fourier Plane . . . . . . . . . . . . . . . . . . . . . 58 Coherent Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60 3 3D Momentum Imaging Technique 63 3.1 Neilson Ion Source . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65 3.2 Vacumn System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66 3.2.1 Scroll Pumps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67 3.2.2 Turbomolecular Pumps . . . . . . . . . . . . . . . . . . . . . . . 67 CONTENTS 4 3.2.3 Hot-Filament Ionization Gauges . . . . . . . . . . . . . . . . . . 68 3.3 Wien Velocity Filter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69 3.4 Ion Beam Manipulation . . . . . . . . . . . . . . . . . . . . . . . . . . . 71 3.4.1 Einzel Lens and Deflectors . . . . . . . . . . . . . . . . . . . . . . 72 3.5 Quadrupole Triplet Focusing Lens . . . . . . . . . . . . . . . . . . . . . 74 3.6 Spectrometer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76 3.7 Ion Beam Alignment Protocol . . . . . . . . . . . . . . . . . . . . . . . . 78 3.8 3.7.1 Collimation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78 3.7.2 Ion Beam Imaging . . . . . . . . . . . . . . . . . . . . . . . . . . 80 3.7.3 Ion Beam Current Measurement . . . . . . . . . . . . . . . . . . 81 Hexanode Delay-Line Detector . . . . . . . . . . . . . . . . . . . . . . . 82 3.8.1 3.9 Hexanode Delay-Line Detector Calibration . . . . . . . . . . . . 84 Data Acquisition and Electronics . . . . . . . . . . . . . . . . . . . . . . 87 3.9.1 Constant Fraction Discriminator (CFD) . . . . . . . . . . . . . . 87 3.9.2 Time to Digital Converter (TDC) 3.9.3 Timing Sequence . . . . . . . . . . . . . . . . . . . . . . . . . . . 89 . . . . . . . . . . . . . . . . . 89 3.10 Measuring T 0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91 3.11 Resolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91 3.12 Femtosecond Laser System . . . . . . . . . . . . . . . . . . . . . . . . . 92 3.12.1 Oscillator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93 3.12.2 Amplifier . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93 3.12.3 Compressor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94 3.13 Pulse Shaper . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94 3.13.1 Determining the Central Pixel . . . . . . . . . . . . . . . . . . . 97 3.14 Laser Pulse Characterization (GRENOUILLE) . . . . . . . . . . . . . . 99 3.15 Parabolic Mirror and Alignment and Imaging . . . . . . . . . . . . . . . 100 3.16 Z-Scan . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102 CONTENTS 5 4 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses 4.1 4.2 104 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses . . . . 105 4.1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105 4.1.2 Fourier Transform Limited Pulses 4.1.3 + + Molecular structure of H+ 2 , HD and D2 4.1.4 Linear Chirp . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111 4.1.5 Third Order Dispersion . . . . . . . . . . . . . . . . . . . . . . . 117 4.1.6 Combined Linear Chirp and Third Order Dispersion . . . . . . . 132 . . . . . . . . . . . . . . . . . 106 . . . . . . . . . . . . . 110 Conclusions and Future Outlook . . . . . . . . . . . . . . . . . . . . . . 136 5 Laser Induced Fragmentation of CD+ 140 5.1 Laser-Induced Fragmentation of CD+ . . . . . . . . . . . . . . . . . . . 141 5.2 Longitudinal and Transverse Field Imaging Technique . . . . . . . . . . 143 5.3 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147 5.4 5.3.1 Dissociation Channels . . . . . . . . . . . . . . . . . . . . . . . . 147 5.3.2 Branching Ratio . . . . . . . . . . . . . . . . . . . . . . . . . . . 153 5.3.3 Ionization Channels . . . . . . . . . . . . . . . . . . . . . . . . . 154 5.3.4 Charge Asymmetric Dissociation . . . . . . . . . . . . . . . . . . 158 Conclusions and Future Outlook . . . . . . . . . . . . . . . . . . . . . . 159 Research Publications 160 Research Presentations 162 References 164 Chapter 1 Diatomic Molecular Structure and its Interaction with Light Molecules are the fundamental building blocks of nature and constitute everything from the air we breathe to the interstellar clouds of neighboring galaxies. Recent advancements in laser technology has witnessed the onset of ultrafast bursts of electromagnetic radiation which can be delivered on a timescale shorter than the typical molecular vibration and with a magnitude comparable to the binding energy of these systems. Understanding how molecules interact with quantized light can provide a direct insight into their internal structure and used as a tool to characterize and manipulate such systems. This chapter outlines the basic properties of molecules and the nomenclature commonly used to describe them. Furthermore, a plethora of laserinduced dynamics surrounding photodissociation and ionization have been unveiled, and dependence on various laser properties investigated. These dynamics are introduced in this chapter and form a platform to understanding the results presented in chapter 4 and 5 of this thesis. 1.1 Introduction 1.1 3 Introduction It was Lord Kelvin just before the turn of the 20th century who reputedly stated “There is nothing new to be discovered in physics now. All that remains is more and more precise measurement.” But of course, like much of Science and most definitely in physics, many of the most interesting discoveries came shortly after these infamous words. As physicists did indeed take more and more precise measurements and began to probe deeper into the microscopic world. J.J Thomspon discovered the electron and thus began the journey to understand and catalogue the mechanisms involved in atomic and molecular physics. Through theory’s and postulates and experimental trial and error the picture of the atom became clearer. The story of the atom is an interesting one and won’t be covered here, however the knowledge gained from that rich history forms the basis of understanding for the work carried out in this thesis. So as all good physicists tend to ‘stand upon the shoulders of giants’, it is a good starting point to begin by recapping some important information with regards to the underlying mechanisms involved in the atom before moving on from this ‘simpler’ case to the somewhat more complicated molecular picture. 1.2 Atomic Structure An atom is the most fundamental unit of matter in existence and is made up of three subatomic particles; protons, neutrons and electrons. In a classical picture, it can be envisaged as a nucleus, consisting of the protons and neutrons, located amongst orbiting electrons. It is the balance between the electrostatic attraction of these two entities and the outward centrifugal force of inertia which binds them together. Since the electrostatic attraction varies proportional to 1 , r2 where r is the distance of separa- tion between the nucleus and electron, the electrons can be depicted as bound to the nucleus inside a potential well, with an energy corresponding to their distance from the nucleus. In a quantum mechanical model however, the wave-particle duality behavior of the 1.2 Atomic Structure 4 electron must be taken into account when describing the characteristics of the orbiting electrons. This changes dramatically the conceptual image of a particle orbiting the nucleus on a single path trajectory. Instead, we have a probability cloud inhabiting a spatially confined region which can only exist at discrete energies [1]. Such regions are known as atomic orbitals, energy levels or quantum states. These can be adequately described by the quantum numbers contained within the solution to the Schrödinger equation, a wavefunction Ψ. The square of the absolute value of the wavefunction, |Ψ|2 , is interpreted as a probability density, where |Ψ(x)|2 gives the probability density for finding the particle at position x. The conserved quantities of these orbitals can be eloquently expressed using four discrete sets of interrelated quantum numbers n, l, m and s, which can only take the form of an integer or a half integer [2]. The principal quantum number, n, indicates the total energy of the system. This can be physically interpreted as the most probable distance between the electron and the nucleus, therefore defining the size of the orbital. An electron can transfer from one orbital to another through the emission or absorption of a photon. The region of space delineated by each orbital is determined by the wave-like properties of the electron, in particular the number of nodes created. For more energetic systems, a greater number of nodes are created, and the angular momentum quantum number l, which takes the form of an integer value is assigned to quantify this. The values of l = 0,1,2 3 are then substituted by the letters s, p, d, f respectively . The corresponding orbital shapes created by these differing values of l are shown in figure 1.1. Since angular momentum has directionality, the magnetic quantum number, m is used to define the spatial orientation of each orbital. Each electron which inhabits an orbital has an intrinsic spin s, which can pertain a value of s = ± 12 to define its orientation. The electrons are then placed into orbitals abiding to the Aufbau Principle, Pauli Exclusion Principle and Hunds rule to ensure the lowest possible energy configuration is achieved, thus promoting stability [3]. 1.3 Diatomic Molecules 5 Figure 1.1: The shape of the atomic orbitals resulting from an angular momentum quantum number of l = 0,1 and 2 which corresponds to s, p and d respectively. Figure adapted from [4]. 1.3 Diatomic Molecules If two atoms are in close enough proximity to form a bond then a molecule can be created. Diatomic molecules are comprised of two atoms which can either be of the same (homonuclear) or differing (heteronuclear) species, and are bound in a linear arrangement. There are additional degrees of freedom associated with molecules which not only increases the complexity of theoretical calculations, but also makes depicting this information effectively a more intricate task. 1.3.1 Molecular Orbital Theory For diatomic molecules the two nuclei are separated by an internuclear distance R, and the electronic distribution no longer occupies the orbital for a specific atom. Instead it extends throughout the entire molecular domain. A technique known as the linear combination of atomic orbitals (LCAO) can be used to characterize properties such as the respective energies and shapes of the molecular orbitals [5]. This procedure is schematically depicted in figure 1.2 and is based on the superposition of constituent atomic orbitals. In other words, it is simply a weighted sum of the contributing atomic 1.3 Diatomic Molecules 6 Figure 1.2: The respective energy levels and corresponding schematic orbital diagrams for the two atomic 1s wave functions of hydrogen and the resulting molecular orbitals obtained from the linear combination of atomic orbitals (LCAO) molecular orbital (MO) method. The black circles represent both the constructive and destructive interference of the two atomic orbitals, where the wavefunction phases are designated either (+) or (-) relative to their wave ‘up’ or wave ‘down’ displacements. The nodes are regions in which there is zero probability of finding an electron. The diagrams on the far right show the inversion through the geometric center for both orbitals where (a) illustrates odd parity and (b) even parity. orbitals. The molecular hydrogen ion is a one electron system which contains two equivalent 1s atomic orbitals within which the electron can reside. The probability of finding the electron on either of the atoms, for a large internuclear distance R, can be determined by squaring either of the atomic wavefunctions (ψa or ψb ). The stability of the system is determined by the location of the electron. As the distance R between the two nuclei is decreased, the wavefunctions of the two atomic orbitals begin to interfere [6]. The lowest energy solution can be modeled as the linear combination ψa +ψb which constructively overlap to form a high electron density region between the two nuclei. This 1.3 Diatomic Molecules 7 forms what is known as a bonding orbital, denoted σ. It is the electrostatic attraction between the electron and the two independent positively charged nuclei which binds them together. Conversely, in the ψa -ψb combination the two wavefunctions overlap with opposing phases and the creation of nodes, which are otherwise known as the annihilation of the overlap amplitude in the internuclear region, forcing the electron to reside elsewhere. The lack of any electron density between the two nuclei causes them to repel each other, and no bond can be formed. This unstable orbital is higher in energy than the original two 1s orbitals and is referred to as an anti-bonding state σ ∗ . The symmetry of the electronic part of the wavefunction from each state can also be described by an important feature known as parity [3]. If after an inversion through the geometric center of the system the wavefuntion remains the same, it is considered to have even parity (see figure 1.2 (b)), and if any change occurs is denoted as odd parity (see figure 1.2 (a)). This model can be extended to incorporate heteronuclear diatomic molecules. It should be noted however that unlike homonuclear diatomic molecules, the contribution from each orbital is not equal and must be weighted. For the formation of a bonding or anti bonding orbital to occur, sufficient overlap between the constituent atomic orbitals must be achieved. This is more probable for atomic orbitals of similar symmetry and energy. 1.3.2 Molecular Kinematics and Potential Energy Curves Although the LCAO method can provide some insight into the shape of the orbitals, it does not offer any quantitative analysis to support experimental findings. The relationship between the energy of a molecule and its geometry can be graphically represented in a quantitative manner using potential energy curves (PEC). The kinematics surrounding the internal energy Eint of a diatomic molecule is more complex than that of an atom. In addition to the electronic energy Eelec there are nuclear energy components as a result of the vibrational Evib and rotational Erot motions, and these must be taken into account. The total energy of the system can therefore be expressed as Eint = Eelec + Evib + Erot where Eelec Evib Erot . 1.3 Diatomic Molecules 8 In a classical picture, the motion of a diatomic molecule can be considered similar to two balls on a spring which subsequently rotate. As this spring is compressed or stretched through different configurations, the potential energy of the system will vary accordingly. This motion resembles that of a simple harmonic oscillator which obeys Hooke’s Law. However, to explicitly describe the quantum state of a molecule the Schrödinger equation must be solved to calculate the quantized eigenenergies for a collection of mobile electrons at various internuclear distances. This task is incredibly taxing from both a mathematical and computational viewpoint. However, the Born-Oppenheimer approximation proposes that the electronic and nuclear motion can be decoupled due to their significant mass difference. Since the timescales of an electron (10−18 s) and the nuclear motion (∼ 10−15 - 10−12 s) differ by orders of magnitude, the electron is capable of adapting to a modified nuclear position in an instantaneous fashion. In this way, the electronic potential for a given separation can be determined solely by the position of the nuclei. It is therefore assumed that the wavefunction can be separated into its electronic and nuclear (Ψvibrational , Ψrotational ) components and treated independently. This reduces the problem of solving the Schrödinger equation to that of electrons residing in a stationary potential. The potential energy curves of the nuclear configuration as a function of internuclear distance for a hypothetical diatomic are shown in figure 1.3. The origin signifies the center of mass of the system, and the internuclear distance R is defined with respect to this. The equilibrium internuclear separation, R0 , is at the minimum of the potential energy curve, and the strength of the bond is indicated by the depth of the well. This potential well, along with any other bonding orbital can support vibrational states so the population never resides at the very bottom of the well. The temperature of the molecule determines which states are most likely populated and the energy required to fragment each state is different. It is therefore important, in any situation, to understand which states are initially populated, as this will influence the observations. The vibrational energy associated with each individual level can be described by Evib = h̄ωv (υ + 21 )υ where υ is the vibrational quantum number (which must be an integer), h̄ is Planck’s 1.3 Diatomic Molecules 9 Figure 1.3: (a) A schematic diagram of an anharmonic potential well created by the bonding state σ which supports the vibrational levels (red) and the corresponding rotational levels (black). The light blue box outlines the region of the continuum of states within which the anti bonding state σ ∗ exists. The purple line highlights the required dissociation limit from the ground vibrational level and R0 is the equilibrium internuclear separation. (b) Illustrates the potential energy curves for a more complex molecule which possesses a mixture of both singlet and triplet states and demonstrates how the molecular nomenclature system is applied in practice. constant and ωv is the vibrational frequency. Since the strength of the electrostatic attraction between the electron and the nuclei does not vary proportionally with an increase in internuclear distance, the potential well is anharmonic and can be conveniently modeled using the Morse potential [7]. This causes the spacings of the vibrational energy to decrease as they approach the dissociation energy, D0 . Furthermore, for each vibrational level there are numerous associated rotational states. The rotational energy of a molecule is also quantized and is given by Erot = J(J + 1)B where J is the rotational quantum number and B is the rotational constant unique to that molecule. The anti-bonding states σ ∗ (as shown in figure 1.3 (a)) do not form a potential well, hence any population promoted to this state will experience a repulsive force and tra- 1.3 Diatomic Molecules 10 verse out to larger R values, until it is no longer bound. The shape of the potential energy curves mimic the forces of electrostatic forces within the system. As the internuclear distance between the two nuclei approaches zero, the coulomb repulsion between them tends to infinity. Conversely, the electrostatic attraction between the nuclei and the electron diminishes with increasing internuclear distance, until no resultant binding force exists. For more complex molecules the wide range of different electronic configurations which can exist within a molecule can be represented on the same potential energy curves diagram as shown in figure 1.3 (b). Some of the states included may be ionic where one or more electrons have been ejected from the molecule. Since a quanta of energy must be absorbed or emitted for the population to move from one state to another, these curves can be used to determine how much energy is required to pursue these transitions. This enables the expected kinetic energy release Eker shared amongst the dissociating fragments to be estimated from; Eker (eV ) = nh̄ω − Evib (1.1) where n is the net number of photons absorbed, Evib is the energy of the initially populated vibrational level υ of the molecule and ω is the angular frequency of the laser. The potential energy curves can then be efficiently used to deduce the most probable dissociation pathways of diatomic molecules. 1.3.3 Molecular Term Symbols The spectroscopic labels for potential energy curves may give some insight into the electronic configuration of a molecule in that electronic state. Since the four interrelated quantum numbers n, l, m and s used to characterize a particular atomic state (see section 1.2) assume spherical symmetry, an additional set of operators are used to describe diatomic states, where the internuclear axis provides a fixed reference coordinate. The angular momenta of the individual electrons are vectorially added to give Λ, the 1.3 Diatomic Molecules 11 projection of the orbital angular momentum along the internuclear axis, which has a magnitude Λh̄. The assigned integer values Λ = 1, 2, 3... correspond to the designated spectroscopic terms Σ, Π, ∆, Φ respectively, which are analogous to the s, p, d and f orbitals in atomic notation. The spin degeneracy of a diatomic is given by 2S + 1, akin to atoms, where the vector projections are defined in terms of Σ, of magnitude Σh̄, where Σ is a quantum projection for molecules, analogous to m for atoms. Finally the sum of the projections Σ and Λ on the internuclear axis is defined as Ω, of magnitude Ωh̄. All of these properties can be represented in the following molecular term symbol shorthand, 2S+1 (+/−) ΛΩ,(g/u) (1.2) except the Ω term which is normally omitted. For homonuclear molecules the symmetry property parity is an important feature (see section 1.3.1). Since the two nuclei are equivalent the electron cloud distributions have point symmetry about the midpoint of the internuclear axis. The corresponding orbitals are either symmetric with respect to inversion about the midpoint, in which case they retain their original sign; or they are antisymmetric in which case they change sign. Orbitals which are symmetric with respect to inversion are designated g for gerade (even) and those which are antisymmetric are designated u for ungerade (odd). A symmetry property which is included only for heteronuclear molecules for states Λ>0, is the reflection through a plane containing the internuclear axis. The symmetry in this plane can be either symmetric or antisymmetric and is labeled ‘+’ or ‘-’ respectively. Finally, an identity character which precedes the moleculeular term symbol is used to differentiate between states of equivalent properties but situated at differing energies. The ground state is always designated X (orX̃) and excited states of different multiplicity are generally labeled a,b,c,d... [8] in ascending order of energy (see figure 1.3 (b)), although, for historical reasons, this is not always the case. 1.3 Diatomic Molecules 1.3.4 12 Transition Dipole Moment and Selection Rules If a molecule is exposed to a time varying electric field E(t) a transient, oscillating electric dipole moment can be induced and is given by µ = −er, where e is the electron charge and r is the location relative to the nucleus. There are two properties which can be imparted from a photon to a molecule when absorbed. That is energy EP hoton = h̄ω and a unit of angular momentum. The internal energy of the molecule is subsequently raised and the excess energy is dissipated between different modes. This can cause an electron to be promoted from its initial state ψi to one of higher energy ψf . Although molecules can exist in a multitude of states, not all transitions are allowed. The probability of a certain transition occurring is proportional to the amplitude arising from the evaluation of the matrix transition dipole moment shown in equation 1.3 below where V (r, t) = µ · E(t) and hψf |V | ψi i is the matrix element of the transition dipole moment. Z hψf |V | ψi i = ψf∗ V ψi dr (1.3) This is an integral over all space and ultimately provides an estimate of the overlap between the wavefunctions of the initial ψi and final states. For some particular combinations of quantum numbers the matrix element will yield zero, corresponding to a forbidden transition. The relationship between the quantum numbers of the initial and final states for which the transition matrix element is not zero are known as the transition selection rules. In principle, these selection rules apply conservation arguments to the momentum and symmetry of the system. Table 1.1 below presents the electronic transition rules for diatomic molecules and outlines a few possible pathways. However, since the properties of angular momenta are added vectorially there are corresponding selection rules for the rotational quantum number J which must be taken into consideration and are given in table 1.2. In addition to the increase in angular momentum (unit momenta ±h̄) that ensures 1.3 Diatomic Molecules 13 Allowed Transitions ∆Λ = 0, ±1 ∆Σ=0 +↔+ -↔− g↔u Examples Σ ↔ Σ, Π ↔ Π, Σ ↔ Π, ∆ ↔ Π 1 Σ ↔1 Σ,3 Π ↔3 Π,1 Σ ↔1 Π,3 Σ ↔3 Π Σ+ ↔ Σ+ Σ− ↔ Σ− + Σ+ g ↔ Σu , Σg ↔ Πu Table 1.1: Selection rules for the electronic transitions in diatomic molecules for the absorption of a single photon. Electron Transition Σ↔Σ All Others Allowed Transitions ∆J = −1 ∆J = 1 ∆J = 1 ∆J = 0 Table 1.2: Selection rules for rotational quantum number J in electronic transitions for the absorption of a single photon. the parity of the initial and final states must be taken into account to ensure that the overall parity of the interaction is even to prevent the resultant integral from tending to zero. Parity is a multiplicative operator and since both the center of mass and symmetry of a homonulear diatomic molecule are the same, the selection rules are more stringent. The parity of the dipole moment µ is odd so for the absorption of an even number of photons in this case, the initial and final states must differ in parity g ↔ u. Conversely, this rule becomes redundant if an odd number of photons are absorbed as here it is imperative that the two states are of equivalent parity g ↔ g and u ↔ u. For heteronuclear diatomic molecules this inversion symmetry is broken due to the coupling between the electronic motion with asymmetric rotational and vibrational motion of the nuclei around the center of mass. This therefore defies the parity transition rules described above rendering Σ+ ↔ Σ+ and Σ− ↔ Σ− as allowed transitions. Finally, the dipole moment operator does not incorporate any spin-orbit interactions and the integral of transitions involving states of different spin multiplicities is zero. In principal however these singlet-triplet transitions can occur, but the probability is very weak. 1.4 Photodissociation Dynamics 1.4 14 Photodissociation Dynamics In recent years, a plethora of laser induced dynamics has been unveiled, and various theoretical models have been developed in a bid to gain a more comprehensive understanding of the principal mechanisms behind these scientific findings. Owing to its simplicity, the molecular ion H+ 2 has been subjected to extensive experimental and theoretical scrutiny, details of which can be found in several references including [9, 10, 11]. The theoretical models supporting these laser induced dynamics can be segregated into different regimes depending on the pulse characteristics, primarily wavelength, intensity and temporal duration. The polarisability of a light, diatomic molecule is small making it difficult to excite with non-resonant radiation. The majority of studies involving H+ 2 concentrate on Ti:sapphire pulses centered at ∼ 795 nm, for which an appreciable transition dipole moment 1 a.u. - 2.3 a.u [12, 13] can be created for the vibrational levels ν = 0 − 9, respectively. The strength of the radiative coupling of a system can be evaluated by considering the Rabi frequency, ωR , which is given by; h̄ωR = E0 · µ (1.4) where µ is the dipole moment and the magnitude of the electric field E0 = E0 cos(θ) (1.5) which is related to the intensity of the pulse via 1 I = c0 E02 2 (1.6) It is common practice to differentiate between the strong and weak coupling regimes by comparing the Rabi frequency ωR and the vibrational frequency of the molecule ωv [11]. If these two quantities are equivalent, then the field is considered to be intense. 1.4 Photodissociation Dynamics 15 At low intensities the perturbation theory can be acceptably used to describe the transpiring dynamics. In this regime, the dissociation rate is proportional to the laser intensity, and the transition rates from the initial to final states can be adequately obtained from Fermi’s golden rule. For higher intensities where the irradiating field is comparable to the binding energy of the system, the potential well in which the electrons reside can become distorted. The onset of multiphoton processes introduces several nonlinear effects, and subsequently ascertains a breakdown in the perturbation theory and requires a more demanding time dependent treatment. Diabatic processes must be modeled using time dependant calculations, as the temporal duration of the interaction is so short that the system has insufficient time to adapt. This is particularly perceptible for the case of ultrashort pulses (1×10−15 s) where the intensity varies on the same timescale as the molecular vibrational motion. Alternatively, the system is considered to be adiabatic if the temporal duration of the pulse exceeds the timescale of the molecular vibrational motion, leaving the molecule ample time to respond in the presence of the electric field. For the latter case, time-independent methods such as the Floquet or molecular field dressed state formalisms (see section 1.4.2) are suffice and offers a more intuitive representation of the light-induced molecular potentials in strong fields. If we consider the irradiation of H+ 2 by a 795 nm pulse of photon energy 1.55 eV, we need only be concerned with the two lowest electronic states: the bound state 1sσg and the repulsive 2pσu , as the others are inaccessible due to being 11 eV higher in energy. These two states are subsequently coupled and the population undergoes an oscillitary motion which is temporally synchronized with the electric field. At an internuclear distance of ∼ 4.7 a.u these two states become resonantly coupled for vibrational level v=9. The probability of a molecule absorbing a photon during its time located near the resonance is low if the Rabi transition frequency ωR is small compared to the vibrational frequency ωv . However, if the reverse scenario ωR ωv is true, then it is highly probable that the molecule will absorb a photon in this region. The molecule may absorb and emit many times in a process known as Rabi oscillations. The outer turning point of the population for the vibrational level υ = 9 is overlapped with the 1.4 Photodissociation Dynamics 16 resonant internuclear distance, thus creating a sizable Franck-Condon factor for this transition. The vibrational energy of this level is ∼ 0.3 eV with a vibrational period of 29 fs. The Rabi frequency approaches the vibrational frequency at an intensity of 1011 Wcm−2 [12], above which the non-perturbative approach must be employed to describe the laser induced dynamics of H+ 2. 1.4.1 Dissociation It is now a widely accepted phenomena that diatomic molecules irradiated with intense laser light may absorb an excess of energy which promotes the original population to a dissociation continuum and subsequently fragments into its constituent atoms. Despite being the simplest molecular system to study, the photodissociation of H+ 2 has revealed remarkably rich dynamics and will be exemplified here to provide a comprehensive explanation of the underlying mechanisms of the dynamics reported. The photdissociation of H+ 2 into a neutral and an ion can be written in the form: H2+ + nh̄ω → H + H + (1.7) Provided the selection rules are not violated (see section 1.1) the electronic states for a particular transition can couple as a result of the induced polarisability of the molecule. This charge displacement will occur along the direction of the internuclear axis, provided the laser is aligned in the same direction. In the H+ 2 picture the electron is periodically driven back and forth between the two energetically accessible lower lying electronic states, namely the 1sσg and the 2pσu , and this process is known as Rabi flopping. The orbital composition of the bound well, 1sσg (see section 1.3) implies that the electron is located midway between the two nuclei. During irradiation, once the population is projected onto the 2pσu state, the electron can be physically interpreted as having localized itself on one of the nuclei. At this instance, the attractive binding force between the two constituents, (H+ ion and H atom) is very weak, and the laser field can impel the H+ away, slowly (compared to electronic timescales). However, the oscillatory nature of the laser field makes this a continuous process where the electron becomes localized on alternate nuclei, and the internuclear distance increases up each 1.4 Photodissociation Dynamics 17 iteration. Eventually, the electron will be located on one of the nuclei, and the H+ be driven away completely, leading to the breakup of the molecule. 1.4.2 Floquet Formalism There are various techniques which can be employed to solve the time-dependent Schrȯinger equation (TDSE). If the Floquet theorem [14] is adopted, the results provide an intuitive insight into the physical phenomena surrounding dissociation. In this approach, at low intensities where the electronic states of a molecule are regarded as diabatic, they are considered to be ‘dressed’ by ±n photons, meaning they are simply shifted by multiples of photon energy [15, 16]. The field dressed states of H+ 2 are given in figure 1.4 where it is apparent that the 1sσg and 2pσu intersect at υ = 9 and υ = 3 for the 1ω and 3ω shifted curves respectively. These locations indicate the resonant transitions and it is assumed the population will transfer from the 1sσg to the corresponding 2pσu - nω curve and move outward, achieving a greater internuclear separation before dissociating. Since each photon absorbed invokes a parity change of the system, a direct two photon process for H+ 2 is forbidden as an overall odd function would be created, thus contravening the selection rules [10] (see section 1.1). Alternatively, if three photons are initially absorbed and one is emitted prior to reaching the 1sσg - 2ω and 2pσu -3ω interception, (highlighted by the green circle in figure 1.4), the population can cross onto the 1sσg - 2ω curve. The molecule can then dissociate via a net two-photon absorption in a process known as above threshold dissociation ATD (see section 1.4.4). The consequence of irradiating a molecule with a stronger electric field can be sought through diagonalizating the TDSE matrix. The resulting adiabatic Floquet curves qualitatively demonstrate the bond softening effect(see section 1.4.3) and its intensity dependence is illustrated by the broken lines in figure 1.4, where only the pulse intensity envelope has been considered. Despite having been a powerful tool to model the coupling between molecules and monochromatic continuous wave (CW) fields (or at least where the laser period is much shorter than the timescale of the nuclear motion), 1.4 Photodissociation Dynamics 18 Figure 1.4: Molecular potential energy curves for H+ 2 dressed by 792 nm photons. The solid lines represent the field-free curves which have been shifted by multiples of photon energy (0ω, 1ω, 2ω and 3ω). This diagram elucidates three possible dissociation pathways for H+ 2 . Two of these are resonant transitions which require the absorption of 1ω (red circle, υ = 9) or 3ω (blue circle, υ = 3) photons to proceed. In the third scenario, three photons are initially absorbed thus promoting the electron onto the 2pσu , but the emission of a photon prior to reaching the green circle causes the electron to cross over onto the 1sσg , resulting in an overall net two-photon process. For higher intensities these resonant transitions are avoided as the potential energy surfaces are perturbed. The broken lines correspond to the modified potential surfaces which are created for the range of intensities given on the right hand side. Adapted from Frasinski et al [18] it is only recently that the Floquet formalism has been suitably adapted for use with ultrashort pulses (≤ 10 fs) [17]. 1.4.3 Bond Softening The vibrational state energy of a molecule and peak intensity of the pulse play a critical role in determining the photodissociation dynamics most likely to occur. Molecules with a vibrational state energy either at or close to the one-photon crossing are not heavily influenced by the properties of the electric field as they dissociate early on in the pulse, even for low intensities. 1.4 Photodissociation Dynamics 19 For high intensities, where the strength of the laser field is comparable to the binding force of the molecule, the potential surfaces can become distorted and figure 1.5 shows this effect in H+ 2 for a range of intensities. The external field causes the two potential curves to repel each other, creating what is commonly referred to in the adiabatic picture as an ‘avoided crossing’. The height of the potential barrier can be reduced to an energy below that of the vibrational state energy and the population can subsequently flow out and dissociate via bond softening [19, 20]. Alternatively, the molecule can tunnel through the finite width of the suppressed barrier, however the likelihood of this is small compared to the dissociation over the barrier. The probability of photodissociation for these levels increases nonlinearly with intensity. Due to the oscillatory nature of the laser field, these avoided crossings are not static, and will vary between the diabatic and adiabatic curves in sync with the intensity profile of the pulse over time. The temporal characteristics of this mechanism cannot be overlooked, as for a significant portion of the population to dissociate over the barrier the potential barrier must be sufficiently suppressed on a timescale comparable to the vibrational motion of the population within the bound well. 1.4.4 Above Threshold Dissociation There are several different mechanisms than can lead to the fragmentation of a molecule, but the resulting pathway largely depends on the initial vibrational state and the intensity and frequency of the laser field. 14 Wcm−2 ) the |2pσ − 1ω> chanIf H+ u 2 is irradiated with a strong enough field (5×10 nel becomes redundant as the |2pσu − 3ω> one opens up allowing the vibrational levels at or below the three photon crossing to escape (see figure 1.4 (blue circle)). The dissociating portion of the wavepacket may proceed to dissociate along the |2pσu − 3ω> curve. Alternatively, if an adequate intensity prevails until the wavepacket reaches the |2pσu − 2ω> pathway, a photon can be re-emitted causing the molecule to dissociate via a net two photon process. This mechanism is known as above threshold dissociation (ATD) and the molecule dissociates with a higher than expected kinetic energy 1.4 Photodissociation Dynamics 20 Figure 1.5: The two lowest lying electronic states of H+ 2 in the absence of an electric field (dashed line). In the presence of a strong enough electric field these electronic surfaces can become perturbed. The extent to which the barrier is suppressed depends on the time averaged intensity, where 1×1013 Wcm−2 (dotted line) and 5×1013 Wcm−2 (solid line) are exemplified here. Reproduced from Posthumus and McCann [21]. release (KER). Since an excess of the net minimum number of photons required to break the bond energy were absorbed, the peaks in the KER spectrum are separated by the photon energy [19]. Theory suggests that for a wide range of intensities this could be the dominant dissociation mechanism [22]. At greater intensities even higher order ATD process in H+ 2 , such as the net 4ω can be invoked and even dynamically controlled [23]. In general, ATD has a higher KER and lower probability of occurring than BS because of the larger number of photons involved. Despite the experimental evidence for the existence of ATD being manifested in the KER spectra of H+ 2 where the peaks are separated by the photon energy [19], these results were considered ambiguous as the product fragments of the coulomb explosion channel also span the same KER range. This issue was resolved by trapping and cooling HD+ to its ground state and subsequently measuring the neutral fragments [24]. Furthermore, clear evidence of a 2ω ATD channel in the experimental KER spectrum of CO2+ was concurrent with theory at these high intensities [25]. 1.4 Photodissociation Dynamics 21 [h] Figure 1.6: Mechanism of vibrational trapping. (a) As the intensity increases on the rising edge of the pulse, an avoided crossing is initiated, and a portion of the wavepacket can traverse through. (b) Near the peak of the pulse the avoided crossing is accentuated, creating a shallow potential well which essentially traps the traversing population. (c) This shallow well becomes inverted as the intensity decreases on the trailing edge of the pulse and the population can either dissociate or revert to the bound potential well.[18] 1.4.5 Vibrational Trapping If a molecule is exposed to an electric field of sufficient intensity, the adiabatic surface of the bound state forms a maximum turning point which alleviates the liberation of vibrational wavepackets through the bond softening effect (see section 1.4.1). The adverse is true of the repulsive state as the forces act to bend the surface upwards and form a concave turning point which is capable of confining the population. This phenomenon is commonly referred to as bond hardening, vibrational trapping or ‘stabilization’ [26, 27, 18, 28]. Figure 1.6 presents a schematic diagram of the evolution of a light induced well and hence the mechanism for vibrational trapping for H+ 2 at distinct points in the pulse. It is assumed in figure 1.6 (a) that the vibrational state υ = 4 is populated and as 1.4 Photodissociation Dynamics 22 the intensity increases on the rising edge of the pulse the population travels towards the avoided crossing induced at the |1sσg − 0ω> and |2pσu − 3ω> intercept. At this stage in the pulse, the gap at the avoided crossing is very small and the probability of dissociation is relatively low. An escaping wavepacket from the υ = 4 or vibrational states just above the intersection can cross onto the upper curve, which is materializing towards a shallow potential well. As the peak intensity is approached, the depth of the well increases, causing the population to be momentarily trapped (see figure 1.6 (b)). Thus a reduction in the dissociation probability of highly excited vibrational states has been experimentally observed [28]. As the intensity decreases on the tail end of the pulse, the potential energy surface is inverted as it reverts back to its diabatic configuration. This effectively creates a hill from which some of the population can either return to the bound well or be repelled toward larger R and dissociate. The mechanism of vibrational trapping is a controversial topic despite the experimental evidence showing that pulses of temporal duration in the range 45-500fs, obtained by chirping, can be used to manipulate trapped wavepackets [18]. It was found that the rise time of the pulse intensity plays a critical role in the kinetic energy release (KER) of the fragments as this controls the height of the hill from which the dissociating wavepacket is ejected and hence a greater KER was measured for shorter pulses. It has been suggested that this model is an artifact of an aligned one dimensional quantum mechanical model and that this theory does not hold if nuclear rotations are incorporated [29]. 1.4.6 Below Threshold Dissociation As the potential surfaces from the 3ω avoided crossing revert back toward their diabatic configuration on the trailing edge of the pulse, the light induced potential well transforms in such a way that it can cause the vibrational level υ =4 to dissociate along the 1ω pathway, despite this transition seemingly violating energy conservation. As weaker intensities are approached, the shape of the repulsive state changes from concave to convex (see figure 1.6) causing it to form the bottom part of the 1ω anti 1.4 Photodissociation Dynamics 23 crossing. During this process the trapped wavepacket is lifted up and about half of it falls back into the bound 1sσg well, and the other half escapes along the 1ω trajectory [18]. This process, since intuitively it is below threshold to occur is referred to as below threshold dissociation (BTD). The fragments will emerge with a very low kinetic energy release (KER), which is ultimately determined by the intensity profile of the pulse, where fragments with higher KER are observed for a more rapid decrease in intensity. This effect manifests itself in as a shift from 0eV to 0.3 eV in the low energy proton peak when the pulses are shortened from 540 fs to 45 fs. To comply with energy conservation this energy gain can be described in terms of the dynamic Raman effect. Rabi flopping, as described in section 1.4, is a continuous process in which photons from the field are continuously absorbed and re-emitted. Owing to the bandwidth of the pulse, these photons differ slightly in energy. Thus if a photon from the lower wavelength region of the spectrum, with higher energy, is initially absorbed and a photon of lower energy is re-emitted then these small energy gains are enough to enable this process to proceed. If this same mechanism occurs at the 1ω crossing where the population will emerge via the 0ω trajectory then it is known as zero-photon dissociation (ZPD). 1.4.7 Angular distribution of fragment ions The direction in which the dissociating fragments are ejected is not random but depends upon the overlap between the wavefunctions of the initial and final states. For convenience, in the laboratory frame, the reference axis is given by the polarization of the laser and the angular distribution of the fragments is given with respect to this. The probability of excitation is proportional to |E · µ| where E is the electric field vector and µ is the transition dipole moment. Provided the axial recoil approximation, which assumes that the molecules dissociates in a period which is short compared to the rotational period of the molecule, is valid, then the angular distribution (θ) of the velocity v of the fragments with respect to the laser polarization is given by: I(θ) α 1 + βP2 (cosθ) (1.8) 1.4 Photodissociation Dynamics 24 Figure 1.7: (a) For pure parallel transitions in the axial recoil limit, the bond axis, the transition dipole moment µ, and the recoil direction v are parallel, so that the photofragments posses an axis of cylindrical symmetry about these vectors. (b) for a pure perpendicular transition, µ is perpendicular to the bond axis and v, so that the photofragments possess two orthogonal planes of symmetry. (c) For a mixed transition, where µ is neither parallel nor perpendicular to the bond axis, the photofragments posses only one plane of symmetry. Where θ is the angle between the laser polarization and the velocity, υ, of the fragments, the anisotropy parameter β indicates the correlation between the µ and υ and P2 is the Legendre polynomial of order 2. For transitions between two states with a similar angular momentum quantum number Λ (Σ → Σ or Π → Π) in which ∆ = 0, the direction of the transition dipole moment and internuclear axis must be aligned (see figure 1.7(a)). These are known as parallel transitions and from equation 1.8 it can be shown that the angular distribution of the fragments from a pure parallel transition is proportional to cos2 (θ). Alternatively, a perpendicular transition is one in which the angular momentum quantum number Λ of the states change (Σ → Π or Π → ∆) and corresponds to a ∆± 1. Since this transition stipulates that the laser polarization and internuclear axis must be orthogonal (see figure 1.7(b)), the angular distribution of a pure perpendicular transition is proportional to sin2n (θ), where n is the number of photons involved. 1.5 Ionization Processes in Intense Laser Fields 25 For a collection of heteronuclear molelecules, dissociation can proceed via a multitude of possible pathways (see figure 1.7(a)), which can either be perpendicular or parallel transitions. The angular distribution in this case will be a combination of cos2 (θ) and sin2n (θ), and where an indication of the branching ratio between certain degenerate pathways can be determined by the dominant contribution. The fact that both types of transitions are observed shows the individual nature of the molecules electronic states plays a role in the molecular dissociation. 1.5 Ionization Processes in Intense Laser Fields The process of liberating an electron from a molecule is termed ionization. This phenomena is incredibly rich in dynamics as this procedure can be achieved through various mechanisms which is ultimately determined by a combination of the properties characterizing both the molecule and the radiating field. The binding energy of the valence electron of a molecule is known as its ionization potential Ip , and subsequently indicates the minimum amount of energy required to remove this electron. If the energy E = h̄ω of the photons within the electromagnetic field is adequate such that the following condition h̄ω>Ip may be satisfied, then the electron can effectively be promoted to the continuum after absorbing a single electron. This process is illustrate by the single arrow in the schematic diagram 1.8 (a), where any excess energy manifests itself in the kinetic energy of the liberated electron. However, since the Ip of the majority of molecules is quite high, the criterion above is rarely met. In this case, it is not the photon energy which becomes important, but instead the photon density, ρ which depends upon the intensity I and wavelength λ of the laser field ρ = Iλ3 hυc . It was proposed by Göeppert-Mayer [31] that a sequence of photons can be absorbed consecutively, causing the electron to undergo several transitions through a series of virtual states, as it advances toward the continuum. This multiphoton ionization (MPI) mechanism may proceed, nh̄ω>Ip , where n is the integer number of photons, provided the uncertainty principal (∆E∆t ≥ h̄ 2) is not violated. This condition stipulates that a successive photon must be absorbed within 1.5 Ionization Processes in Intense Laser Fields 26 Figure 1.8: Schematic diagram showing the three different ionization regimes. In (a) at lower intensities < 1014 Wcm−2 , multiphoton processes are dominant whereas in (b) at the highest intensity < 1015 Wcm−2 , one reverts to the tunnel ionization model. In (c) the field completely suppresses the coulomb barrier (> saturation intensity) over-the-barrier [30]. the lifetime of the preceding state. Thus a high photon density ρ is required. It is also plausible that an electron can be promoted to an intermediate state which is a real excited state. If this occurs, then progression to the continuum can cease and the electron can remain in that state. Analogous to dissociation processes (see section 1.4.1), ionization mechanisms can be classified into different regimes depending on the intensity of the electric field. At relatively low intensities < 1013 Wcm−2 where MPI processes are dominant, the system can be described using a first order perturbation theory as the potential surface of the double potential well remains unperturbed. The MPI probability P is given by: P = σN I N (1.9) where N is the number of photons absorbed, I is the laser field intensity and σN is the cross section for an N absorption process. Similar to above threshold dissociation (ATD), the number of photons absorbed may surpass that required for the MPI process and this is termed above threshold ionization (ATI), where the excess energy is transfered to the electron. The system can no longer be treated as a perturbation once the electric field strength becomes comparable to the binding energy of the electron, > 1013 Wcm−2 , as the potential well within which the electron resides becomes distorted, namely suppressed, 1.5 Ionization Processes in Intense Laser Fields 27 and creating a barrier of finite width. The ionization process under such conditions cannot be visualized as the the absorption of photons, but instead as a strong electric field inducing a barrier suppression through with the electrons can tunnel through toward the continuum. As the distortion to the barrier increases the resulting binding potential experienced by the electron is reduced. As the field intensity is raised the barrier suppression increases until the electron is no longer bound and can escape, this is termed over the barrier (OTB) ionization (figure 1.8 (c)). In the presence of a temporally changing electric field E(t), the physical width of the barrier varies dynamically in accordance with the frequency of the sinusoidal oscillations of E(t). If the barrier is sufficiently suppressed, an electron can quantum mechanically tunnel through the finite width and escape to the continuum (figure 1.8 (b)). The rate of these tunneling events can be calculated using the Ammosov, Delonoe and Krainov, the so called ADK, model [32]. For the tunnel ionization mechanism to be effective the wavepacket must have enough time and energy to escape, hence the tunneling time must be short compared to the period of the E(t). The Keyldsh parameter, γ, is commonly used to distinguish between the MPI and TI regimes, where γ is the ratio of the tunneling time to the time at which the electric field remains quasi-static, or equivalently the ratio of the ionization potential to twice the ponderomaotive potential of a laser pulse, which for the case of an atom is given by: s γ = Ip 2Up (1.10) where UP is the ponderomotive or cycle averaged quiver energy of a free electron, of mass me , moving in the presence of the electric field of angular frequency ω and peak amplitude E0 ., where the atomic core is neglected Up (a.u.) = e2 E02 4me ω 2 =⇒ Up (eV ) = 9.33 × 10−14 I(W cm−2 )λ(µm)2 (1.11) The shape of the potential surface is not taken into consideration when calculating γ using the formula above. It employs a zero range potential which over estimates the 1.5 Ionization Processes in Intense Laser Fields 28 value of γ for molecules. A molecular parameter, γM , which corrects the width of the barrier through which the electron must tunnel by incorporating the extended length of molecular electronic orbitals, can be more a appropriate representation, see [33] for further details. For any given system, the value of γ is used to categorize it into a specific regime, according to the following condition; γ 1 =⇒ TI and OTB γ 1 =⇒ MPI The dominant ionization mechanism is effectively depicted by the properties in equation 1.11. The probability of tunnel ionsiation (TI) is higher for lower values of γ. The width of the potential barrier tapers with increasing intensity, and the temporal duration for which the barrier endures this suppression, prior to the reversal of the electric field, is wavelength (frequency) dependent. Thus the preferred criteria for TI is high intensity, E0 , (>1013 Wcm−2 ) and a long wavelength (low frequency). For alternative conditions MPI will be favored, but it is unclear as to which mechanism will dominant around the region for which γ = 1. If during tunnel ionization the wavepacket gains insufficient drift momentum to escape the attractive potential of the newly created ion, then it can subsequently be captured into an excited Rydberg orbital of the atom, in a process termed frustrated tunnel ionization (FTI) [34]. This procedure is thought to occur at the trailing edge of the laser pulse in order to conserve both energy and momentum. The wavepacket is driven by the oscillating electric field, as proposed in the the recollison model (see section), which causes the wavepacket to gradually decelerate, over a series of cycles, toward the trailing edge, and as a result it has insufficient energy to escape the coulomb field of the atom and is subsequently recaptured into an excited Rydberg orbital. Since the outcome of FTI and a direct excitation of an electron to an excited Rydberg state is identical, this process has been studied as a function of the ellipticity of the incident laser light. As the ellipticity of the light increased, the probability of the electron returning to the ionic core decreases and so a reduction in the high KER excited 1.5 Ionization Processes in Intense Laser Fields 29 neutral fragments was measured [35] in accordance with that expected for an FTI mechanism. 1.5.1 Non Sequential Ionization It was realized that the theoretical models such as the ADK and its advanced single active electron (SAE) version which use single ionization exclusively to predict the ionization probability rate of the He → He+ → He2+ process were incorrect as not only was the onset of this process much lower than expected, but that the theory did not match experiment [36]. Furthermore, the presence of a ’knee’ structure was indicative of a non-sequential ionization mechanism in which the transition proceeds as follows He → He2+ and the conventional He+ step is evaded. Over recent years, multiple photoionization processes have been under intensive scrutiny and the exact mechanism underlying this non-sequential ionization process is a controversial topic since a photon can only couple to a single electron and which results in the simultaneous ejection of two or more electrons after the absorption of a single photon and is therefore thought to be driven entirely by many electron correlations for which the following three mechanisms have been proposed, shake-off, collective tunneling and rescattering model; The shake − off model was proposed by Fittinghoff et al [37] as a source of nonsequential ionization in an intense laser field. It claimed that if the ionization of the first electron occurred rapidly enough, then the remaining electrons would have insufficient time to react adiabatically to the instantaneous change in their molecular environment. This could cause the initial wavefunction to be projected onto the new system whose unaltered potential may have some overlap with the continuum state, resulting in one or more electrons being subsequently ’shaken off’. Furthermore, in an analogous mechanism, the electrons may not be removed simultaneously but instead have a finite probability of being ’shaken up’ onto a bound excited state, which decays via further ionization. It has also been suggested that whilst the barrier is suppressed, in the presence of the electric field, more than one electron may simultaneously quantum mechanically 1.5 Ionization Processes in Intense Laser Fields 30 tunnel to the continuum in a Collective tunneling process [38]. Although these two proposed methods provide a viable explanation for the onset of an increased yield in multiply charged ions, they are not compliant with the observation that this effect is strongly correlated with the ellipticity of the incident light. It has been experimentally observed that non-sequential ionization processes are strongly suppressed for circularly polarized light [39] [40]. However, the third proposed mechanism, electron rescattering is coherent with the observation that ionization has an intrinsic dependence on beam ellipticity. The three-step model proposed by Corkum [41] introduced a ‘plasma perspective’ on strong field ionization processes, and has been effectively used to eloquently communicate non-sequential ionization mechanisms and high harmonic generation on a classical level. Figure 1.9 provides a schematic illustration of this ‘simple man model’ and each step will be referred to during the following explanation. In the first step, a molecule undergoes tunnel ionization within the first few cycles of a linearly polarized pulse, leaving the liberated electron at the disposal of the lingering electric field. In the second step, the electron is driven by the laser field, which oscillates away from the ionic core in the first instance. The accelerating electron accumulates ponderomotive energy (see equation 1.11), and then when the oscillating laser field changes direction, the motion of the electron also reverses, causing it to revisit the ionic core. The acceleration a, velocity v, and distance x, displaced from the ionic core (typically of the order ∼ 10−10 m) are highly dependent on the release phase ϕ, and can described, in a one-dimensional plane at an instantaneous time t, by the following expressions; a(t, ϕ) = − (1.12) Eo [cos (ωt + ϕ) − cos ϕ] ω (1.13) Eo [sin (ωt + ϕ) − sin ϕ − ωt cos ϕ] ω2 (1.14) v(t, ϕ) = x(t, ϕ) = eE0 [sin ωt + ϕ] me 1.5 Ionization Processes in Intense Laser Fields 31 Ultimately, the trajectory the electron follows depends upon the phase at which it was born in the field. The interaction between the ionic core and the returning electron constitutes the third step, and may be scattering or recombinative in nature. If an inelastic scattering process occurs, then the ponderomotive energy accumulated by the electron (max 3.17 Up ) may be imparted to another electron, thus liberating it via a collisional ionization event. Alternatively, a recollision induced excitation with subsequent field ionization (RESI) event may occur whereby the returning electron excites a bound electron to a higher energy state which is subsequently ionized later in the pulse. Furthermore, from various photoelectron spectra, it is evident that the returning electron can be elastically scattered from the charge cloud of the newly formed ion. The cut-off energy for this process was found to be 10Up , where the electron has been backscattered through 180◦ off the core. If the electron recombines into the ground state of the parent ion, then a photon with a frequency which is a high-order integer multiple of the fundamental frequency of the incident light will be emitted. The energy of these photons is determined by the ponderomotive energy of the electron Up and the ionization potential Ip of the state to which the electron recombines. Since the maximum kinetic energy the electron can acquire in the field is 3.17Up this cusps an upper limit for the energy of photons created in the high harmonic generation (HHG) process. h̄ω = Up + Ip =⇒ (h̄ω)max = 3.17Up + Ip (1.15) The x-ray pulses produced from HHG are of the shortest possible temporal duration, being in the attosecond regime (10−18 s), and the prospect of using these pulses as an imaging tool for time resolved studies of electron dynamics is regarded an exciting endeavor. For more details on HHG and its applications the reader is referred to [42] and references therein. 1.5 Ionization Processes in Intense Laser Fields 32 Figure 1.9: Summary of the three step model. The electron is ionized in step 1 at some particular phase of the electric field. It is then driven away from the parent ion in the laser field (step 2). After sign reversal of the ac-field, the electron stops far from the atom, possibly returns an recombines to emit a photon carrying a kinetic energy of the electron plus its ionization potential (step 3). The kinetic energy of the returning electron can be as high as 3.17Up , defining the so-called cut-off photon energy in the harmonic spectrum, figure adapted from [42]. 1.5.2 Charge Asymmetric Dissociation It is widely accepted that a molecule can fragment via a charge-symmetric channel (CSD): AB+ → A+ + B or A + B+ However, in recent years, evidence to support the existence of charge-asymmetric (CAD) dissociation has been reported [43]: AB+ → AB2+ + e− → A2+ + B + e− It has been proposed that the electronic states involved with CAD processes closely resemble the charge-resonant states introduced by Muliken in 1935 [44]. These can occur when there is a difference in charge state in the dissociation limit of a diatomic molecule. They can exist as a gerade and ungerade pair which become asymptotically degenerate at large R and possess a dipole coupling which increases proportional to 1.5 Ionization Processes in Intense Laser Fields 33 R. Energetically, these states lie significantly higher than the symmetric dissociation limits, an example of this can be seen in figure 1.10, where the the N 2+ + N 0+ channel is situated 12.74 eV higher than the N 1+ + N 1+ , which equates to the absorption of an excess of 9 photons, at 800nm. It is for this reason that the existence and an understanding of the mechanism by which these states become populated has remained, to a large extent, elusive. It is believed that the mechanism by which these states are populated is caused by the fact that in the presence of a strong enough field a large induced dipole represents an asymmetric displacement of charge which would lead to a natural asymmetry in charge states of dissociating products. This explanation concurs with the experimental findings that the yield of the asymmetric channels increases with increasing intensity [45]. Furthermore, it was found that for 600ps interaction of 248 nm with N2 , that the CAD processes dominated [43]. It was suggested that in a system behaving in this manner that the preferred fragmentation mechanism does not depend on the energetics of the system, but instead selects the dynamical mode of interaction with the strongest coupling. Initially CAD was not considered to be a ubiquitous process as it was only observed + for N+ 2 and I2 , both of which were assumed to be special dications with accessible metastable states and that CAD process arises from a special precursor state. However, the single ionization charge-asymmetric (2,0) channel for a series of molecules N+ 2, CO+ , NO+ and O+ 2 for the interaction with 790nm laser pulses for a range of intensities up to 7×1015 has been recent;y reported [45]. They note that these channels are considerably weaker, with a < 10% yield than the symmetric channel (1,1) and a higher appearance intensity [45]. The kinetic energy release (KER) for the CAD (2,0) channel was found to average at ∼ 3eV which is less than half that of the CSD (1,1) channel at ∼7 eV. It was proposed that the Coulomb repulsion between the (1,1) charged centers will result in a higher KER. 1.5 Ionization Processes in Intense Laser Fields 34 Figure 1.10: A partial potential-energy diagram of N2 indicating the ground N2 X1 σg+ and the coulombic contributions to several ionized species. Assuming vertical excitation, the ion energies observed will be one-half the amount indicated 1.5.3 Coulomb Explosion and Multielectron Dissociative Ionization The ionization mechanisms discussed in the previous section 1.5 are applicable to both atoms and molecules as they are non-dissociative in nature, implying the removal of one or more electrons results in the creation of a cation of a higher multiply charged state p, as outlined, for the absorption of n photons, in the process below; AB + + nh̄ω =⇒ AB p+ + (p − 1)e− (1.16) However, owing to the additional degrees of freedom of a molecule, dissociative ionization may occur, whereby in addition to the removal an electron, the molecule can break apart. AB + + nh̄ω =⇒ A2+ + B + e− or AB + + nh̄ω =⇒ A + B 2+ + e− (1.17) 1.5 Ionization Processes in Intense Laser Fields 35 In the presence of a strong enough field, a multielectron dissociative ionization (MEDI) event may proceed, which can effectively be described as ionization followed by coulomb explosion. AB + + nh̄ω =⇒ Ap+ + B q+ + (p + q − 1) e− (1.18) The removal of multiple electrons from a molecule, as outlined in equation 1.18, results in a mutual repulsion between the two charged fragments causing them to coulomb explode apart. The kinetic energy release (KER) of these fragments is higher than that obtained from other dissociation pathways and can be determined from equation 1.19 below; KER1 = µ q1 q2 4π0 R where µ= m2 m1 + m2 (1.19) where qi and mi are the residual charge and mass of the two fragments respectively, R is the internuclear separation at the moment of coulomb explosion and µ is the reduced mass to ensure the KER is shared between the fragments so that momentum is conserved. However, the experimentally measured KER was found to be lower than expected at Ro , indicating that the MEDI occurred at a larger internuclear distances than expected. Some of the various mechanisms which have been identified to explain multielectron dissociative ionization of a diatomic molecules are illustrated in figure 1.11, where each step can be accomplished through either tunneling or multiphoton absorption. Each different pathway can be considered as either a direct or indirect ionization process, depending on the evolution of the internuclear separation R throughout the ionization procedure. Direct ionization processes [46] occur around the equilibrium separation Ro , meaning the molecule is not stretched in the field and that the nuclei can be considered as effectively frozen. A conventional direct ionization via TI or MPI is illustrated in figure 1.11 (a), where the KER of the fragments produced in this way is higher as the repulsive forces of two nuclei is distributed over a shorter distance. In addition to this, figure 1.11 (b) shows a schematic representation of electron rescattering [47], as discussed in section ??. From this it is apparent that there is insufficient time for the 1.5 Ionization Processes in Intense Laser Fields 36 Figure 1.11: Schematic diagrams illustrating the different mechanisms for the multiple ionization of a typical molecule, AB+ , in an intense laser field (a) direct ionization, (b) electron rescattering (nonsequential ionisation, NSI), (c) enhanced ionization, (d) stretch only in dissociation and (e) stairstep process [45] molecule to stretch within the return time of the electron, which is approximated to be ∼ 3/4 of a laser cycle, and so the the ionization will occur at an R similar to that of the initial ionization step. Conversely, in indirect ionization mechanisms the molecule will undergo a stretch, prior to the ejection of the electrons, a widely accepted example of which is charge-resonance enhanced ionization (CREI) [21], [48]. In this instance, the internuclear distance R, is increased to some critical value Rc , at which the electron becomes localized at one of the potential wells and subsequently the removal of an electron becomes easier. Hence the ionization rate at Rc is enhanced (see figure 1.11 (c)). Since Rc >Ro , this accounts for the lower KER which has been experimentally observed, and several quantum mechanical treatments of this mechanism exist to fully quantify the process (see discussion below). In addition to CREI, other possible MEDI mechanisms include the molecule being promoted to the continuum after having being stretched to an internuclear distance in the range Ro <R<Rc , at the initial dissociation stage (see figure 1.11 (d)). Alternatively, MEDI may proceed via a stairstep process in which the electrons advance toward the continuum via a series of sequential transitions which are preceded by a small increase 1.5 Ionization Processes in Intense Laser Fields 37 in internuclear separation (see figure 1.11 (e)) [10], [49]. Figure 1.12: Schematic diagram of an I+ 2 molecule ion in a laser field. The full curve represents the potential energy of the outermost e− ina combined field of two point like atoms I+ 2 and the laser. The internuclear distance and the laser field Fc have critical values such that the electron energy level EL touches both the inner potential barrier Ui and the outer potential barrier Uo [21] In the classical field ionization (FICE) model proposed by posthumus et al [21], the molecular double-well potential U, for a molecule aligned with an electric field of amplitude Fc is given by; Q1 Q − 2 R − Fc x U = − R x + x − 2 2 (1.20) where Q1 and Q2 are the residual charges on the two fragment ions, x is the distance along the internuclear axis and R is the internuclear separation and all entities are in atomic units. The outer valence electron of the molecule is regarded as a freely moving entity which can span the entire double potential well, defined by U, at a specific electronic energy level EL given by equation 1.21 below, whereas the remaining electrons are confined to their constituent atomic orbitals (see figure 1.12). EL = (−E1 − Q2 R ) + (−E2 − 2 Q1 R ) (1.21) 1.5 Ionization Processes in Intense Laser Fields 38 Figure 1.13: Double potential wells for I+ 2 at three different internuclear separations in a strong external field, illustrating the calculation of the classical appearance intensity of I+ + I+ . The numbers in the figure denote the laser intensity, eg. 7.9e12 represents 7.9 × 10 12 W cm−2 . Adapted from [50] where E1 and E2 are the ionization potentials of the two constituent ions and EL is considered as the average ionization potential distributed across the internuclear separation, R. In the CREI model the critical internuclear distance Rc , of a diatomic system is defined as the separation at which the energy of the valence electron, EL and the height of both the inner (U1 ) and outer (UO ) potential barriers are equivalent, such that ionization via over the barrier (OTB) mechanism can prevail. Figure 1.12 illustrates these conditions for the (1,1) channel of I2+ which occur at an internuclear distance of 7 a.u. and an applied intensity, Fc , of 4 × 10 13 Wcm−2 [21]. The appearance intensity of each ionization channel (Q1 , Q2 ) is highly dependent on the internuclear separation, as the inner potential barrier Ui , between the two atomic potential wells, heightens with increasing R. Thus attaining the scenario depicted in figure 1.12 becomes an interplay between the appropriate R and corresponding intensity of the laser field, such that both barriers are adequately lowered. This effect is illustrated in figure 1.13 which shows how the appearance intensity of the (1,1) channel of I+ 2 varies with R. 1.5 Ionization Processes in Intense Laser Fields 39 Figure 1.14: Classical appearance intensities of the fragmentation channels of I2 (full curves) and classical trajectories (broken curves). The laser parameters corresponding to each of the trajectory calculations are as follows [21] Around the equilibrium internuclear separation Ro , the inner potential barrier Ui poses no hindrance to the motion of the valence electron as it can occupy the entire region between the two atomic wells. The electronic energy level EL in this case is located at a relatively low energy, and the only stipulation in this scenario is that the electric field is strong enough to suppress the outer potential barrier enough to enable ionization to occur (similar to R = 5 in a.u. in figure 1.13 (a) and (d)). As the nuclei are moved further apart the inner potential barrier Ui will eventually rise above the electronic energy, EL causing the electron to become localized in one of the atomic potential wells. However, the onset of stark shifts can cause the electronic energy EL in the upper potential well to be raised above the inner potential barrier Ui . These shifts combined with a more effective lowering of the outer potential barrier Uo caused by the increased distance over which the electric field acts creates a minimum in the appearance intensity of this ionization process. Hence the critical separation in this model is when the stark shifted EL and both the Ui and Uo are equal (see figure 1.13 (b) and (e)). If the internuclear separation is increased further still, then the continually rising inner potential becomes insurmountable for the electron to pass and an increase in laser intensity is required (see figure 1.13 (c) and (f)). The appearance intensities as a function of internuclear separation R, as obtained from the FICE model, are plotted in figure 1.14 for various ionization channels (Q1 , 1.6 Molecular Alignment 40 Q2 ) of I2 . The distinct minimum in all curves around ∼ 10 a.u. correspond to the critical internuclear separation Rc and illustrate explicitly that the required equilibrium between the the physical entities described above occurs over a very small range of R. 1.6 Molecular Alignment Laser-induced alignment is defined as the molecular axis being oriented at an angle (θ=0) parallel to the direction of the laser polarization [51]. The response of a molecule to an electric field can differ according to this angle as many dynamics demonstrate a strong alignment dependence. Alignment can be considered to be either a geometric or dynamic mechanism, ultimately however a combination of both are expected. In geometric alignment the molecules are initlally aligned parallel to the field. For geometric alignment the laser polarization couples with the molecule dipole to exert a torque on non-aligned molecules. Irradiating a molecule with an electric field of insufficient intensity to ionize, polarizes the molecule creating an induced dipole moment: Vµ = −µ · E(t) (1.22) where µ is the the dipole operator and E(t) = E0 cos(ωt) the electric field. The fieldfree molecular Hamiltonian Ho of the molecule then evolves in the electric field into the eigenstates of the new Hamiltonian Hmol [52]: Hmol = Ho + Vµ + Vα (1.23) where the induced dipole interaction potential is given by: 1 Vα = − E 2 (t)(∆α cos2 θ + αper ) and ∆α = αpar − αper 2 (1.24) The αpar and αper are the components of the polarizability tensor parallel and perpendicular to the molecular axis, respectively. The induced dipole moment interacts with 1.6 Molecular Alignment 41 the electric field and rotates the molecule to a position that minimises its energy. It is the interference from Rabi-type osciallations and the continual exchange of a number of units of angular momentum with the field which populate the rotational wavepackets. The degree of rotational excitation is determined by either the pulse duration or the balance between the laser intensity and the detuning from resonance. 1.6.1 Adiabatic Alignment If the temporal duration of the pulse is greater than the rotational period of the molecule (τ >τrot ) then the alignment is considered to be adiabatic. The electric field oscillates on a femtosecond timescale which is too rapid for the molecule to follow on its picosecond response time. This condition requires that the interaction be averaged over several laser cycles. This in turn leads to the permanent dipole term being negated and the induced dipole dominates. As the electric field is adiabatically switched on the field-free eigenstates H0 evolve to eigenstates of the complete Hamiltonian Hmol . These so-called ‘pendular’ states librate about the polarization axis [53]. As the electric field is adiabatically switched off, the eigenstates return to the field-free Hamiltonian H0 . 1.6.2 Non-adiabatic Alignment Laser pulses which are shorter than the rotational period of the molecule (τ <τrot ) can be used to induce non-adiabatic alignment. Within a quantum mechanical framework, a coherent superposition of rotational states (rotational wavepacket) is created via a series of Raman excitations and deexcitations. All the states within the wavepacket acquire phase at integer multiples of the fundamental angular frequency. It is the dephasing and rephasing of the individual components of the rotational wavepacket that act to align the molecule along the direction of polarization. Thus ‘field-free’ alignment can be achieved shortly after the peak of the pulse. Furthermore, a ‘revival’ can occur at a time commensurate with the ground state rotational period of the molecule, even after the field has passed. Chapter 2 Ultrafast Measurement Techniques The development of highly sophisticated spectrometric methods has enabled the various laser-induced dynamics described in the previous chapter to be eloquently explored. This chapter gives an overview of the experimental properties which inevitably make interpreting the data complicated. Some potential solutions to overcome these predicaments and extract the relevant information are then provided. The nature of these experiments make measuring all the resulting fragments simultaneously and unambiguously extremely challenging. A few widely used experimental techniques to investigate topical areas in this field are outlined. In addition to the advancements in detection systems, the development of laser technology has propelled the ideology of moving from passive to active quantum control methods. The realization of using laser pulses to drive chemical reactions is now tangible due to the onset of pulse shaping. The principal behind pulse shaping and a few commonly used techniques are also discussed in this chapter. 2.1 Experimental Work 2.1 43 Experimental Work There has been impressive progress made towards gaining a comprehensive understanding of how atoms and molecules behave in the presence of a strong laser field. However, the experimental observations are a convolution of various parameters concerning the target molecules and the laser. These include the central wavelength and spectral bandwidth of the laser, peak intensity and temporal duration, the focusing of the laser beam and the interaction volume, the vibrational and rotational population of the target ions and the resolution of the detector. 2.1.1 Neutral vs. Ionic Targets There are two experimental methods used to prepare ions for laser induced fragmentation. In this discussion, H+ 2 is used as an example to elucidate the two processes. The H2 molecules can be ionized on the rising edge of the pulse to form H+ 2 , and then fragmented around the peak of the same pulse [54, 55]. One drawback with this method is that it restricts analysis to intensities greater than that required to ionize the neutral H2 (> 1×1014 Wcm−2 ). Hence obtaining any information on low intensity dynamics is inaccessible. It was found that the lower vibrational states [56] are predominantly populated using this method. Moreover, the exact distribution was found to be wavelength and intensity dependent [57]. Alternatively, the development of a fast ion beam target [58] has permitted the dy14 Wcm−2 ) to be thoroughly investigated. namics of H+ 2 at lower intensities (< 1×10 Thus any ambiguities introduced in the intermediate ionization stage are eradicated. The most powerful aspect of this technique is the initial translational velocity of the ions which means that the neutral dissociation fragments have enough energy to be efficiently detected. Ions formed in an electron impact ion source have their vibrational states populated according to a Franck-Condon distribution [59, 60, 61]. The nature of the ion production however means that coherence between the vibrational states of separate ions cannot be achieved, unlike with the laser. Furthermore, heteronuclear molecules can undergo internal transitions during the flight time from the ion source 2.1 Experimental Work 44 Figure 2.1: Profile of a focused laser pulse showing the isointensity contours bounded by the saturation intensity of ion species A+ , A2+ etc. The slit (ion beam) therefore probes a particular intensity and volume region. Figure adapted from [62]. to the interaction region and so the vibrational population of such molecules upon interaction with the laser can be ambiguous. A major drawback of this technique is the low target density of ion beams which limits the rate of measurements and promotes time consuming experiments. Focal Volume Effect The spatial intensity distribution of a focused Gaussian laser pulse is extremely complicated. It varies in a Lorentzian manner along the direction of propagation z, and Gaussian along the radial direction, r. This effectively creates a series of isointensity contours, as shown in figure 2.1. The intensity experienced by an ion depends upon the position at which it interacts with the pulse and can be determined from equation 2.1. This makes extracting information for a particular intensity a non-trivial task. I(r, z) = 1+ I0 2 exp −2r2 2 ω0 1 + zz0 z z0 (2.1) The dissociation yields measured are therefore a convolution of the dissociation rate and the intensities enclosed within the signal producing volume. High intensity regions 2.1 Experimental Work 45 occupy a lesser volume than those of lower intensity. This leads to a problem known as the ‘volume effect’. By exposing the entire laser pulse to the detector, the dissociation signal becomes dominated by the large-volume, low-intensity regions which tend to an I 3/2 trend at saturation. These contributions can be reduced by employing the intensity selective scan (ISS) technique [63, 64]. The interaction volume exposed to the detector can be restricted by the FWHM of the ion beam (see the slit in figure 2.1). A series of cylindrically symmetric thin slices of width ∆z are created and vary only in the radial direction. The laser pulse can then be translated along the direction of laser propagation z by translating the focal lens and effectively probe different intensity regions. The volume effect can be removed from an angular-resolved kinetic energy release (KER) spectra (eg., see figure 2.2 (A)) using a technique known as the intensitydifference-spectrum (IDS). In principal IDS is the difference between two KER-cosθ distributions obtained under the same conditions but with different peak intensities. This allows the intensity dependence of a feature to be qualitatively investigated. The spatial and temporal distribution of the laser beam are kept constant and beam splitters and neutral density filters are used to attenuate the pulse energy. A sequence of measurements are made for various intensities, and the appropriate subtractions allow exactly the contribution from the intensity range between the two peak intensities to be extracted. This method stipulates that the size of the ion beam must be larger than the laser focus, but much smaller than the Rayleigh range of the laser beam. Thus the interaction region is actually in a 2D configuration [65]. 2.1.2 Pulse Intensity Effects An insight into the dynamics of photodissociation has been attained from studying its intensity dependence. Obtaining a quantitative comparison between published data is a non-trivial task due to the numerous techniques employed to change the peak intensity probed. Neutral density (ND) filters can be used to decrease the peak intensity and have the added advantage of maintaining a constant spatial and temporal profile. Alternatively, the intensity selective scan (ISS) technique described in section 2.1.1 can 2.1 Experimental Work 46 Figure 2.2: (A) The KER-cos θ distributions for the laser-induced dissociation of H+ 2 for 135 fs pulses of peak intensity I0 = 2.4×1014 Wcm−2 . The peak intensities are then labeled in each respective panel. (B) The KER distribution of (a) the filled circles and (c) the open circles. The arrows mark the KER from the v=8 level at each peak intensity. Only the dissociation inside |cosθ| >0.9 is taken into account. Figure adapted from [66]. be used at the expense of a variation in the interaction volume for different intensities. Furthermore, if the temporal duration of the pulse is increased, the peak intensity decreases but the overall pulse energy can be kept constant. The latter method is usually achieved by chirping the pulse (see section 4.1.4 ). However, recent studies have shown that the dissociation dynamics can be influenced by the instantaneous frequency distribution of the pulse and care should be taken when interpreting such data [67]. The intensity dependent dissociation phenomena for H+ 2 is qualitatively illustrated in figure 2.2 (A). The various KER-cosθ distributions were measured for 135 fs pulses starting at a peak intensity of I0 = 2.4 ×10 14 Wcm−2 and reduced using ND filters. Molecules that dissociate via bond-softening demonstrate a continuous shift toward lower kinetic energy release (KER) for increasing intensity. This effect is better observed in the vibrational structure of the one-dimensional KER distribution shown in figure 2.2 (B). Only molecules dissociating inside |cosθ| >0.9 around the laser polarization direction are taken into account and the arrows indicate the shift in the v=8 2.1 Experimental Work 47 Figure 2.3: The KER-cos θ distributions for the laser-induced dissociation of H+ 2 for 14 −2 a peak intensity I0 = 2.4×10 Wcm . (a) The intensity averaged spectrum, for a pulse duration of 135 fs. (b) The IDS spectrum for (a). (c) The intensity averaged spectrum for a pulse duration of 45 fs. (d) The IDS spectrum for (c). Figure adapted from [66]. vibrational level. This shift can be explained by the widening of the gap at the avoided crossing for higher intensities. Furthermore, above threshold dissociation (ATD) is seen to peak at around 1.2 eV. This is observed only for the higher intensities as this effective two-photon dissociation feature is the result of the opening of the 3 photon curve crossing and subsequent emission of a photon. 2.1.3 Temporal Duration The effect of the temporal duration τ on the dissociation of H+ 2 was qualitatively investigated by comparing the KER-cos θ distributions for 135 fs and 45 fs pulses of equivalent peak intensity. Considering the laser-induced avoided crossing changes dynamically with intensity and pulse duration, interpreting the data can be problematic due to the interdependence of these two parameters. The measured spectra and corresponding IDS spectra (see section 2.1.1) shown in figure 2.3 demonstrate the dramatic differences observed. The IDS spectra in figure 2.3 (b) and (d) demonstrate that dissociation induced by 2.2 Experimental Imaging Techniques 48 shorter pulses is dominated by ATD, whilst bond-softening prevails for the longer pulses. Although vibrational structures can be observed for the long pulses (135 fs) this is not true of the shorter pulses. This is a consequence of the vibrational period for the vibrational levels from which the H+ 2 molecule dissociates (20 fs) approaching the pulse duration (45 fs). Furthermore, the narrow bond softening channel below 0.4 eV observed for the shorter pulse does not appear for the longer pulses. 2.2 Experimental Imaging Techniques There are various experimental techniques used to study the behavior of molecules in strong laser fields. These all contribute some additional information or clarity to established scientific findings. In principal, to study ultrafast dynamics it is necessary to identify the resultant fragments and measure their kinetic energy. There are various experimental methods capabable of delivering this criteria but they vary drastically in complexity, sophistication and cost. 2.2.1 Time of Flight Arguably the easiest and most effective way to analyze a full mass range of resulting fragments simultaneously is using time-of-flight (TOF) mass spectrometry. Such devices operate on the principle that ions can effectively be temporally and spatially separated unambiguously using their mass to charge m/q ratio. A linear mass spectrometer is comprised of two adjacent regions. The first acceleration region contains a uniform electrostatic field E of potential U and length S. The second is a field free ‘drift’ region of length D as shown in figure 2.4 (a). Ions of charge q are produced at some potential Ep = U q in the acceleration region and gain some kinetic energy Ek = 12 mv 2 . The ions then traverse the ‘drift’ region with a constant velocity v. The ions TOF is governed by the m/q ratio as higher charged states are accelerated to higher energies in the extraction region, thus they have a shorter TOF. Similarly, heavier ions have a lower velocity and longer TOF. The measured TOF for the ions 2.2 Experimental Imaging Techniques 49 Figure 2.4: (a) Schematic diagram of the principle behind a linear time-of-flight (TOF) device, where an ion of charge q, is ‘born’ (purple circle) amidst a homogeneous electric field Es and accelerated over a distance S before traversing a field free region D and striking the detector at a corresponding t time later. (b) Typical TOF spectra from a neutral xenon target showing charge states Xe+ → Xe8+ and illustrating the concept that higher charge/mass states have a shorter flight time and hence will arrive at the detector first (adapted from [68]). incorporates the distance from the position of formation to striking the detector. The sequential detection of the different ions leads to a temporal spectrum, that may easily be converted to a plot of signal intensity versus ion mass-to-charge ratio, an example of which is shown in figure 2.4 (b). The TOF technique is advantageous as theoretically it has an unlimited mass range and a high transmission. Furthermore, it enables a complete m/q spectrum for molecular fragmentation and is relatively cost effective. The resolution of this device however is hindered by ions formed at non-zero velocity. This is a consequence of the focal volume 2.2 Experimental Imaging Techniques 50 of the laser producing fragments at slightly different potential energies U q within the electric field. To compensate for the initial distribution of energies and increase the resolution, a Wiley-McLaren spectrometer comprised of two acceleration regions was developed, see [69] for details. 2.2.2 Covariance Mapping Figure 2.5: Two dimensional covariance map of carbonyl sulphide measured using 790 nm light 55 fs pulses at an intensity of 2 × 1014 Wcm−2 (adapted from [68]). The covariance mapping technique was initially used to unambiguously identify the fragmentation channels of the multi-electron dissociative ionization (MEDI) process in intense sub-picosecond laser fields [70]. The TOF of fragments for each laser shot is measured but not averaged over a number of shots as in conventional averaging procedures. While averaging TOF results improves the signal-to-noise ratio, any inherent statistical fluctuations from shot-to-shot caused by the laser pulse can become lost. Instead, covariance mapping compares the changes in one measurement with another measurement, by way of a shot-by-shot analysis. Covariance is defined as a measure of association between two random variables. The covariance between two points on the TOF spectrum is determined by the average vector minus the vector product of the 2.2 Experimental Imaging Techniques 51 averages (for mathematical details see [71]). A typical spectra is shown in figure 2.5 where the fragment ions which are statistically correlated with themselves appear on the diagonal line and are meaningless. The real events corresponding to the forwardbackward and backward -forward pairs are situated off-diagonal. The momenta of each pair of ions can be determined and used to identify the coulomb explosion (CE) channels, as labeled on the right hand side of figure 2.5. False coincidences can occur if the target gas pressure is high enough to allow more than one CE event to occur per laser shot. An example of this is highlighted by the red circle in 2.5 where a correlation between C+ and C2+ from an OCS molecule is unfeasible. 2.2.3 COLTRIMS Figure 2.6: Schematic diagram showing the layout of a COLTRIM. The ability to perform kinematically complete laser-ion interaction measurements with no limitations on the collection angle of the reaction fragments can be achieved using COLTRIMS (Cold Target Recoil Ion Momentum Spectroscopy). This imaging technique has the capability of measuring the momentum vectors of all charged fragments (electrons and ions) over a 4π solid angle. The device is essentially comprised of an ion and electron TOF spectrometer located at opposing ends of the interaction region, as shown in the schematic diagram in figure 2.2 Experimental Imaging Techniques 52 2.6. The copper plates generate a weak uniform electrostatic field and following the interaction, are used to extract the charged fragments in a direction corresponding to their charge. Then after traversing their respective acceleration region, the fragments traverse a field free drift region before impinging a 2D position sensitive channel plate detectors with multi-hit capability. Due to the considerably large kinetic energy of the electrons, achieving a high acceptance of these particles is challenging. The pair of Helmholtz coils situated along the spectrometer axis act to confine the electron motion in space by enforcing them to traverse in cyclic trajectories. The resolution and acceptance angle of the ion and electron can be changed by modifying the magnetic field strength and the spectrometer voltage. These instruments are highly versatile as this can be done independently for each particle. The nature of these experiments stipulates that to increase the resolution the target ions must be cold (below thermal motion at room temperature). Usually they are injected into the interaction region using a cooled supersonic gas jet. Thus only a small initial momentum (≤ 0.05 a.u.) is projected along the laser polarization axis. From the measured position and TOF information the fragments trajectories can be reconstructed and the initial momentum vector calculated unambiguously. This technique is particularly useful when studying three-body breakup as the momentum vectors of all components can be determined. 2.2.4 Velocity Map Imaging The velocity map imaging (VMI) technique is a powerful tool used to study molecular dissociation as it provides the resolution needed to determine the structure and energy levels of the ion. The energy imparted to a molecule, in excess of the dissociation threshold, is partitioned between the kinetic and internal energies of the constituent atoms. Following dissociation the ejected fragments from a single vibrational level lie spatially distributed on a Newton sphere with a radius R, defined by the conservation of momentum. This sphere then expands at a rate determined by the velocity 2.2 Experimental Imaging Techniques 53 v0 of the fragments as it traverses towards the detector. The three-dimensional velocity distribution of the fragments is projected onto the detector. Thus creating a circularly shaped pattern as shown in figure 2.7 for the neutral dissociation fragments of H+ 2 . The detector consists of a multichannel plate (MCP) mounted in front of a phosphorus screen. As the localized electron avalanche exiting the MCP impinges the phosphorus screen, luminous photons are emitted and their positions recorded using a charge-couple device (CCD) camera. The full three-dimensional information can be reconstructed from the projected two-dimensional image using a mathematical transformation called an Abel inversion. It should be noted that this method does not provide a means of completely distinguishing between the dissociation and ionization processes due to an overlap in their kinetic energy release (KER) distributions. Figure 2.7: Two-dimensional momentum projection of the neutral photofragments at a pulse energy of 1.0 mJ and a wavelength of 785 nm. (a) τ = 575 fs, I0 = 3.5 ×1013 Wcm−2 (b) τ = 135 fs, I0 = 1.5 ×1014 Wcm−2 . Adapted from [72]. 2.2.5 Pump-Probe Schemes An optical technique which can be coupled with the experimental set-ups described above to initiate and image time-resolved vibrational and rotational dynamics of molecules is a pump-probe scheme. For these experiments two (or more) laser pulses are required. The pump pulse creates a molecular wavepacket, and the probe pulse images how this wave packet evolves with time. Usually the variable in these studies is the the delay τ between the pulses. 2.3 Laser Pulse Shaping 54 For this scheme to work, the effective duration of the probe and the pump pulse has to be shorter than the time scale of the process of interest. Furthermore, the temporal resolution of the scheme is increased for shorter laser pulses. A major advantage of this scheme is that following the pump pulse, the system evolves by the natural eigenstates of the field-free system. Also, addition information can be obtained from so-called ‘two color’ pump-probe measurements where two synchronized sources of short pulses are used. 2.3 Laser Pulse Shaping Pulse shaping has played a critical role in the quest to control chemical reactions using laser pulses. There are two types of pulse shaping schemes which can be employed. The closed-loop system can be implemented using a programmable pulse shaper and a learning algorithm to find the specific pulse shape suitable for optimizing a reaction. The signal from a particular process is fed within a feed back loop into a computer and analyzed using an evolutionary algorithm. The pulse shape is then modified based on the outcome of the algorithm. This iterative or adaptive process is then effectively used to find the pulse shape required to maximize the yield of the predefined target product. Although this is a powerful tool, obtaining any insight into the physical mechanism induced by that particular pulse is a formidable task. Alternatively, in an open-loop scheme the user theoretically predicts the desired pulse shape and then creates and tests the hypothesis. This is a more intuitive method and relies on prior knowledge of the molecular system in question. 2.3.1 Time and Frequency Domains The electromagnetic waves of linearly polarized femtosecond (1×10−15 s) laser pulses can be fully described by the time and space dependent electric field. A complete description of the pulse can be given in either the time or the frequency domain. e + (t) and its complex The real electric field E(t) is composed of a complex electric field E 2.3 Laser Pulse Shaping 55 e − (t) contain the positive and negative values respectively, where E(t) is conjugate E given by: e + (t) + E e − (t) E(t) = E (2.2) The spectral amplitude of the pulse is centered around a carrier frequency ω0 , and the complex electric field can be represented by the product of an amplitude function A(t) and a temporal phase term φ(t): E + (t) = A(t)e−i[ω0 t−φ(t)] (2.3) E − (t) = A(t)ei[ω0 t−φ(t)] (2.4) and Where the latter two properties contain all the information about the laser pulse. The complex electric field is given by the Fourier integral: 1 2π Z 1 E (t) = 2π Z E + (t) = ∞ e + (ω)eiωt dω E (2.5) e − (ω)eiωt dω E (2.6) −∞ and − ∞ −∞ e + (ω) and E e − (ω) are a complex representation of the electric field in the Where E frequency domain. Similar to equation 2.2 the complex electric field for the frequency domain can be written as: e e + (ω) + E e − (ω) E(ω) =E (2.7) where each of these spectrum components can be expressed as: iϕ(ω) e + (ω) = A(ω)e e E (2.8) −iϕ(ω) e − (ω) = A(ω)e e E (2.9) and e Where A(ω) is the spectral amplitude and ϕ(ω) the spectral phase. The relation 2.3 Laser Pulse Shaping 56 between the spectral and temporal representations of the pulse is then given by the Fourier integral: Z + ∞ E (ω) = e + (t)e−iωt dt E (2.10) e − (t)e−iωt dt E (2.11) −∞ and Z − ∞ E (ω) = −∞ e and The description of laser pulses in the time domain using a temporal amplitude A e and a phase a phase φ(t) or in the frequency domain using a spectral amplitude A ϕ(ω) are equivalent. If the spectral phase ϕ(ω) of a pulse varies slowly with frequency ω, it can be expanded in the frequency domain into a Taylor series around the carrier frequency ω0 : 1 00 1 000 0 ϕ(ω) = ϕ(ω0 ) + ϕ (ω0 )(ω − ω0 ) + ϕ (ω0 )(ω − ω0 )2 + ϕ (ω0 )(ω − ω0 )3 + . . . 2 6 (2.12) By analogy, this Taylor expansion can be performed in the time domain and the time derivative of the temporal phase φ(t) defines the instantaneous frequency ω(t): ω(t) = ω0 + Pulses with dφ(t) dt dφ(t) dt (2.13) = 0 are known as Fourier transform limited and contain no phase φ(t) variation. However, for dφ(t) dt = f (t) the carrier frequency varies with time and the corresponding pulse is said to be frequency modulated. The terms in equation 2.12 can be used to characterize laser pulses as follows: • Absolute Phase ϕ(ω0 ): The carrier envelope phase describes to the phase between the envelope of the electric field and the carrier frequency. This is particularly important for extremely short pulses as they only have a few cycles within the pulse. 0 • First Derivative ϕ (ω0 ): The Linear delay term or so-called “group delay” generates a constant group delay. This shifts the entire pulse in time with respect to an arbitrary origin. 2.3 Laser Pulse Shaping 57 00 • Second Derivative ϕ (ω0 ): The group delay dispersion (GDD) is a quadratic spectral function. A frequency sweep, commonly known as linear chirp is created by the corresponding linear group delay. Each frequency component ω(t) experiences a linearly increasing delay as the spectral components of the pulse are scanned. 000 • Third Derivative ϕ (ω0 ): Third order dispersion (TOD) is a cubic function which causes the spectral components of the pulse to be redistributed according to a quadratic group delay. The interference between the high and low frequencies causes beating phenomena which manifests itself as post/pre pulses in the time domain. Shaping femtosecond pulses directly in the time domain is difficult due to the response time of electronic devices. Instead, pulse shaping can be more readily achieved by multiplying the spectral components in the frequency domain by a transfer function M (ω) of a device or medium. 2.3.2 Passive Optical devices; Materials, Prisms and Gratings Femtosecond pulses can be considered as being composed of several groups of quasi monochromatic waves. Each group can be envisaged as a wavepacket with a narrow spectrum, all of which can be added together coherently. In vacuum, the group velocity of the wavepackets are constant and equal to the speed of light. In reality however, the refractive index of each group of quasi monochromatic waves is wavelength (hence frequency) dependent. This causes each group to acquire a different group velocity and subsequently broadens the temporal profile of the pulse. As the pulse traverses transparent media in the optical domain, positive GDD can be introduced. This occurs as the higher frequencies travel faster through the medium than the low frequencies. This creates what is known as an ‘up-chirped’ pulse. The time reversal of this situation known as a ‘down chirped’ pulse is induced for negative GDD and can be created by the angular dispersion of prisms or gratings. Although this technique can be used to chirp pulses, any higher order spectral terms (see equation 2.3 Laser Pulse Shaping 58 2.12) introduced in this process cannot be eliminated. Furthermore, the intensity of the pulses cannot be too high or else new frequencies are generated due to self phase modulation. 2.3.3 Acousto-Optic programmable Dispersive Filters (AOPDF) A method used to control the phase and amplitude of an ultrashort laser pulse is an acousto-optic programmable dispersive filter (AOPDF), often referred to as a dazzler (see the schematic diagram in figure 2.8). It relies on an acousto-optic interaction between an acoustic wave, whose spatial and temporal shape is controlled by a radio frequency transducer, and a laser pulse [73]. The laser pulse traverses the ordinary (fast) axis of the acousto-optical crystal alongside the acoustic wave. The phase matching between the two results in diffraction which causes the laser pulse to transit from the ordinary (fast) to the extraordinary (slow) axis. The frequency components will experience a delay if the velocity of the laser pulse on these two axes differs. Figure 2.8: Principle of operation of a dazzler. In an acousto-optic interaction the phase matching between the two waves cause the optical wave to transit from the ordinary (fast) axis to the extraordinary (slow) axis. A time delay is introduced between each frequency component if the velocity along the two axes differ. (adapted from [73]). 2.3.4 Masks in the Fourier Plane The conventional 4-f pulse shaper configuration used for Fourier transform pulse shaping is shown schematically in figure 2.9. The incoming pulse Ein (t) is dispersed by 2.3 Laser Pulse Shaping 59 Figure 2.9: Standard design for the conventional 4-f setup for Fourier-transform femtosecond pulse shaping. Adapted from [74] the first grating creating several groups of spatially separated quasi monochromatic waves. These waves are then focused to a minimum beam waist at the Fourier plane – focal plane of a converging optical lens – by a lens (or curved mirror to eliminate chromatic abberations) of focal length f. Each individual spectral component can then f(ω)) located at the Fourier plane. After be manipulated by the pulse shaping mask (M traversing the mask, the laser pulse is reconstructed by performing an inverse Fourier transformation back into the time domain. The most representative spatial light modulator pulse shaping masks used to modify the optical path of each group of quasi monochromatic waves are described below; Liquid Crystal Spatial Light Modulators (LC-SLM) This programmable pulse shaping tool is comprised of a linear array of independently controlled pixels. The applied voltage changes the refractive index of each pixel. The spectral components are therefore retarded with respect to one another by the frequency dependent phase added by the specific pixel they are spatially mapped to. To a first approximation a phase-only LC-SLM does not change the spectral amplitudes and the integrated pulse energy remains constant for different pulse shapes. Acousto-Optic Modulator (AOM) The AOM crystal is oriented at the brag angle to the Fourier plane in the 4-f setup. An acoustic (sound) wave created using 2.4 Coherent Control 60 a radio-frequency (RF) electrical signal to drive the piezoelectric transducer is propagated through the crystal. This induces a change in the refractive index of the crystal, effectively creating a grating, and light passing transversely through the medium is then diffracted. The diffracted beam is shifted in frequency by an amount equal to the electrical drive frequency (typically in the one hundred MHz range), ideally with an amplitude and phase that directly reflect the amplitude and phase of the RF drive. Flexible Membrane Mirrors The front plate (reflective side) of a mirror composed of flexible thin metal coated silicon nitride can be reshaped by the voltage applied to the electrodes beneath the surface. A movement accuracy of a few microns is achieved using electrostatic actuators to control the voltage. The different spectral components are dispersed as they are reflected off the mirror, and the different distances traversed by different spectral components acts to shape the pulse. This device is widely used a corrective optic within the laser system. 2.4 Coherent Control Research into laser-ion interactions is motivated by the ambition to understand and optically drive chemical reactions to the highest degree of specificity [75, 76, 77]. Over the years various active coherent control strategies have been developed and exemplified. They largely rely on effects such as coherence, interference and time-delays between laser pulses [78, 79, 80]. The realization of pulse shaping, and in particular learning algorithms [81, 82, 83, 84, 85, 86], has expedited the control of molecular dynamics such as the bending and stretching of molecules (e.g., for CO2 [87]) and electron localization [88]. Furthermore, the ability to perform selective-bond cleavage and rearrangement as well as optimize branching ratios has been readily achieved [85, 89]. Although learning algorithms are relatively simple to implement and require no prior knowledge of the molecules Hamiltonian, obtaining a mechanistic explanation is a formidable challenge. In recent years, our understanding of the behavior of small molecules in the presence of 2.4 Coherent Control 61 strong laser fields has been acquired through varying laser parameters such as the peak intensity, duration, central frequency, bandwidth and the carrier envelope (see section 2.1). Our strategy to achieve and understand coherent control of photodissociation relies on the interaction of the theoretically tractable H+ 2 , with analytically shaped, well characterized pulses. The spectral phase function φ(ω) used to shape the pulses discussed in this thesis can be expressed by certain terms of the Taylor expansion described in equation 2.12. Using an LC-SLM pulse shaper (see section 2.3.4) the magnitude and sign of these terms were used as a tool to control the dissociation yield of the low lying vibrational levels v≤6 of H+ 2 and its isotopic variants. Furthermore, the combination of experimental results and theoretical calculations elucidate the underlying mechanism responsible for these findings. 2.4 Coherent Control 62 Chapter 3 3D Momentum Imaging Technique The experimental results presented in chapters 4 and 5 of this thesis were carried out at two different institutes; the Weizmann Institute of Science, Rehovot, Israel (WIS) and Kansas State University, Manhattan, Kansas, USA (KSU). Although the respective experimental setups were used to explore different aspects of laser-induced dissociation, the 3D momentum imaging technique employed to measure the momentum components of the fragments was the same. The two set-ups differ only in ion optic configurations and the laser systems used. Such experiments involve generating, extracting and focusing ions to form a well collimated ion beam. The laser beam is then perpendicularly focused onto these ions amidst a longitudinal spectrometer located in the interaction region. The longitudinal dc electric field of the spectrometer acts to accelerate the charged fragments relative to their neutral counterparts. This allows homogeneous dissociation channels to be unambiguously identified. The position and time of flights of the dissociation fragments is measured. From this information the 3D momentum components of each fragment can be calculated and used to determine the kinetic energy release (KER) of the dissociation process. Furthermore, the angle between the molecular axis and laser polarization at the time of dissociation can be obtained. The experimental apparatus 64 can be considered as two separate sections; the ion beam and the laser system. A schematic diagram of the whole experimental layout is shown in figure 3.1. The following chapter describes the various experimental components and outlines the setup procedure followed at WIS. Figure 3.1: A schematic diagram of the experimental system. 3.1 Neilson Ion Source 3.1 65 Neilson Ion Source The molecular hydrogenic ions (H2 + ,HD+ and D2 + ) were generated simultaneously in a Neilson-type ion source [90]. The construction of the ion source is shown in a schematic diagram in figure 3.2. A tungsten filament is mounted coaxially within a hollow, cylindrical anode which is subsequently enclosed within a magnetic field which acts parallel to the axis of the anode. The ion source region was maintained at a base pressure of ∼ 7×10−8 mbar. Then the target gas mixture (composed of 45% H2 , 45% D2 and 10% Ar) was admitted through a long, thin walled gas inlet into a region close to the filament to achieve a pressure of ∼ 3×10−6 mbar. During operation, the filament is incandescently heated so that the chamber is maintained at a constant temperature of ∼ 800◦ . The anode is held at a slightly positive potential relative to the filament (∼ 100 volts) and the surge of electrons, emitted via thermionic emission of the filament, are accelerated towards it. These energetic electrons can then collide with the background gas with sufficient energy to knock electrons from the neutral target atoms and produce positive ions. The presence of the magnetic field causes the positively charged ions and electrons to gyrate in opposite directions. This increases the frequency of collisions and the concentration of electrons, which in turn augments the ionization rate. When collecting data from these experiments the ion beam is expected to maintain stability for a number of consecutive days. Maintaining a constant pressure is a critical aspect of achieving beam stability as under certain operating conditions a plasma can be created and the ionization rate becomes proportional to the gas density. All the apparatus was grounded with the exception of the ion source which is floated at some potential (typically 5keV). The ions were extracted and accelerated between the anode and the conical shaped extraction electrodes and then traverse the beam line with an energy equivalent to the source potential. The anode current was used to monitor the stability of the plasma conditions. Any minor fluctuations were corrected by adjusting the filament current or anode voltage to maintain the pressure stability within the anode. This method will not affect and focusing and steering properties of 3.2 Vacumn System 66 Figure 3.2: A schematic diagram of the Neilson-type Ion Source as viewed from (a) above (b) the side. the beam. It should be noted that the electron impact ionization mechanism generating the ions in this source involve vertical transition where the vibrational population approximately follows the Franck-Condon factors [59]. 3.2 Vacumn System A ‘perfect vacumn’ is a space which is completely empty of matter, however these conditions cannot be experimentally achieved. The term vacumn is therefore considered to be the number of particles contained within a certain volume and can be calculated using the Ideal Gas Law P V = nRT (3.1) where P is pressure in Pa, V is volume in m3 , n is the number of moles of gas where 1 mol is 6×1023 particles, R is the universal gas constant = 8.314J kg−1 mol−1 and T is the temperature in Kelvin. In these experiments the entire beam line is maintained under vacumn to prevent unwanted interactions and background collisions from being measured, to maintain clean working surfaces and to prevent any chemical re- 3.2 Vacumn System 67 actions. Depending on the capability of the vacuum technology, the lowest possible pressures achieved are generally classed as either vacuum (∼ <10−3 mbar, V), high (∼ <10−8 mbar, HV) or ultrahigh (∼10−12 mbar, UHV) vacuum. There are four gate valves along the beam line effectively separating slightly different vacumn regions. During operation the ion source is maintained at a relatively high pressure (∼ 3×10−6 mbar) compared to the rest of the beam line and differential pumping is required. A small pipe (conductance limiting apertures) creates a transition region between between the two chambers (Wien filter to adjacent chamber see figure 3.1) so that the pressure ratio across this connecting stage is manageable. This strategy is possible due to the longer mean free path in the lower pressure region. Two (name) scroll pumps (see section 3.2.1) are used to initially back out the system to a pressure of <10−3 mbar. Then four turbo pumps (see section 3.2.2) are used to reduce the pressure further along the beam line until UHV (7∼10−9 mbar) is achieved in the detection region. Several ion gauges are used to obtain pressure readings (see section 3.2.3). 3.2.1 Scroll Pumps Before the turbomolecular pumps can be switched on the chamber must be evacuated to a rough vacumn or to achieve the ’molecular flow’ and such that the molecules don’t interact with each other(∼1×10−3 mbar). A scroll pump is an oil free backing pump composed of two interleaved archemidean spiral shaped scrolls. One scroll is fixed whilst the other encompasses an elliptical path without rotating. The air becomes trapped and compressed between the scrolls and is eventually successively compressed out through a central exhaust. 3.2.2 Turbomolecular Pumps To achieve and maintain HV/UHV turbomolecular pumps can be used to reduce the pressure from the ‘molecular flow’ (10−3 mBar) to the ‘viscous flow’ (10−7 mBar) region. Turbo pumps consist of a series of rotors with angled blades. Inbetween these 3.2 Vacumn System 68 rotors are fixed stators with blades oriented in the opposite direction. As the rotor begins to spin momentum is imparted to those gas molecules hit by the rotating blades. A net gas flow is then created in a particular direction due to the angle of the blades and the molecules are driven toward the exhaust where they are collected by a backing pump. The volume flow rate (pumping speed) of these momentum transfer pumps is constant. The throughput and mass flow rate of the system decreases exponentially as less mass mass becomes available for evacuation after pumping has commenced. However, the constant throughput into the system introduced from real and virtual leakage, evaporation, sublimation, desorption of materials and back streaming rates means the turbo pump will reach a maximum compression ratio and the pressure will become constant. 3.2.3 Hot-Filament Ionization Gauges Several hot-filament ionization gauges are located in different regions of the beam line to monitor the pressure in the chamber (the one located in the detection region should be switched off during data collection). These sensitive pressure gauges can be effectively operated in the region ∼10−10 - ∼10−3 mBar. The electrons generated from thermionic emission of the filament are accelerated (by ∼700 volts) with sufficient kinetic energy to ionize any background molecules. The ions created are then attracted to and measured at an appropriately biased electrode. Hence for a fixed temperature the current measured is proportional to the number of incident ions and the pressure as shown in equation 3.2. I + ∝ I −P (3.2) However, hot-filament ionization gauges must be carefully calibrated as each different molecular species in the chamber has a different ionization cross section which could bias the measurement. 3.3 Wien Velocity Filter 69 Figure 3.3: (a) Top view of the Wein filter illustrating the operation principle, where m is the mass of the target ion [?]. (b) Wein filter viewed from the beam entrance. 3.3 Wien Velocity Filter A Wien filter is a mass selection device comprised of a magnet and a pair of electrostatic deflection plates. They are assembled so that the electric and magnetic field lines are directed perpendicular to each other, as shown in figure 3.3 (b). It operates on the principle that an ion beam of velocity v, passing through the center, will be deflected in one direction by the electrostatic field, and in another by the magnetic field. The force experienced by an ion traveling perpendicular to a magnetic field, causing it to deflect, is given by, FB = Bq × v (3.3) FE = Eq (3.4) and similarly for an electric field, Where B and E are the magnetic and electric field strengths, and v and q are the velocity and charge of the ion respectively. However, if FB and FE are of the same magnitude, the two forces will cancel and 3.3 Wien Velocity Filter 70 the ion will experience no net force and traverse unperturbed through the two crossed fields and straight through the filter. It follows from equation 3.5 qE = qvB ⇒ v0 = E B (3.5) that this condition is achieved for an ion with a velocity v0 equal to the ratio of the electric to magnetic field. However, ions with a velocity v0 which is not equal to this ratio experience a non-zero force and are deflected away from the exit of the filter. The kinetic energy of the extracted ions is defined by the source potential, but the velocity of each species differs according to equation 3.6 below: 1 (2qE) 2 v= m (3.6) During operation the electric field is kept constant and the magnetic field is tuned to allow the desired ion to pass through the aperture. This is an important characteristic for the work described in this thesis as switching between hydrogenic isotopes is an integral part (see section 4.1.5). The H2 + , HD+ and D2 + ions are created simultaneously in the source with the same energy. A small electric field (16 V) can be applied to spatially separate the isotopes and the desired species can then be alternated by switching the magnetic field of the Wien filter. During the experiment the magnetic field is automatically changed periodically using a Labview program. This is timed in accordance with the pulse shaper (see section 3.13) so it can be assured that each isotope are interacted with the same pulses, thus accounting for any long term laser drifts. These experiments are typically run over a duration of days and the magnet may undergo some hysteresis. This acts to reduce the ion beam current, and slight modifications to the magnetic field may be required to counteract this effect and maintain maximal ion current. The major advantage of using a Wien filter as a mass selection device as opposed to a magnet in this setup, is that the transmitted ions are not subjected to spatial dispersion and exit through the aperture on the same trajectory. It follows that since the remaining ion optic elements are electrostatic, it is possible to focus all three 3.4 Ion Beam Manipulation 71 Figure 3.4: (a) Schematic of an operating Einzel lens. (b) Simion simulation of the saddle potential created by an Einzel lens. isotopes into the interaction region using the same conditions. 3.4 Ion Beam Manipulation To achieve a measurable rate of dissociation events the extracted and mass selected ions are manipulated using a series of electrostatic ion optic elements to form a well collimated, pulsed beam of sufficient current density in the interaction region (see section 3.6). Although roughly 300 nA of each hydrogenic species exits the Wien filter, subsequent to traversing the beam line optics described below, typically only ∼0.5 nA remains. 3.4 Ion Beam Manipulation 3.4.1 72 Einzel Lens and Deflectors An einzel lens is an electrostatic focusing element comprised of three cylindrically symmetric electrodes (cylinders, rectangular prisms or plates) assembled in series, as shown in figure 3.4 (a). The two outer electrodes are held at the same potential (usually grounded) whilst a different (positive or negative) voltage is applied to the central electrode. The electric field lines disseminate from the central electrode to the two outer ones and create a saddle like potential energy surface (see figure 3.4 (b)). The ions traverse the center of these electrodes and those on the outer edges will experience a greater deflection towards the central axis. In this way, all the ions will converge on the axis at a focal distance f, which is determined by the applied voltage and energy of the beam. The symmetry of the device ensures that the ions will regain their initial energy as they exit the lens, hence the translational energy of the ions is unaltered. Located directly after the Einzel lens are a set of horizontal and vertical deflector plates. These act to guide as many ions as possible into the entrance aperture of the 90◦ quadrupole deflector (see section 3.4.1). In this configuration the horizontal and vertical deflector plates act independently. One plate in each direction is grounded, whilst a voltage (+ve or -ve) is applied to its constituent counterpart to generate an electric field which will deflect the ions by an appropriate amount in a given direction. Ion Beam Chopping A pulsed ion beam is used to reduce the rate of scattered particles from the undissociated beam hitting the detector. It also reduces the likelihood of exposing the detector to a DC beam and causing permanent damage. An ion beam chopper (deflector) composed of two vertical parallel plates is used to generate bunches of ions (typically 350 µs). When the DC voltage is applied the ions are deflected away from the entrance of the 90◦ quadrupole deflector (see 3.4.1). However, the voltage is switched off at a time toff synchronized with the laser (photdiode). 3.4 Ion Beam Manipulation 73 Figure 3.5: The SIMION geometry configuration of a 90◦ quadrupole deflector and the simulated trajectory of an ion beam as it propagates through the two-dimensional electrostatic quadrupole field. This enables an ion bunch which is temporally overlapped with the laser within the interaction region to propagate along the beam line (see section 3.9.3). Two-Dimensional 90◦ Quadrupole Deflector The ion beam must be deflected through 90◦ with respect to the initial propagation direction to reach the interaction region. This is done using a two-dimensional 90◦ quadrupole deflector. This device is composed of four circular electrodes arranged in a square (see figure 3.5). The shim electrodes are installed to produce hyperbolic equipotentials and are used to control the fringing fields which may cause significant abberations to the ion beam profile. The arrangement of +ve and -ve voltage pairs creates a two-dimensional electrostatic quadrupole field between the four rods. An ion beam of energy Ubeam enters the deflector on a straight trajectory between two adjacent rods through a grounded rectangular 3.5 Quadrupole Triplet Focusing Lens 74 aperture strategically placed to reduce any fringing effects (see figure 3.5). The combination of electrostatic and centrifugal forces acting on the ion beam cause it to bend by 90◦ around one of the circular electrodes in a trajectory of radius R0 given by equation 3.7 below. qE = mv 2 R0 or E(V m−1 ) = 2Ubeam (eV ) R0 (m) (3.7) Where E is the electric field strength, and v, m and q are the velocity, mass and charge of the ion respectively. Furthermore, the curvature of the two-dimensional electrostatic quadrupole field lines causes the ions to experience a focusing effect in the y direction and can lead to an astigmatic beam profile. This electrostatic device will not be subjected to hysteresis and can be used to deflect all isotope beams of equivalent energy along the same trajectory. It is also particularly advantageous as it facilitates differential pumping and removes any neutrals transmitted from the ion source. 3.5 Quadrupole Triplet Focusing Lens A single quadrupole lens is comprised of four parallel, equi-spaced, hyperbolic rods arranged in a square, with each adjacent pair held at the same potential but alternate polarity. The ions traverse through the center of these rods and the magnitude of the electric field varies proportially to the distance from the central axis (V =0) (see figure 3.6 (a)). This configuration effectively creates a linear lens for a cylindrical geometry. The potential at any location within this field can be determined from V (x, y) = V0 (x2 −y 2 ), b2 where b is the radius of the rod, and V0 its potential. The force exerted on the ions is perpendicular to the equipotential lines and the velocity components given to the ions cause them to focus in either of the two planes of symmetry (y-z and the x-z). Hence, the quadrupole acts as a perfect converging lens in one symmetry plane and a diverging lens in the other. To overcome this problem and achieve a symmetrically focused beam, two or more of these devices, with reversed polarities, can be arranged 3.5 Quadrupole Triplet Focusing Lens 75 Figure 3.6: (a) Electrodes and equipotential lines in an electrostatic quadrupole. Adapted from [91]. (b) Focusing of a positive ion by a single quadrupole, where f is the focal length given by the z value at which the ion crosses the axis (y = 0). Adapted [92]. in close succession. The focal length of such a configuration can be found from transfer matrices. In a quadrupole triplet, the two outer lenses are of the same length, L1 , with a slightly longer middle component L2 (see figure 3.7 (a)). As the ion beam traverses the first quadrupole it will undergo diversion in one plane which will subsequently be counteracted by the restoring force in the next lens, resulting in an overall converging effect. Furthermore, an additional steering force can be incorporated if the voltage difference between the two positive and negative electrode pairs of the last element are not the same. 3.6 Spectrometer 76 Figure 3.7: (a) Longitudinal section through an electrostatic triplet quadrupole lens. Charged particles move in the z direction from the entrance plane (position 0) to the exit plane (position 5) and beyond adapted from [93]. (b) Focusing of a positive ion by a single quadrupole adapted from [92]. 3.6 Spectrometer The breakup momentum of the ions in the transverse and longitudinal directions cause their relative positions and time of flight to the detector to differ, respectively. The laser beam is focused perpendicularly onto the ions within a spectrometer which creates a weak, uniform static electric field along the ion beam direction (z). This field acts to accelerate the ions with respect to their neutral counterparts. This enhances the temporal separation (z axis) between the fragments according to their mass-to-charge ratio and reduces the distance from the center of mass in the spatial axes (x,y) for the faster fragment. In this way, the ionization and dissociation channels can be measured simultaneously and unambiguously identified, even if an overlap in kinetic energy release of the two channels occurs. The spectrometer is composed of 14 concentric ring electrodes (1 mm thickness) arranged in series with a 5 mm separation, as shown in figure 3.8. The first and last electrodes are grounded and a voltage Vs (typically 600V for H+ 2 to create a ∼150 3.6 Spectrometer 77 Figure 3.8: Schematic diagram of the spectrometer (ariel view) where the ion beam propagation direction and the direction of the electric field are shown by the blue and purple arrows respectively. Also marked is the location of the two circular apertures (yellow rectangles) and the interaction point z0 where V = 0.85Vs and the laser beam is perpendicularly focused ont to the ions. fs temporal separation) is applied to the fourth electrode. A resistor chain serves to connect the electrodes either side of Vs and divide the voltage such that a uniform potential drop per unit length is created. Two thin foil circular apertures which are 1.5 mm and 2 mm in diameter are mounted on the first and fourth electrode rings respectively. These not only act as a beam alignment tool, but help reduce scattering and create a more uniform electric field. The interaction point z0 (origin of z-axis(z=0)) is selected to be midway between electrodes 6 and 7 (see figure 3.6) where the voltage is equal to 0.85 Vs and the electric field is uniform. In reality however, due to the finite aperture of the last electrode the effective electric field extends beyond the physical spectrometer dimensions and does not drop linearly to zero. Furthermore, the radial electric field behaves as a magnifying glass causing dispersion of the ions from the spectrometer axis but is compensated for in the analysis procedure. 3.7 Ion Beam Alignment Protocol 3.7 78 Ion Beam Alignment Protocol The kinematics of the dissociation process are calculated from the time and position information of the the ion and neutral fragments. This imaging technique is limited to measurements of KER above some minimum value (∼ 0.1 eV) as fragments with low transverse velocity cannot be cleanly separated from the undissociated beam and are captured by the small Faraday cup FC4 (see figure 3.6). The formation of a well collimated beam is imperative to reduce the rate of scattered particles hitting the detector and increase the signal-to-noise ratio (SNR). This in turn decreases the probability of false coincidences. A histogram of the time of flight (TOF) of the photodissociation fragments is recorded as a live image during the ion beam alignment procedure, as shown in figure 3.9 (a). This is used to verify that the timing sequence (see 3.9.3) has been set correctly and spatial overlap between the ion beam and the focused laser has been achieved. Furthermore, the coincidence TOF density plot (see figure 3.9 (b)) can be used to estimate P Counts the SNR and the dissociation rate for a specific channel. Consequently, the Time overlap between the ion beam and laser can be obtained by scanning the laser in the y direction to find the highest dissociation rate. 3.7.1 Collimation The ion beam current is initially optimized on Faraday cup 3 (FC3 see figure 3.1) which is located after the spectrometer. It is positioned along the same line as the center of the detector and the two circular apertures of the spectrometer. The apertures are used as a tool to collimate the ion beam. Once the current has been optimized the spatial profile is reduced in a symmetrical manner to a cross section of ∼0.6 × 0.6 mm2 using the set of four jaw slits denoted slits 1 (see figure 3.1). To improve the signal to noise ratio the scatter incident on the detector must be minimized. The ion beam viewer described in section 3.7.2 can be used to visualize the position and extent of the scatter. Most importantly, the small Faraday cup (FC4 ) 3.7 Ion Beam Alignment Protocol 79 Figure 3.9: (a) Histogram of the time of flight (TOF) of H+ 2 photodissociation fragments. (b) Coincidence TOF spectra for the different fragmentation channels of CD+ induced by intense laser pulses. Where t1 and t2 is the TOF of the first and second fragment respectively. located in front of the detector must be positioned correctly. Following this, slits 2 can be closed to collect any scatter induced from the closure of slits 1. The two ways this can be done is outlined below, the latter of which is usually the most effective. 1. Each jaw on slits 2 was closed independently until they cut into the ion beam. They were then slowly retracted to the point that the initial current can be retrieved. If required these positions can be adjusted by monitoring the scatter on the ion beam viewer. Any charging effects can be corrected by opening the corresponding jaw further. An indication of how well collimate dthe ion beam is can be obtained by comparing how symmetrically slits 2 are closed around the center of slits 1. 2. By observing the coincidence TOF density plot, each jaw can be fully closed and then opened very slowly to the point where the compromise between the SNR and dissociation rate is operational. 3.7 Ion Beam Alignment Protocol 80 Figure 3.10: Schematic diagram showing the honeycomb structure of an MCP and the electron avalanche created by an electron propagating through one of the individual glass capillary electron multipliers. 3.7.2 Ion Beam Imaging A multi-channel plate (MCP) is a specially fabricated plate which can amplify an incoming signal due to particles or radiation. It is composed of an array of several individual electron multipliers arranged in a honeycomb structure as shown in figure 3.10. Each electron multiplier is effectively a thin glass capillary of very small diameter (roughly 10 micrometers) internally coated with a high resistance, low work function material. If an ion of sufficient energy strikes the inner wall of the capillary it will liberate secondary electrons. These electrons are then accelerated towards the exit due to the potential difference (approx ∼ 1.8 kV ) applied along the length of the capillary. As the electrons propagate toward the exit they repeatedly hit adjacent walls creating a charge cloud of approximately 106 electrons at the outlet. The capillaries are mounted with a slight angle of impact (roughly 8◦ to the plate) and one method of increasing the degree of electron multiplication is to produce a chevron (v-like) shape by positioning two plates back to back. The electron cloud which exits the first plate subsequently initiates the cascade in the next. The overall amplification achieved depends on the applied voltage and geometry of the micro-channel plate. The structure of the MCP allows the position of the incident ion to be transformed into a spatially well defined charge cloud that can be accurately measured using a phosphorus screen or delay-line anode. Overall, MCP’s are highly desirable as they can provide high gain signals with great spatial and temporal resolution. Furthermore, despite a dead time being associated with each individual channel electron multiplier, 3.7 Ion Beam Alignment Protocol 81 Figure 3.11: Schematic diagram of ion beam viewer technique the multiplicity of this device makes it capable of handling multihit events. The ion beam viewer is composed of a 40 mm diameter MCP positioned in front of a phosphorus screen readout device and mounted at 45◦ to the ion beam axis (see figure 3.11). The phosphorus screen is biased (+ 3.6 kV) and the electron clouds emerging from the outlet of each individual electron multiplier are accelerated towards it. The kinetic energy of each impinging charge cloud releases an intense green photon from the phosphor screen. This fluorescence can be captured by a charge coupled device (CCD) connected to a television screen enabling live images to be obtained. The whole readout device is mounted on a verticle manipulator which when lowered terminates behind FC4 allowing it to be accurately positioned. 3.7.3 Ion Beam Current Measurement During the experiment the undissociated ion beam is collected and measured using a metal receptacle known as a Faraday cup (FC). The FC is composed of an electrically conducting hollow cup (diameter 2mm) which is isolated from the thin metallic rod which acts to hold it in place. In this way, only the ions in the cup itself contribute to the measurement. The whole FC device is located at a distance of 20 cm from the front of the detector and is mounted at a 45◦ angle to the vertical direction using a movable x,y and z manipulator. 3.8 Hexanode Delay-Line Detector 82 Figure 3.12: The HEX80 delay-line detector by RoentDek Gmbh (a) Schematic of the delay-line wire array construction, illustrating the 60◦ angle between each dimension (u, v and w), which consists of a pair of parallel wires (reference (red) and anode (green)), between which is a potential difference (+ 90 V). (b) The delay-line detector assembled for use in vacuumn. An ammeter can be used to measure the electrical current induced for a DC ion beam. However, the substantially lower ion densities for a bunch of ions in a pulsed ion beam means that a boxcar integrator must be used instead. The boxcar integrator is a sampling instrument which integrates an amplified input signal for a predefined gatewidth after an applied trigger. This integrated signal is then averaged over a number of cycles. 3.8 Hexanode Delay-Line Detector There is currently no single device which can provide both timing and position information simultaneously. However, the combination of a multichannel plate (MCP) and a helical delay-line anode can be used to perform kinematically complete measurements of dissociating molecules. If a multichannel plate (MCP) is used to amplify the signal of an incident fragment then the spatial information is preserved and the localized charge cloud which exits the back side of an MCP can be collected on a helical delay-line wire arrangement and 3.8 Hexanode Delay-Line Detector 83 used to obtain two dimensional position information. The RoentDek hexanode delay line detector consists of three layers (u, v and w) of wire pairs (reference and anode) wound in parallel around two ceramic rods (see figure 3.12). The layers are oriented at 60◦ relative to each other and separated vertically by 1mm. The charge cloud emerging from the MCP is distributed equally between each layer and the position x (from the center of the wire) of the incident fragment for that respective dimension, is defined by the center of the charge distribution. The inter-wire voltage difference ensures the anode (+ 90 V) acts as the collector, whilst the reference serves as a background indicator. The difference between these two signals is determined using an RF pulse transformer (AMP TP-104) resulting in a comparatively clear signal. The induced signals will traverse parallel to the wire with velocity vpar which is close to the speed of light in vacuum. However, due to the helical construction of each layer of wire pairs, an effective propagation speed vper is introduced as the signal is required to traverse 250 mm of wire around a loop in order to cover a 1 mm distance in the transverse direction (typical values of signal on the anode 0.7 mm/ns). Since vper is geometry dependent, a calibration procedure is required to determine the correct value for each respective dimension (see section 3.21). The delay line technique relies on the relative delay experienced by a propagating signal as it traverses to adjacent ends of a wire ((u1 u2 ), (v1 v2 ), (w1 w2 )). The time required for a signal to travel from position x to one end of the wire is given by u1 = lx + x vper (3.8) and similarly the time taken to reach the opposite end of the wire (where lx is the half-length of the wire). u2 = lx − x vper (3.9) Therefore the position of a signal in the hexagonal frame, along a particular wire (in 3.8 Hexanode Delay-Line Detector 84 this case u) can be determined from: u= vper × (u1 − u2 ) 2 (3.10) The two-dimensional information can be achieved by reading a second wire in coincidence. Despite having three wire pairs, only two combinations are required to convert from the hexagonal frame to the Cartesian coordinate system. This can be done using the equations given below, where a(u), b(v) and c(w) are the calibration factors described in section 3.21. xuw = u.a(u) yuw = √1 (2w.c(w) 3 − u.a(u)) xuw = u.a(u) yuw = √1 (2w.c(w) 3 − u.a(u)) xvw = (v.b(v) + w.c(w)) yvw = √1 (w.c(w) 3 − v.b(v)) In practice, the signals from the third wire are a redundant source of information that can be used to reconstruct signals lost due to electronic dead-time and inadequate electronic threshold conditions. The variables measured for each dissociation event are the timing ( t1 , t2 ) and spatial (x1 , y1 , x2 , y2 ) components for both fragments. 3.8.1 Hexanode Delay-Line Detector Calibration The exact geometry of each Hexanode Delay-Line Detector manufactured may differ slightly. For this reason, each detector must be independently calibrated to determine the effective propagation speed vper for each respective dimension (u, v and w). It is imperative that this value is determined correctly as it scales the timing signals to a well defined spatial coordinate system and plays a critical role in the quality of the resolution of the images. An image of a calibration mask mounted ∼ 5 mm in front of the MCP of the detector, comprising of 0.25 mm holes arranged linearly with 5 mm spacing, is generated by irradiating it with a uniform distribution of H+ 2 ions (see figure 3.13). If a sufficiently 3.8 Hexanode Delay-Line Detector 85 Figure 3.13: Image of the calibration mask, as measured using the uw wire pair of the delay line detector and the calibration parameters as defined in equations 3.12 and 3.14. The dashed lines are to placed to guide the eyes to straight lines. high voltage (+ 4 keV) is applied to the Einzel lens (see section 3.4.1) the traversing ions will be very tightly focused at some focal length f, on the exit of the 90◦ quad deflector. These ions will then disperse as they propagate through the unobstructed, field free beam line (the deflector and spectrometer were removed from the beam line) and strike the detector. The 2 m distance traversed is adequate to minimize the divergence angle of the impinging ions. Since the physical length (∼ 20 m) and propagation velocity (close to the speed of light) through each wire is invariant, the overall travel time of a signal within each wire (∼ 100 ns) is a constant Tsum and given by the ‘sum rule’ below: Tsum = tright + tlef t − 2tmcp (3.11) Where tright and tlef t are the signals measured on adjacent ends of each wire respectively, and tmcp is measured from the MCP. In practice, Tsum was found to be position 3.8 Hexanode Delay-Line Detector 86 Figure 3.14: The‘sum rule’ as calculated from equation 3.12 for the measured calibration data for u, v and w wires corresponding to a, b and c respectively. dependent, as shown in figure 3.14. The ‘sum rule’ can be used to correlate ‘good’ signals and eradicate any undesirable background data introduced as a result of ringing and electronic noise. Any data which lies outside the two parallel, horizontal, red dashed lines in figure 3.14 is excluded. To determine the vper for each respective dimension (u, v and w) two calibration procedures were completed. One relies on an unconstrained, non-linear optimization fitting procedure (in Matlab) to optimize the cross-correlation of a background reduced image of the mask measured from the detector and an accurately scaled replica. A linear correction term was introduced to the calibration parameters as Tsum is position dependent. The scaling and linear correction parameters of the experimental image were then varied to align these two matrices. The algorithm converged on the minimized value for the cross-correlatation once the appropriate calibration parameters had been obtained. In the second procedure a 2D Gaussian was fitted to the signal from each hole of the background reduced image of the mask measured from the detector and a non-linear least squares fitting procedure was used to match it to an accurately scaled replica. 3.9 Data Acquisition and Electronics 87 The optimal calibration factors for each dimension were found and are substituted into equations 3.11 to convert from the hexagonal frame to the Cartesian coordinate system. 3.9 a(u) = 0.3660 × (1 − u · 3.0 × 10−4 ) (3.12) b(v) = 0.3791 × (1 − v · 3.0 × 10−4 ) (3.13) c(w) = 0.3537 × (1 − w · 3.0 × 10−4 ) (3.14) Data Acquisition and Electronics The time of flight (TOF) measured for each fragment is initiated at the dissociation event and terminated as the fragments strike the MCP. Such measurements can only be achieved using a sophisticated timing sequence. In this setup, all internal and external triggers are controlled using a labview program and are initiated every laser shot (see section 3.9.3 for details). This timing scheme ensures temporal overlap of the target ions and the laser in the interaction region. Furthermore, it facilitates the the relevant data to be recoded with minimal background contributions. The signals obtained from the Hexanode delay-line detector are temporal readouts. They are initially amplified (Ortec Fast Amp) before being processed by a constant fraction discriminator (CFD Ortec 845) and then measured using a multi-hit time to digital converter (TDC) which is then read using a labview program. 3.9.1 Constant Fraction Discriminator (CFD) Discriminators can be used as a means of reducing background and non desirable effects such as electronic ringing. They operate by generating precise logic pulses in response to input signals exceeding a given threshold. In a leading edge discriminator the output pulse corresponds to the time at which the input pulse crosses a given threshold voltage. This method however can lead to a shift known as a ‘time walk’ between the timings of pulses with equivalent rise times but differing amplitudes. Figure 3.15 (a) illustrates 3.9 Data Acquisition and Electronics 88 this effect where it is evident that the pulse of smaller amplituted crosses the threshold at a later time. Where accurate timing signals are imperative a constant fraction discriminator (CFD) can be used to alleviate this problem. The amplified input signal is split into two and one half is attenuated to a certain fraction of the original amplitude while the other is inverted and then delayed by an amount determined by the cable length. The addition of these two constituent parts creates a bipolar signal which produces an output pulse as it crosses the baseline. This zero crossing time tcross is effectively given by equation 3.15 below, tcross = td (1 − f ) (3.15) where td is the delay and f the fraction by which it has been attenuated, hence tcross is always independent of amplitude (see figure 3.15 (b)). Figure 3.15: (a) In a leading edge discriminator two input pulses (dotted lines) of equivalent rise time but differing amplitudes will trigger at dissimilar times tHigh (blue) and tLow (green) as they cross the threshold (solid yellow line). This time difference is known as a time walk and can be alleviated through use of a constant fraction discriminator (CFD). This operation causes each pulse (solid black line) to cross the baseline tcross (solid red line) and hence trigger simultaneously, regardless of amplitude. (b) Outlines the principle of the CFD by illustrating the input (dash line), delayed (dotted) and attenuated and inverted (dash-dot) pulses and then finally the resulting bipolar pulse (solid line) created upon addition of the latter two pulses. The output pulse of the CFD occurs at tcross (solid red line) as the bipolar pulse crosses the baseline. 3.9 Data Acquisition and Electronics 3.9.2 89 Time to Digital Converter (TDC) For each event, the output pulses from each channel of the CFD (t1 , u1 , v1 , w1 and t2 , u2 , v2 , w2 ) are fed into a separate channel of the 16-channel time to digital converter (TDC). This device provides a digital representation for the time of arrival of an incoming pulse with respect to a well defined reference. In this case the laser is used as the start trigger, effectively defining the interaction as t=0 and then the time denoted by the TDC is given by the the time of flight of the fragment to the detector (i.e. from CFD) plus some time associated with the electronics (td see section 3.10). This information is then read and saved by a labview program. 3.9.3 Timing Sequence The timing sequence for the electronics and data acquisition associated with this experiment is illustrated in a schematic diagram in figure 3.16. A signal from each incoming laser pulse is obtained by focusing the leakage from a dielectric mirror onto a photodiode (see figure 3.1). This is then used to initiate a labview program which provides the required delays (tdelay ) and gating widths (tgate ) for all components involved. Each incident pulse redefines the time as t=0 and effectively signifies the time of the interaction. The chopper delay is syncronised with the previous pulse. The repetition rate of this timing sequence is 1 kHz. There are three components for which the delay and timing gate width is controlled using the labview program: Chopper The delay between the trigger from the laser and the chopper is such that the subsequent laser pulse and ions are temporally overlapped and reach the interaction region simultaneously. A continuous voltage is applied to the chopper deflecting the ions elsewhere until it is switched off at a time tdelay later for a time duration tgate permitting a well defined bunch of ions to proceed to the interaction region. CFD The CFD is vetoed so that the only signals processed are those which arrive after tdelay but within tgate . Thus the relevant species are incorporated within 3.9 Data Acquisition and Electronics 90 Figure 3.16: A schematic diagram of the timing sequence for the data acquisition. Every incoming laser pulse activates the labview program which redefines a zero time (t=0) and assigns the correct time delay to each component. The first laser pulse will not be recorded as an event as the ions and laser pulses will not be simultaneously present in the interaction region. However, the subsequent timing sequence will be repeated on a 1kHz rate for the duration of the experiment. tgate due to their expected time of flight (ToF) with respect to the trigger. In this way the surrounding background signals recorded are minimized. TDC The output signals sent from the CFD to the TDC are no continually read. Instead the TCD is triggered after all events from each pulse is collected and sent to the VME computer to be read by a labview program. 3.10 Measuring T 0 3.10 91 Measuring T 0 The measured times of flight (TOFi ) are shifted from their actual values (ti ) by a constant delay (t0 ) caused by the signal propagation time through their associated electronics (mainly cable lengths). Thus the actual time of flight of each fragment is determined by: ti = T OFi − t0 (3.16) In practice (t0 ) is measured at the end of each experimental run. The parabolic mirror is moved parallel to the ion beam (z-direction) using a translation stage. Eventually the incident photons strike one of the spectrometer electrodes and they are reflected toward the detector. The value for t0 can then be determined from the histogram of the time of flight (TOF) of the photodissociation fragments (see section 3.7). 3.11 Resolution The accuracy of the measured timing signals from the opposite ends of each delay-line wire are position dependent. The length of each wire is ∼ 20 m which equates to an overall signal propagation time of ∼ 100 ns. Those signals which travel a greater distance along the wire may undergo a broadening effect. Since the CFD relies on pulses of similar shape, this may have a slight adverse affect on the times measured. There is however no alternative solution to providing exact timing signals independent of amplitude (see section 3.9.1). The precision of the calibration is also a critical factor as it scales the signals to a well defined spatial coordinate system whilst correcting for any non-linear effects. The resolution obtained directly from the multi-channel plates and the delay-line wire signals is ∼ 0.3 ns and ∼ 0.5 ns respectively. This results in an overall energy resolution of 25 meV in the measured kinetic energy release (KER) spectra. 3.12 Femtosecond Laser System 92 Figure 3.17: The Femtosecond Laser System (adapted from Femtolasers Gmbh). 3.12 Femtosecond Laser System The Fourier transform limited femtosecond pulses used for the dissociation experiments are 33 fs in duration with a spectral FWHM of 40 nm centered around 795 nm. The pulses have an energy of approximately 1.5 mJ and are delivered at a repetition rate of 1 kHz. Such pulses are created using a Ti:Sapphire based multi-pass amplifier (Femtolasers Gmbh) system consisting of three main components; an oscillator, amplifier and compressor. These pulses are shaped using a conventional 4-f phase-only pulse shaper with a programmable liquid crystal spatial light modulator (Jenoptik Phase SLM-640) situated in the Fourier plane (see section 3.13). The laser polarization is rotated to the vertical direction and the laser beam is focused using a 20cm off-axis cylindrical mirror (section 3.15) through an anti-reflective (AR) coated glass window onto the ion beam target. 3.12 Femtosecond Laser System 3.12.1 93 Oscillator Typically, an oscillator configuration (see figure 3.17 (a)) is comprised of a gain medium (with a specific emission spectrum) positioned between an end mirror and an output coupler. The gain medium is then optically pumped causing a large range of frequencies to resonate within the cavity. A specific phase relationship between the stable modes (mode-locking) must then be established to amplify the signal and create a Fourier transform limited pulse (minimum temporal length). Other than the achievable pulse width, stability and reproducibility are the most important demands on a practical femtosecond source. The femtosecond pulses at WIS are generated in a Kerr lens mode-locked Ti:Sapphire oscillator [94] which is pumped by a diode-pumped solid-state laser (Spectra-Physics Millenia V). The seed pulses created are 10 fs and centred around 795 nm with a spectral FWHM of 100 nm. They are produced at a repetition rate of 75 MHz with average output power of ∼400 mW when the medium is pumped with 4.2 W, thus creating 5 nJ pulses. 3.12.2 Amplifier This seed beam is then passed through the amplifier (see figure 3.17 (b)) amplified using the chirped pulse amplification (CPA) technique [95]. To avoid generating very high peak powers which can damage the optics and introduce nonlinear distortions to the spatial and temporal profile of the beam during the amplification process, the pulses are first temporally stretched from 10 fs - 10ps. The power can be reduced by three orders of magnitude by using dispersive optics and third order (TOD) mirrors to create a frequency-chirped pulse and introduce high-order phase terms. This elongated beam then enters a multi-pass configuration. It is amplified via a simulated emmison process at each pass through the Ti:Sapphire crystal which is pumped by 11W from a Nd:YLF laser (Spectra-Physics Empower 30). After traversing the gain medium 4 times, a single pulse from the 75-MHz pulse train is picked off by a Pockels cell every millisecond, a rate which is in unison with the pump laser (1kHz). These selected pulses 3.13 Pulse Shaper 94 then undergo 5 more passes of the Ti:Sapphire crystal to achieve a typical energy of 1.5 mJ at the exit of the multipass stage. The spectral bandwidth of the pulse is reduced from 100 nm to 50 nm due to gain narrowing and some wavelength mismatch of the optics in the amplifier stage. 3.12.3 Compressor Optical pulse compression relies on accurately reversing the stretch factors induced in the amplifier. The objective is to create an overall zero group-delay dispersion and create Fourier transform limited pulses (FTL). There are 4 low-dispersion Schott prisms located in the compressor which applies negative TOD to counteract the positive TOD which is introduced by previous optical components in the system (materials and mirrors)[96]. The compressed and amplified beam consists of pulses 33 fs in duration with a repetition rate 1 KHz around 795 nm. 3.13 Pulse Shaper The pulses are shaped using a 4-f phase-only pulse shaper with a programmable liquid crystal spatial light modulator situated in the Fourier plane. This configuration is comprised of a combination of gratings and lenses (or mirrors) assembled as shown in figure 3.18. In the 4f-line configuration the spectral components of the incident pulse are angularly dispersed by the first set of gratings (10001/mm). This creates several spatially separated groups of quasimonochromatic waves coupled into a given direction. The first 20 cm cylindrical lens is then used to focus each spectral group to a small diffraction spot at a specific position on the Fourier plane (i.e. creating a spatio-temporal coupling). Any mask located in the Fourier plane can be used to manipulate the spectral components of the pulse. These components are then recombined and recollimated by the second lens and grating combination in the second half of the 4f-line, fabricating the desired pulse shapes. 3.13 Pulse Shaper 95 Figure 3.18: A schematic diagram illustrating the principle of the 4f-line. The first grating disperses each frequency in a given direction and the lens maps it to a given position in the Fourier plane (i.e. creating a spatio-temporal coupling). In the second half of the 4f-line all the components are recombined. An aperture placed in the Fourier plane can be used to reduce the spectral bandwidth of the pulse, creating what is termed a ‘narrowband’ pulse. The interdependence of laser pulse parameters means this increases the temporal duration. For a Fourier transform limited pulse the temporal duration is inversely proportional to the spectral width. These ‘narrowband’ pulses demonstrate an enhanced resolution in the data obtained (see section 4.1.2). Alternatively, to accurately introduce higher orders of dispersion a programmable liquid crystal spatial light modulator was used as a mask. The pixels are constructed from a thin layer (9 µm) of nematic liquid crystals enclosed within two glass substrates. One substrate is coated with a thin layer of transparent ITO (indium tin oxide) as this material is electrically conducting. The entire LC-SLM contains 640 of these pixels arranged parallel to each other. An independent voltage can be applied to each individual pixel which causes the rod-like molecules to orient themselves with the corresponding electric field (see figure 3.20 (a)). This results in a change in the refractive index for light polarized in the y direction. This retards the spectrally dispersed components 3.13 Pulse Shaper 96 Figure 3.19: Schematic diagram showing how the rod-like molecules in each pixel of a liquid-crystal spatial light modulator (LC-SLM) reorient themselves along the direction of the applied electric field. according to wavelength. The performance and accuracy of this configuration depends on careful alignment, and the quality of the phase voltage and frequency φ(ω, U ) calibration. A correct calibration ensures an appropriate retardation is applied for the corresponding wavelength. The 50 nm bandwidth of the incident pulses means only 150 pixels are actually active during the experiment. The pixel resolution is 0.34 nm for frequency domain pulse shaping. In principle, linearly chirped pulses (frequency sweep) can be generated by inserting additional glass into the optical path. However, this technique provides no means of compensating for higher orders of dispersion. To accurately shape the pulses which will propagate the interaction region are first aligned through a replica AR coated window into a GRENOUILLE (see section 3.14). The corresponding frequency-time relation is used as a tool to eliminate any undesirable higher order dispersion terms introduced due to self phase modulation from various downstream optical components and create an FTL pulse. The desired spectral phase was then applied to this FTL. Furthermore, the programmable nature of the LC-SLM allows the pulse shape to be alternated periodically throughout the experiment (typically every 2 minutes). This allowed drift biases in the laser or the ion beam pointing stability to be compensated 3.13 Pulse Shaper 97 Figure 3.20: Schematic illustration of shaping the temporal profile of an ultrashort laser pulse by retardation of the spectrally dispersed individual wavelength components in a phase only LC-SLM. for. 3.13.1 Determining the Central Pixel 00 000 The quadratic (linear chirp, φ ) and cubic (TOD, φ ) spectral phase functions used to produce the shaped pulses described in this thesis can be expressed by certain terms of the Taylor expansion given in equation 3.17 and is discussed in detail in section 2.12. 1 00 1 000 0 φ(ω) = φ(ω0 ) + φ (ω0 )(ω − ω0 ) + φ (ω0 )(ω − ω0 )2 + φ (ω0 )(ω − ω0 )3 + . . . 2 6 (3.17) Equation 3.17 is incorporated into an algorithm in Labview and used to supply the correct voltage to each pixel in the 640 LC-SLM arrangement. This ensures that the induced changes in the refractive index of the pixel arrangement retards the spectrally dispersed individual wavelength components of the pulse with respect to each other according to the applied spectral phase function (as illustrated schematically in figure 3.20 (b)). The spectral phase function is applied with respect to some central pixel of the LC-SLM, and it is imperative that the central wavelength of the pulse matches 3.13 Pulse Shaper 98 Figure 3.21: Calibration showing the pixel number to wavelength correlation. this pixel. This is particularly important when applying asymmetric spectral phase functions so the frequency components are temporally redistributed around the central frequency of the pulse, and not an arbitrary one. The central pixel is defined as the one through which the central wavelength of the pulse propagates. This value can be experimentally determined by applying a pi gate (delta function) to a known pixel. The transmitted spectrum (wavelength and intensity) was measured using a highly sensitive spectrometer (ANDO) at the exit of the shaper. All wavelength components of the pulse traverse the LC-SLM unaffected, with the exception of the pi gated pixel. This component is highly dispersed and a clear dip in the spectrum can be observed. This procedure was repeated for several different pixels, and a linear relation with a gradient of 0.34 nm (corresponding to the width of each pixel δλ nm) was obtained (see figure 3.21). The central wavelength of our spectrum was measured as ∼795 nm, which from figure 3.21, implies the central pixel should be set to ∼405, for this particular alignment. To quantify the sensitivity of this effect consider an asymmetric function applied 10 pixels offset from the central wavelength. This corresponds to the phase function 3.14 Laser Pulse Characterization (GRENOUILLE) 99 being ∼3.4 nm off center. Experimental data illustrating the effect an off-centered TOD phase function has on the dissociation of H+ 2 can be seen in section REF. 3.14 Laser Pulse Characterization (GRENOUILLE) An inherent problem with measuring femtosecond pulses is that to measure an event in time, a shorter event is required to measure it. To combat this problem autocorrelation and frequency resolved optical gating (FROG) techniques use the pulse to gate itself. The full time-dependent intensity and phase of the shaped pulses can be retrieved using a GRENOUILLE (grating-eliminated no-nonsense observation of ultrafast incident laser light electric fields) a schematic diagram of which is shown in figure 3.22 [97]. Figure 3.22: (a) Schematic diagram of the GRENOUILLE components and configuration. (b) Schematic diagram illustrating the two pulse replicas (different colors) being focused and crossed into the SH crystal by the Fresnel biprism. (c) Crossing the beams at an angle maps delay onto transverse position. The input pulse is split into two replicas and focused at a mutual angle (crossed) into a thick second harmonic (SH) crystal by a Fresnel biprism (a prism with an apex angle close to 180, see figure 3.22(a)). This causes the light at the center of the laser beam 3.15 Parabolic Mirror and Alignment and Imaging 100 to undergo a greater retardation than the light at the edges (see figure 3.22 (b)). This spatiotemporal overlap creates a relative delay τ between the input E(t) and gated pulse E(t- τ ), and is mapped in the x-direction. The second harmonic signal generated by the overlapped components of the electric field (2E(t)E(t-τ )) exit the crystal on axis and are measured as a function of delay. The signal can be described by following equation, Z 2 iωt S(τ, ω)α E(t)G(t − τ )e dt (3.18) Since the crystal has a relatively small phase-matching bandwidth, the phase matched wavelength produced by the crystal varies with angle and can be used as a spectrometer. The cylindrical lens positioned after the SHG crystal maps the position of each wavelength to a location on the y-axis of the CCD camera. The final trace (i.e., spectrogram) is a 2D map of pulse intensity as a function of delay time and frequency. The advantage of GRENOUILLE devices is that they are compact, affordable and contain zero degrees of alignment. 3.15 Parabolic Mirror and Alignment and Imaging The laser beam must be focused within the interaction region to achieve intensities high enough to induce dissociation. This was done using a 20 cm off-axis parabolic mirror as opposed to a conventional glass lens to avert the onset of any further distortions due to chromatic abberations. Alignment of the off-axis parabolic mirror was initially conducted using a HeNe laser to accurately mark the relevant reference points. The laser beam is aligned through the center of the first two irises (to set the height and tilt of the beam) onto the center of the dielectric mirror (see figure 3.23). It is crucial that the incident laser beam is parallel to the table and centered on the off-axis parabolic mirror. The laser beam is then reflected 90◦ through the center of the AR coated entrance window (the abberations of which have been recompensed using the GRENOUILLE see section 3.14) and focused amid two spectrometer electrodes. The beam then exits the interaction chamber and should terminate (concentric with a closed iris) on a pre-aligned (using HeNe) cross 3.15 Parabolic Mirror and Alignment and Imaging 101 Figure 3.23: (a) Schematic diagram showing the laser being focused into the interaction region. The dashed red line is where the laser beam was diverted into the CCD for imaging purposes. (b) The focusing dimensions of the laser beam corresponding to certain translations of the off-axis parabolic mirror. on the wall. However, unlike a convex lens, the parabolic mirror could potentially introduce astigmatisms if not aligned properly. A glass plate inserted before the chamber entrance window was used to deflect an attenuated beam (ND filters inserted as shown in figure 3.23) into a CCD camera to image the focus. A high neutral density filter (ND = 8) was needed to protect the camera when locating the focus and creates an overall higher sensitivity to the beam profile. To achieve a non-astigmatic beam, the parabolic mirror can be adjusted in the θ and φ directions to correct profiles which are tilted or appear horizontal/vertical, respectively. However, such modifications can cause undesirable changes to the trajectory of the laser beam through the interaction region. To combat this issue the position of the focal image should be kept at a constant position on the 3.16 Z-Scan 102 Figure 3.24: (a) The area of the focused laser beam at various z positions obtained from fitting an ellipse (red line) or a contour (blue line) to the CCD images. (b) The corresponding peak intensity for these z positions and fit type. screen (thus essentially implying constant directionality). In practice this is accomplished by focusing using θ and φ in the appropriate directions according to figure 3.23 (b) and then using x and y manipulators to bring the image back to the same position on the screen. After each iteration the new focus must be found and the procedure repeated until the best focusing conditions are reached. Images of the beam profile are obtained for a range of z positions (where z= 0 is the focus) and an ellipse and contour are fitted around each profile to determine the area. The area is then used to provide an estimate of the intensity of the laser across the various z positions (see figure 3.24). 3.16 Z-Scan The profile of a focused laser pulse varies both spatially and in intensity along the axis of propagation (z-axis) depending on the strength of the focusing lens. The spatial contour defining each specific intensity region In can be described as a ‘peanut’. The volume vn occupied by In is vn minus the volume of all lesser intensities VP i , as shown schematically in figure 3.25 (a). As In increases, the volume occupied by the 3.16 Z-Scan 103 Figure 3.25: (a) Profile of a focused laser pulse showing the isointensity contours bounded by the saturation intensity of ion species A+ , A2+ etc. The slit (ion beam) therefore probes a particular intensity and volume region. Figure adapted from [62]. (b) The measured dissociation rate of H+ 2 as a function of laser position. In contour decreases and tends towards a finite point as I tends to Io. Therefore, at a given position along the focal axis, the laser- ion interactions are unique to that z coordinate. The finite ion beam can effectively be used as an aperture to restrict the interaction region to a cylindrically symmetric thin slice of width ∆z (see figure 3.25 (a)). The laser focus can then then be translated with respect to this aperture enabling ion interactions unique to specific z positions to be probed. The signal measured at each z increment is a compromise between the volume of interaction and range of intensities exposed. This gives rise to the z-scan shown in figure 3.25 (b). In practice, this z-scan is carried out to establish the position of the focus and subsequently determine the z-position of the experimental overlap region to achieve a particular peak intensity. Although more than a single intensity is probed at each z increment, this method allows the lower intensity regions to be appropriately masked from the detector for high intensity measurements. Typically, the pulse shaping experiments were carried out at a 19 mm distance from the laser focus. This region is still within the Rayleigh radius of the laser where the volume is still less than double the beam waist. Chapter 4 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses The long-outstanding ambition of scientists to understand and optically drive chemical reactions to the highest degree of specificity has lead to extraordinary advancements in coherent control strategies. Ultimately, the extent to which molecules can be controlled is governed by how precisely the electric field can be manipulated. The development of generic algorithms has proven a useful technique in finding the specific pulse shape required to optimize a specific chemical process. The complexity of the electric fields created however, makes extracting a mechanistic explanation a formidable task. This chapter presents a systematic study into the interaction of the theoretically tractable H+ 2 and well characterized pulses which can be described analytically. Specifically, 00 Fourier transform limited pulses are shaped via the application of quadratic (ϕ ) and 000 cubic (ϕ ) spectral phase functions described by the corresponding terms of the Taylor expansion given in equation 2.12. The magnitude and sign of these parameters are varied and the ability to use them as a tool to control the photodissociation of H+ 2 is demonstrated and the underlying control mechanisms are identified and supported by theory. 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses 4.1 105 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses 4.1.1 Motivation Momentum imaging techniques can be eloquently used to study how a particular pulse shape influences the fragmentation of a molecule. One form of pulse shaping involves redistributing the frequency components in time according to the group delay imposed by an applied spectral phase function. In turn, the modified temporal profile can be obtained from the inverse Fourier transform (see section 2.3.1). With only two energetically available potential energy curves at 800 nm, H+ 2 is highly theoretically accessible. The interaction of H+ 2 and analytically shaped pulses therefore plays a critical role in gaining an insight into how the instantaneous frequencies and temporal profile of shaped pulses can be used to control dissociation dynamics. Furthermore, 00 000 experimentally the quadratic (ϕ ) and cubic (ϕ ) spectral phase functions used are accurately applied and the shaped pulses well characterized. Polarization-Gate Frog Various geometries of the frequency resolved optical gating (FROG) technique can be used to fully characterize ultrafast pulses. However, the retrieved FROG traces from each variant contain non-trivial differences [98, 99, 73]. The Polarization-Gate (PG) technique is the most intuitive method as the traces do not contain a direction of time ambiguity. Therefore the sign of the spectral phase function can be clearly determined. For this reason PG FROG traces have been simulated to elucidate the time-frequency relationship for the different pulse shapes used in an unequivocal manner. A measured spectrum of the pulse is incorporated and the phase function is applied with the same resolution as the pulse shaper in the experiment. 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses 106 Figure 4.1: (a) Simulated Polarization-Gating FROG traces showing a 33 fs FTL pulse with a 40 nm spectral bandwidth. (b) The corresponding Gaussian temporal profile is normalized to 1 (black line). 4.1.2 Fourier Transform Limited Pulses The shortest, most intense laser pulses a particular laser system can produce are termed Fourier Transform Limited (FTL). The shorter the pulse duration, the larger the spectral width of the pulse. This relationship is given by the time-bandwidth product which can be described in terms of the frequencies of the pulse [74, 73] and is given in equation 4.1 below for an ideal Gaussian pulse: ∆t∆ν = 2ln2 = 0.441 π (4.1) Where ∆ν is the spectral frequency width and ∆t is the temporal duration of the pulse. In an FTL pulse the spectral phase ϕ(ω) is constant and each frequency component ω(t) has an equal probability during the pulse duration (see the PG FROG trace in figure 4.1 (a)). The temporal intensity profile is Gaussian and defined by the full width half maximum (FWHM) of the intensity distribution (see figure 4.1(b)). The addition of any non-linear phase terms results in the inequality ∆t∆ν ≥ 0.441 which stretches the pulse in the time domain and acts to reduce the peak intensity. Before the behavior of H+ 2 in a shaped electric-field can be understood, its response to these simple FTL pulses must be explicated. In the three-dimensional momentum 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses 107 Figure 4.2: The KER-cosθ density plots for FTL pulses with (a) I0 ∼ 1×1012 Wcm−2 and 18 nm bandwidth measured off-focus. (b) I0 ∼ 1×1014 Wcm−2 and 40 nm bandwidth measured off-focus. (c) I0 ∼ 3×1014 Wcm−2 and 40 nm bandwidth measured on-focus. (d)-(f) Projection of the corresponding KER (|cosθ| > 0.9) spectra. imaging technique employed for the purpose of these experiments the time-of-flight (TOF) and impact position of the photofragments from each laser pulse were measured in coincidence. Thus the full 3D momentum components of both fragments can be retrieved. Subsequently, the KER of the dissociation process as well as the angle θ between the dissociation velocity and the laser polarization were determined. These experiments are therefore very sensitive to the precise experimental conditions. The factors which are known to influence the measured kinetic energy release (KER) spectra are a convolution of the vibrational and rotational population of the target molecular ions, the vibrational levels which can be accessed by the bandwidth of the laser, the resolution of the detector, the laser beam focusing and the interaction volume. 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses 108 Figure 4.3: Angular distribution for the dissociation events of H+ 2 for fragments with KER 0.54 eV and 0.82 eV (blue and red data points respectively). The corresponding fitted functions (dashed line) demonstrate a cos2 and cos8 distribution indicating dissociation via a near-resonant transition or bond softening, respectively. Dissociation of H+ 2 Induced by FTL Pulses The angular distributions and kinetic energy release (KER) spectra for the dissociation of H+ 2 using FTL pulses of different bandwidths and intensities are presented in figure 4.2. Each spectrum can be categorized into two different regimes depending on the KER and angular distribution of the fragments. Since perturbative transitions demonstrate a linear intensity dependence and are not obligated to geometric or dynamic alignment, a broad cos2 θ angular distribution signifies their occurrence. This is illustrated by the cos2 θ fit to the angular distribution of the dissociated fragments with KER = 0.82 eV in figure 4.3. The bandwidth of the laser defines the frequency range of the pulse and the vibrational levels that can be accessed by a particular h̄ω transition. The condon point is defined as the internuclear separation where the energy spacing between two electronic states is equivalent to the incident photon energy. For H+ 2 this occurs at Rc =4.8 a.u. corresponding to a vibrational level v=9 for a pulse centered at 795 nm. The cross section for the one photon transition is maximal at this location, and these results are compliant with previous studies [66]. The main KER peak therefore appears around 0.82 eV for an FTL pulse in figure 4.2, as expected. 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses 109 The contributions from the narrow angular distribution from vibrational levels v≤7 with KER <0.61 eV have no overlap with the laser bandwidth and dissociate via bond softening (see section 1.4.3). The cos8 θ fit to the angular distribution for fragments with KER = 0.54 eV verifies that these are multiphoton processes (red dashed line in figure 4.3). The dissociation rate of these lower vibrational levels are strongly enhanced with increasing intensity (also shorter pulses) as the potential barrier is suppressed further. Changing the Bandwidth of the Pulse The spectral bandwidth of the laser can be manipulated by placing an amplitude mask in the Fourier plane of the shaper. This reduces the spectral content of the pulse with respect to the 795 nm central frequency, see section 3.13 for details. This procedure was used to transform 33 fs, ∆λF W HM = 40 nm bandwidth (∆E∼77 meV) FTL pulses of peak intensity I0 ∼ 8×1013 Wcm−2 into 120 fs ∆λF W HM = 18 nm (∆E∼36 meV) ‘narrowband’ pulses of I0 ∼ 5×1012 Wcm−2 . Their effect on H+ 2 dissociation is discussed below. Narrow Spectral Bandwidth (18 nm) The contribution from bond-softening events for ‘narrowband’ ∆λF W HM =18 nm (∆E∼36 meV) pulses is minimal as I0 is substantially lower (see figure 4.2 (a)). The most salient feature however, is the apparent enhancement in energy resolution compared to the pulses with a greater spectral bandwidth. Since the emanating photons across the full spectral bandwidth of an FTL pulse are distributed with equal probability (see PG FROG trace figure 4.1 (a)), the vibrational resolution can be obscured by their energy range. It is therefore expected that vibrational structure becomes indiscernible with increasing bandwidth (shorter pulses) and increasing laser intensity (i.e., field strength) [100]. These differences are compliant with previously published KER spectra which also demonstrate a higher energy resolution involving FTL pulses of longer temporal duration, predominantly due to a smaller spectral bandwidth [72, 101] . 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses 110 Full Spectral Bandwidth (40 nm) The bandwidth ∆λF W HM =40 nm (∆E∼77 meV) of the original FTL pulses is smaller than the vibrational energy spacing of H+ 2 . However, the vibrational structure (see figures 4.2 (e) and (f)) is no longer as clearly resolved due to the greater energy spread in the bandwidth of the laser. Intensity inhomogeneity of the laser pulses as a function of spatial position is inevitable, and the observed spectrum is hence the sum of the physical phenomena occurring at different intensities ranging from zero up to the peak intensity at the focus. This is commonly referred to as intensity averaging and becomes more pronounced for increased intensity and its effect is largely dependent on the zposition (3.16). Figure 4.2 (b) - (f) demonstrate the differences in the KER spectra for the same FTL pulses but probed 19 mm off-focus at I0 ∼ 5×1013 Wcm−2 and I0 ∼ 8×1013 Wcm−2 on-focus (see figure 3.15). In both cases, the v=9 contribution arises from the larger volume, weaker intensity confocal shells of the pulse (see figure 3.25) which have a larger overlap with the ion beam. The contribution of low KER fragments related with strong field phenomena is substantially higher and extends to lower vibrational levels for an increased intensity, as expected. It should however be noted that the lowest KER measurable is limited to 0.1 eV due to the small Faraday cup collecting the on-axisundissociated ion beam. 4.1.3 + + Molecular structure of H+ 2 , HD and D2 + + The isotopes H+ 2 , HD and D2 share the same Born-Oppenheimer potential energy curves and dipole matrix elements. However, the vibrational energy spacing for the heavier molecules are smaller, as the vibrational frequency depends inversely on the reduced mass. The vibrational states are populated according to the Franck-Condon distribution. This is determined by the ground vibrational states of their respective neutral counterparts. Essentially, the vibrational state distribution as a function of the vibrational energy is the same. The KER spectra for these isotopes is therefore similar in shape, but the peak positions change with the respective vibrational spectra. When dressed with a 795 nm laser, the condon point for each isotope will be associated with 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses 111 Figure 4.4: Schematic diagram illustrating the Gaussian fit to the measured power 00 spectrum I(ω) normalized to 1. The applied quadratic phase function ϕ(ω) = 500 fs2 (red dash) results in the frequencies being redistributed in time according to a linear group delay Tg (ω) (blue dashed), which subsequently leads to a frequency sweep within the pulse. Figure adapted from [100]. + + a different vibrational level (H+ 2 ; v=9 τ =29 fs, HD ; v=11 τ =36 fs, D2 ; v=13 τ =41 fs). Any results presented in this chapter comparing two or more isotopes was obtained by alternating the pulse shapes (see section 3.13) and the ion beam (see section 3.3) periodically. This kept experimental conditions such as the ion source temperature (vibrational population), ion beam current, interaction volume, laser focusing, peak intensity, pulse shape and any long term laser drifts consistent. Otherwise, these parameters can fluctuate significantly between measurements. 4.1.4 Linear Chirp A chirped pulse contains a linear increase in frequency with time (frequency sweep). 00 00 This is achieved by applying a quadratic spectral phase function ϕ(ω) = 12 ϕ (ω0 ) · (ω − ω0 )2 to a pulse of central frequency ω0 . The chirp rate is defined by the group dispersion 00 delay (GDD) parameter ϕ . The frequency components of the pulse are temporally redistributed according to a linear group delay. This concept is illustrated in figure 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses 112 Figure 4.5: Simulated Polarization-Gating FROG traces of a 33 fs FTL pulse shaped 00 00 with (a) φ = -1050 fs2 and (b) φ = 1050 fs2 . Figures (c) and (d) show the temporal profile of the 33 fs FTL pulse (black line) normalized to 1 and the relative temporal 00 00 profiles of the chirped pulse (red dashed line) obtained for φ = -1050 fs2 and φ = 1050 fs2 respectively. 4.4 where the blue dot-dashed line indicates that the higher the frequency component, the greater the Tg (ω) acquired, for positive chirp. The PG FROG traces shown in figure 4.5 (a) and (b) can be used to elucidate the frequency-time relation of the redistributed frequencies of the chirped pulses. Heuristically speaking, we can envisage this as the ‘red’ detuned frequencies of the pulse 00 leading the ‘blue’ for a positive chirp (+φ , see figure 4.5 (a)) and the time reversal of 00 this for a negative chirp (-φ , see figure 4.5 (b)). The temporal profile of a chirped pulse is Gaussian, as indicated by the red dashed line in figure 4.5 (c) and (d). The temporal duration τ is elongated compared to the FTL and the peak intensity reduced. The final temporal duration ∆τout for an FTL 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses GDD [fs2 ] 0 780 1040 1300 ∆RM S [fs] 33 78 101 125 113 Relative Intensity 1 0.38 0.28 0.22 Table 4.1: Calculated relative peak intensity and pulse duration for the various GDD pulses used in the experiment. 00 pulse of duration ∆τ , chirped by a magnitude φ can be calculated from the formula below [74]: s ∆τout = 00 2 φ ∆τ 2 + 4ln2 ∆τ (4.2) 00 For the range of GDD used φ(ω) =± 1300 fs2 , the 33 fs FTL pulses were elongated to a maximal duration of 125fs. The fluence and bandwidth of the pulses were kept fixed, but the peak intensity decreased from ∼8×1014 Wcm−2 to ∼3×1012 Wcm−2 for the highest GDD value used, see table 4.1.4. The instantaneous frequency ω(t) of linearly chirped pulses has been used as a tool to control different dissociation dynamics (see review for details [102]). To name a few examples, it has been shown that the directionality of the chirp can be used to manipulate population transfer for transitions which are closely spaced within the chirped pulse bandwidth [103]. The ability to enhance a particular dissociation product by shaping the nuclear wave packet using chirped pulses was demonstrated by Pastirk et al [104]. In addition to this, control of the Landau-Zener (LZ) transitions in NaI predissociation using chirped pulses was studied theoretically. It revealed an enhancement or suppression in the transitions between the excited and the ground state for chirped pulses [105]. The group at WIS recently reported that the KER spectra for the higher vibrational levels of H+ 2 v≥7 close to the Condon point can be manipulated using chirped pulses [106, 67]. The near-resonant transitions for the vibrational levels v≥7 of H+ 2 were shown to dissociate well before the peak of the pulse and are therefore driven by the instantaneous frequency of the pulse E(ω) [67]. This observation manifests itself as shifts in the positions of the KER peak positions to higher or lower 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses 114 KER values for negative and positive chirp respectively. It should be noted however that many of the observed effects depend on the intensity, chirp rate and frequency of the laser. Figure 4.6: The cos θ vs KER density plots of the dissociation of HD+ for FTL pulses shaped with (a) -1300 fs2 and (b) 1300 fs2 creating 125 fs pulses with I0 ∼3×1013 Wcm−2 . The black dotted lines represent the expected field-free vibrational levels of HD+ . 00 The density plots of the HD+ dissociation signal for φ(ω) = ±1300 fs2 pulses as a function of cos θ and KER are shown in figure 4.6, and the differences are striking. A significant increase in the contribution from the low lying vibrational states is observed for positive GDD. The narrow distributions can be expressed as higher cosine powers which suggest geometric or dynamic alignment has taken place. To quantify the dissociation probabilities a yield analysis is presented in figure 4.7. The signal along the axis of the laser polarization in a cone with an angle of 25◦ (|cos θ| > 0.9) from figure 4.6 was integrated over certain KER ranges and normalized to the FTL pulse. The results show the relative photodissociation yield for the one-photon transitions located close to the Condon point with KER ≥ 0.75 eV. This corresponds to vibrational levels + v≥9 for H+ 2 and v≥11 for HD and the yield proves independent of pulse shape. This 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses 115 Figure 4.7: Dissociation probabilities comparing H+ 2 and the two dissociation channels of HD+ for various GDD pulses normalized to the yield of the FTL of peak intensity 5 ×1014 Wcm−2 pulse for (a) One-photon transitions (KER ≥ 0.75 eV) and (b) bondsoftening driven dissociation (KER ≤ 0.41 eV). is expected as it is a one-photon absorption process and does not require any intricate dissociation mechanism. A maximal 20% enhancement in the dissociation rate for low lying vibrational levels (v≥8) of H+ 2 with KER ≤ 0.41 eV was recently reported for positively chirped pulses [107]. These observations were attributed to the manipulation of the avoided crossing using the instantaneous frequency of the pulse E(ω) [107]. The Condon point, which defines the position of the avoided crossing, was found to be dependent on E(ω). The frequency sweep of a chirped pulse can therefore be used to displace the internuclear distance at which the avoided crossing is induced RAC , and the dynamics vary according to the chirp rate. The ‘red’ detuned frequencies will shift RAC to larger values with a higher potential barrier. As a consequence the dissociation rate relative to the FTL pulse is suppressed. For the ‘blue’ frequencies RAC is displaced in the opposite direction and experiences a shift toward smaller RAC . This results in an increased reduction of the potential barrier and thus promotes bond-softening. This concept is 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses 116 Figure 4.8: Modified light induced potential energy curves of H+ 2 (black solid line) illustrating the initiation of an avoided crossing which allows vibrational levels 7 (dotted orange) to dissociate via bond softening. The position (internuclear separation R) of the crossing is defined by the dressed diabatic potential curves (grey). Inset: For positively chirped pulses, the adiabatic potential curves are dynamically modified from red (dashed red) to blue (solid blue) shifted potential curves, according to the direction of the frequency sweep. A larger gap size for blue detuned frequencies develops as the temporal alignment evolves close to peak intensity. Figure adapted from [107]. illustrated pictorially in the inset of figure 4.8. Furthermore, the gap at the avoided crossing is opened wider for higher intensities and if the light-molecule interaction is adiabatic, the molecule can align along the laser polarisation direction. This study was extended to explore any differences in the dissociation rate of H+ 2 and its isotopic variant HD+ using chirped pulses. The main discerning feature between these two molecular ions is the permanent dipole moment of HD+ . This means that + the conservation of parity which applies to H+ 2 can be violated in HD . Two-photon dissociation is therefore possible and has been observed at intensities around 5.0 × 1012 W cm−2 . As the intensity is increased to 1.5 × 1015 W cm−2 four-photon absorption begins to dominate [108]. Furthermore, the nuclear mass correction to the BornOppenheimer approximation means that the 1sσ state lies 3.7 meV below the 2pσ state at the dissociation limit. The ability to use a chirped pulse to control the difference between these two channels would allow realization of chemical control. The H+ 2 and the HD+ ion beams and the pulse shapes were alternated periodically every few minutes (see section 4.1.3) allowing a direct comparison of the isotopes to be accurately made and thus any long term drifts compensated for. 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses 117 + Figure 4.7 shows that the dissociation rate for the H+ 2 and HD fragments with KER ≥ 0.41 eV is enhanced by ∼20%. This is in agreement with that previously reported for H+ 2 [107] and demonstrates the ability to manipulate the dissociation of low lying vibrational levels for HD+ using linearly chirped pulses. However, no difference in the + relative dissociation yield between H+ 2 and HD (or either of its two dissociation chan- nels) was observed for chirped pulses in this range. This supports the idea [108] that permanent dipole transitions play an insignificant role in the strong-field dissociation of HD+ . Recently, HD+ has attracted a lot of attention as a benchmark molecule to explore the role of a permanent electric dipole moment of heteronuclear molecules and their dissociation dynamics. Although no difference in the branching ratio of the H & D+ and D & H+ dissociation channels was observed in this experiment, the group of Prof. Ben-Itzhak at KSU have observed slight differences due to better statistics [109]. Owing to the sensitivity of these experiments, the higher peak intensities, laser bandwidth or even the chirp parameters studied could be contributing factors for this discrepancy. Furthermore, the effects of a weak third pulse and the carrier envelope phase of the dissociating pulse in the asymmetry in the branching of dissociated fragments of HD+ has recently been reported and several control schemes for above threshold dissociation in HD+ . 4.1.5 Third Order Dispersion An increase in the number of molecules dissociating via the bond softening mechanism was observed for positively chirped pulses (see section 4.1.3). Furthermore, laser induced alignment was considered to be an important aspect of these scientific findings. Alignment in a strong laser field can be achieved through various mechanisms depending on the nature of the electric field. If the temporal duration of the pulse is longer than the rotational period of the molecule, a pendular state which liberates around the polarization vector is created and adiabatic alignment occurs. However, once the field is switched off, the molecule will return to its isotropic state. Alternatively, if a molecule encounters a laser pulse whose duration is less than the rotational period 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses 118 Figure 4.9: Schematic diagram illustrating the Gaussian fit to the measured power 000 spectrum I(ω) normalized to 1. The applied cubic spectral phase function φ = 40500 fs3 (red dash) results in the frequencies being redistributed in time according to a quadratic group delay Tg (ω) (blue dashed). Figure adapted from [100]. of the molecule, the momentum imparted acts to ‘kick’ the molecular axis toward the laser field vector. This process is known as non-adiabatic alignment and the system exhibits field-free, post-pulse alignment and rotational revivals [10, 51]. It follows that the combination of these two effects, in the form of a temporally asymmetric pulse where the electric field is turned on slowly and switched off quickly could potentially be used as a tool to enhance molecular alignment. TOD Pulses Recall from section 2.3.1 that the fourth term of the Taylor expansion of the spectral phase function, which is commonly referred to as third order dispersion (TOD), is cubic 000 000 φ(ω) = 16 φ (ω0 ) · (ω − ω0 )3 . This leads to the frequencies being redistributed in time according to a quadratic group delay. This concept is illustrated in figure 4.9 where the 000 φ(ω) is applied with respect to the central frequency, ω0 of an FTL pulse. The blue dot-dashed line demonstrates that whilst the central frequencies experience no group delay, the high and low frequencies of the pulse acquire a greater delay. The PG FROG traces shown in figure 4.10 (a) and (b) can be used to elucidate the frequency-time 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses 119 Figure 4.10: Simulated Polarization-Gating FROG traces of a 33 fs FTL pulse shaped 000 000 with (a) φ = -40500 fs3 and (b) φ = 40500 fs3 . Figures (c) and (d) show the temporal profile of the 33 fs FTL pulse (black line) normalized to 1 and the relative TOD 000 000 temporal profiles (red dashed line) obtained for φ = -40500 fs3 and φ = 40500 fs3 respectively. relation of the redistributed frequencies of the TOD pulses. Heuristically speaking, 000 for negative TOD (-φ ) the ‘red’ and ‘blue’ detuned (high and low) frequencies from the edge of the spectrum are shifted towards the leading edge of the pulse (see figure 000 4.10 (a)). For positive TOD (+φ ) the time order of this process is reversed and these frequencies are shifted towards the trailing edge of the pulse, so that the central frequencies arrive first. This effectively creates a narrower FTL bandwidth on the rising edge of the pulse (see figure 4.10 (b)). In both cases the detuned frequencies arrive simultaneously and the interference between them leads to beating in the time 000 000 domain. This creates a sequence of pre (-φ ) or post (+φ ) pulses depending on the TOD sign (see figure 4.10 (c) and (d)). 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses TOD [fs3 ] 0 13500 27000 40500 ∆RM S [fs] 33 54 97 142 120 Relative Intensity 1 0.71 0.48 0.39 Table 4.2: Calculated relative peak intensity and pulse duration for the various TOD pulses used in the experiment. In the experiment, the fluence and bandwidth of the pulses was kept fixed, but the peak intensity decreased from about 8×1014 W cm−2 for the FTL pulse to 3×1014 W cm−2 for the highest TOD value used. In the latter case, the amount of energy redeployed to the post or pre pulses is increased, and can contain up to half of the overall pulse energy. The highly asymmetric temporal profile of TOD pulses implies that the full width at half maximum (FWHM) is not the most meaningful quantity to characterize the pulse duration. Instead, we use a statistically defined width given in equation 4.3 [74]: s 2σ = ∆τ 2 + 8(ln2)2 2ln2 φ000 ∆τ 2 2 , (4.3) Where ∆τ is the FWHM of the Gaussian FTL pulse before the spectral phase φ(ω) 000 is applied. According to this definition, the 33 fs FTL pulses were elongated to about 50-140 fs for the range of TOD values used, see table 4.1.5. Near-Resonant Transitions The two-dimensional (2D) density plots of H+ 2 dissociation as a function of KER and cos θ for pulses shaped with TOD in the range φ 000 = ± 40500 fs3 are shown in figure 4.11 (a)-(g). The most prominent feature is the significant differences observed in the structure of the v=9 peak. Its evolution can be traced by following the field-free vibrational line (black dashed) from positive to negative TOD values. The initially well resolved peak for large positive TOD values broadens and then splits into two distinct peaks. These changes are more clearly illustrated in figure 4.12, which shows the KER spectrum (for figure 4.11 (a), (d) and(g)) integrated along the laser polarization in a 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses 121 Figure 4.11: The 2D KER-cosθ density plots of H+ 2 dissociation induced by TOD 000 3 3 pulses in the range φ = ± 40500 fs (with 13500 fs increments) as marked on each individual panel. The black dashed lines signify the expected peak position for the field-free vibrational levels at 795 nm. cone with an angle of 25◦ (|cos θ| > 0.9). The KER peak for the FTL pulse appears as expected around 0.82 eV for a central wavelength of 795 nm (see figure 4.12 (b)) but the energy resolution deteriorates because of the 40 nm spectral bandwidth of the pulse (see section 4.1.2). Recall the argument in section 4.1.2 implying that an enhanced resolution can be obtained as a consequence a smaller spectral bandwidth. In figure 4.11 a narrower v=9 peak is observed for increasingly positive TOD despite maintaining a constant bandwidth. This feature and the peak splitting induced by negative TOD can be explained heuristically by considering the temporal behavior of H+ 2 dissociation within the pulse and the 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses 122 limited range of spectral components of the pulse within this time. Time-dependent dissociation probabilities for the v=7–9 vibrational levels of H+ 2 were calculated by the group of Prof. Esry at Kansas State University by solving the timedependent Schrödinger equation (TDSE) numerically. The method used is described in detail elsewhere [66] and assumes that the nuclei have insufficient time to rotate during the pulse and that any rotation after the pulse is negligible. The internuclear axis was kept fixed along the laser polarization allowing nuclear vibration on the coupled 1sσg and 2pσu channels. The calculations were performed for a 30 fs pulse linearly chirped (both in the positive and negative direction) to 120 fs with a peak intensity of 2× 1013 Wcm−2 and the results are shown in figure 4.13. In reality, where rotations play a role, the dissociation probability will saturate at a lower intensity. However, these relatively simple calculations are used to demonstrate that the v=7-9 states are depleted before the peak of the pulse. The exact time window where saturation occurs is not critical for our qualitative model explaining the difference in the structure of the v=9 peak. This early saturation implies that a limited range of frequencies, situated on the rising edge of the pulse are responsible for dissociation. The enhanced energy resolution observed for positive TOD is a consequence of the effective ‘narrow bandwidth’ FTL pulse created on the leading edge. Alternatively, the combination of ‘red’ and ‘blue’ frequency components contained within the rising edge of negative TOD pulses cause the molecules to dissociate at slightly lower or higher KER respectively. Thus the v=9 peak is split into two distinct components. This effect is sensitive to specific pulse parameters such as the bandwidth of the original FTL pulse and the TOD magnitude applied. Since a lower magnitude of TOD will redistribute a smaller range of frequencies around ω0 , a smaller split is observed, as observed in figure 4.11. Non-Resonant Transitions In order to quantify the dissociation probabilities the same yield analysis discussed in section 4.1.3 was performed on the KER and cos θ density plots in figure 4.11. The photodissociation yield for the one-photon transitions located close to the Condon 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses 123 Figure 4.12: Kinetic energy release (KER) spectra (|cos θ| > 0.9) of H+ 2 dissociation 000 000 000 for 33 fs FTL laser pulses shaped with (a) φ = 40500 fs3 (b) φ = 0 fs3 (c) φ = -40500 fs3 . Note the change in the width and the splitting of the v=9 peak for different TOD parameters. 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses 124 Figure 4.13: Calculated time-dependent dissociation probability for a few vibrational levels of H+ 2 , using a 30 fs FTL laser pulse positively and negatively chirped to 120 fs. Negative (dashed) and positive (solid) chirps and plotted for the peak intensity ∼2×1013 Wcm−2 with the intensity envelope depicted by the short dashed line. Adapted from [100]. point and emerging from vibrational levels v≥9 were obtained from integrating those counts along the laser polarization in a cone with an angle of 25◦ (|cos θ| > 0.9) with KER ≥ 0.75 eV. The flat red line in figure 4.14 indicates that these transitions are independent of pulse shape. This again is expected as they proceed via a one photon absorption process and do not require any intricate dissociation mechanism. The highly aligned contribution from the low lying vibrational levels v≤6 of H+ 2 in the KER region ≤ 0.41 eV dissociate via bond-softening and are represented by the black curve in figure 4.14. It is evident that the TOD sign is of paramount importance to the mechanism driving the dissociation. A remarkable 50% increase in dissociation yield was measured for the optimum range of negative TOD ( -27000 to -13500 fs3 ) compared to that of an FTL pulse. This enhancement can be explained in conjunction with the temporal profile of a negative TOD pulse (see figure 4.10). Although the pre-pulses are relatively low in intensity, they can contain up to half of the pulse energy for a peak intensity of 4×1013 Wcm−2 . This is sufficient to induce some degree of alignment 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses 125 Figure 4.14: Dissociation probabilites of H+ 2 for various TOD pulses normalized to the yield of the FTL of peak intensity 5 ×1014 Wcm−2 pulse for (a) One-photon transitions (KER ≥ 0.75 eV) and (b) bond-softening driven dissociation (KER ≤ 0.41 eV). Imperceptible error bars are smaller than the symbols + of H+ 2 and the dissociation is known to be more efficient for H2 aligned with the laser polarization axis as this increases the dipole coupling. Table 4.2 shows how certain pulse parameters vary when TOD of different magnitudes are applied to pulses of fixed fluence. It is clear that increasing the TOD magnitude results in a greater temporal duration. Furthermore, the relative peak intensity decreases from 1 to 0.39 for the highest TOD value used. The overturn in the dissociation yield in figure 4.14 can be explained by considering the two critical factors for inducing dissociation of H+ 2 . This is the compromise between the pulse duration required to align the molecules, and a sufficiently high peak intensity to induce bondsoftening. The overturn in the dissociation yield in figure 4.14 (b) for more negative TOD values therefore suggests the conditions for which molecular alignment becomes the predominant mechanism. Contrary to this augmentation, a dramatic 35% reduction in dissociation yield relative to the FTL pulse is observed for positive TOD. This 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses 126 can be attributed to the absence of pre-pulses and a lesser peak intensity than the FTL pulses (see figure 4.10). A recent study demonstrates theoretically an enhancement in the alignment of N2 for TOD pulses [110], but theses calculations were limited to the rigid rotor regime. The group of Prof. Brett Esry at KSU performed calculations to solve the time-dependent Schrödinger equation (TDSE) for H+ 2 in an intense laser field, replicating the experimental condititions. The method used is described elsewhere [111] and includes radial dynamics (vibrations and in particular dissociation). The time-dependent electric field for these calculations was obtained by adding TOD to the amplitude of the measured power spectrum and performing a Fourier transform. Reproducing the experimental results quantitively requires a non-trivial intensity averaging process and is not within the scope of this work. Instead, the purpose of the calculations is to support the experimental interpretation of the data. Therefore, the theoretical results were obtained by separately propagating all the J=0 bound vibrational states of the 1sσg channel of H+ 2 . It has been shown that the results of such calculations are independent of the initial J state used [111]. The results presented in figure 4.15 (a) show the calculated time-dependent bound state population dynamics of v=6 for H+ 2 in an intense laser field for an FTL pulse shaped with the optimum TOD magnitude φ 000 = 28800 fs3 but opposing signs. The bound state exhibits fast (∼ 1 fs) oscillations which occur with the carrier frequency. The slower oscillations which are locked to the envelope (or sub-pulses of the TOD pulse). A small delay is observed between the slow bound-state population dynamics and the laser field envelope. This delay depends on the pulse parameters, in particular the intensity. The resulting dissociation yield is determined from one minus the bound state population at the end of the pulse. It is clear that the depletion of the vibrational level v=6 is more efficient for the φ 000 = -28800 fs3 pulse, which is congruent with experimental findings. It is interesting to note that qualitatively the difference between the final bound state population for the FTL and φ the FTL and the φ 000 000 = 28800 fs3 is less than that of = -28800 fs3 pulses, similar to that observed in the experiment. To understand how the use of TOD accomplishes photodissociation control, the align- 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses 127 Figure 4.15: Time-dependent calculations for various TOD parameters. The FTL pulse has a peak intensity of 5×1013 Wcm−2 to match the experiment. It is clear the increasing negative TOD leads to a greater pre-pulse alignment and leads to an increase the dissociation for for vibrational level v=6 of H+ 2. ment dynamics have to be considered. The duration of the pre-pulses for negative TOD are < 5 fs and the revival period of H+ 2 is 560 fs. Thus the induced alignment is highly non-adiabatic [112]. Although the separation time between the prepulses is not tailored to ‘kick’ the molecule at the precise revival times [113], each successive pulse acts to gradually increase the degree of alignment [114]. It follows that the degree of alignment achieved for TOD may not be as effective as an accurately timed sequence of pulses. On the other hand, TOD provides a coherent application tool which is experimentally and theoretically accessible. + + H+ 2 and its Isotopic Variants HD and D2 To experimentally explore the concept that prepulses play a principal role in the underlying mechanism responsible for the enhancement observed in the dissociation rate, + + the isotopic variants of H+ 2 (HD and D2 ) were studied simultaneously as a function of TOD (see figure 4.16). For this data set, it is known that the phase of the FTL pulse 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses 128 Figure 4.16: Dissociation probabilities for fragments produced with KER ≤ 0.41 eV + + corresponding to v= 6 of H+ 2 (black), v= 8 of HD (red) and v= 9 of D2 (blue) for various TOD pulses normalized to the yield of the FTL. 000 is not completely flat as it contains some residual TOD of the order of φ ∼ 5000 fs3 . This creates a substantial uncertainty on the x-axis. In the experiment, the ion beam and pulse shapes are alternated periodically (see section 4.1.3) allowing a direct comparison of the hydrogenic ions to be accurately made. The only discernible differences in the dissociation yields occurs in the region of the optimal TOD values as found for 3 H+ 2 (-27000 to -13500 fs ), where alignment was considered to be the dominant control mechanism (see section 4.1.5). The nuclei in heavier molecules move more slower than in lighter molecules. This reduced nuclear motion of heavier molecules allows their evolution to be traced with femtosecond pulses. A qualitative picture has been proposed to suggest that due to the lower velocity of heavier molecules, they effectively experience a shorter pulse duration [115, 116]. This therefore implies that heavier molecules are subject to a lesser degree of laser-induced alignment for the same pulse. This in turn can lead to the molecule experiencing a lesser peak intensity. It is these two contributing factors combined which lead to the reduction in the dissociation yield for the heavier D+ 2. 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses 129 Figure 4.17: Time-dependent laser intensity calculations demonstrating the bound + state population and alignment for the vibrational level v=6 of H+ 2 and v=9 of D2 for 000 000 TOD pulses shaped with φ = -28800 fs3 and φ = -26800 fs3 , respectively. The time-dependent laser intensity calculations (described in section 4.1.5) for the + vibrational levels v=6 of H+ 2 and v=9 of D2 which are located at roughly the same vibrational energy were calculated and compared. Figure 4.17 shows that the alignment and consequently the dissociation is predicted to be significantly lower for D+ 2 , as is experimentally observed. Although the TOD is slightly different this is not believed to be the main cause of the effect. Despite the dissociation efficiency apparently decreasing for more massive molecules, it was theoretically determined that TOD could be used to enhance the alignment of N2 [110]. This supports the concept that TOD can be used as a general application tool to increase the molecular alignment and subsequently molecular dissociation. The effect of changing certain properties of the original FTL pulse which is subsequently shaped with TOD was not explored and could potentially change its efficiency as a tool to enhance bond-softening of heavier molecules. 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses 130 Figure 4.18: Dissociation probabilities of H+ 2 for various TOD pulses normalized to the yield of the FTL pulse of peak intensity 8 ×1014 Wcm−2 for (a) One-photon transitions (KER ≥ 0.75 eV) and (b) bond-softening driven dissociation (KER ≤ 0.41 eV). TOD at Higher Intensity The TOD pulses must maintain a sufficient peak intensity to induce dissociation of the low lying vibrational levels v≤6 of H+ 2 . However, the efficiency of the TOD as a control tool to enhance dissociation decreases relative to FTL pulses for higher intensity. The relative dissociation yield presented in figures 4.14 and 4.18 were measured under the same experimental conditions with the exception of the peak intensity, where 5 ×1014 Wcm−2 and 8 ×1014 W cm−2 was used respectively. As expected, the near-resonant transitions are independent of TOD magnitude or peak intensity, as they are governed by the pulse energy. The same shape of trend line was observed for the bond-softening events for both intensities. However, the quantitative analysis demonstrates that the augmentation in the dissociation yield dramatically decreased from 50% to 10%, for the lower and higher intensities respectively. Recall from figure 4.2 that the dissociation rate of the vibrational levels v≤6 can be increased for FTL pulses of higher intensity. Since the relative dissociation yield is determined with respect to the yield of the FTL, this suggests that intensity is the prevailing factor. Furthermore, this implies that TOD is a more efficient tool for inducing dissociation in the lower intensity regime 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses 131 Figure 4.19: Relative dissociation probabilities of H+ 2 for various TOD pulses where the spectral phase function applied is offset by 10 nm from the central frequency of the laser. Each yield is normalized to that of the FTL pulse for (a) One-photon transitions (KER ≥ 0.75 eV) and (b) bond-softening driven dissociation (KER ≤ 0.41 eV). around 5×1013 W cm−2 . Sensitivity of Wavelength-Pixel Correspondence The experimental technique used to match the central wavelength of the laser to the central pixel of the spatial light modulator (SLM) in the pulse shaper is discussed in section 3.13.1. This procedure is particularly important when applying spectral phase functions which are asymmetric, such as TOD, but not critical for GDD as the linear group delay is monotonically increasing. The results shown in figure 4.14 demonstrate the dissociation yield analysis for TOD pulses where the frequencies are temporally redistributed according to a quadratic group delay around the central frequency, ω0 of the pulse (shown schematically in figure 4.9). However, if the spectral phase is applied around a frequency other than ω0 , an effective additional linear chirp is added. Figure 4.19 shows the dissociation yield analysis for a TOD pulse where the spectral phase is applied 10 nm offset from ω0 . These results contradict the theoretical and experimental findings discussed above disclosing that negative TOD pulses are more efficient for dissociation due to non-adiabatic alignment induced by the pre-pulses. 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses 132 Figure 4.20: Schematic diagram illustrating the Gaussian fit to the measured power spectrum I(ω) normalized to 1 (black line). The applied combination of a quadratic 00 000 φ = 1040 fs2 and cubic φ = 13500 fs3 spectral phase function (red line) results in the frequencies being redistributed in time according to the group delay given by the blue line. Instead, due to the dominance of the higher order of the effective additional linear chirp, the results are similar to those in figure 4.7 for GDD pulses as they suggest that positive TOD is ∼20% more effcient. 4.1.6 Combined Linear Chirp and Third Order Dispersion The ability to enhance the dissociation rate of the low lying vibrational levels v≤6 of H+ 2 using GDD and TOD as independent control tools has been demonstrated in sections 4.1.4 and 4.1.5 above. The underlying control mechanisms were identified as being manifested in the temporal order of the instantaneous frequencies within the pulse and the asymmetric temporal profile respectively. A pertinent question however is how the 00 000 combination of these two terms φ(ω) = 12 φ (ω0 ) · (ω − ω0 )2 + 16 φ (ω0 ) · (ω − ω0 )3 can be used to manipulate the dissociation of H+ 2. Since GDD is the higher-order term it dominates the spectral phase. The addition of TOD enforces an asymmetry to this parabolic spectral phase, as shown by the red line 00 000 in figure 4.20 for φ = 1040 fs2 and φ = 13500 fs3 combined. The corresponding group delay becomes a mildly curved line as represented by the blue line in figure 4.20. 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses 133 Figure 4.21: Simulated Polarization-Gating FROG traces of a 33 fs FTL pulse shaped 00 000 00 000 with (a) φ = -1040 fs2 and φ = -13500 fs3 and (b) φ = 1040 fs2 and φ = 13500 fs3 . Figures (c) and (d) show the temporal profile of the 33 fs FTL pulse (black line) normalized to 1 and the corresponding relative TOD temporal profiles (red dashed line) respectively. For greater TOD values, the asymmetry of the parabola increases and the function as it tends towards a more cubic phase. The corresponding PG FROG traces and temporal profiles in figure 4.24 demonstrate that these pulses contain a frequency sweep and triangular temporal profile. For these particular GDD and TOD parameters there is no beating phenomena on the leading or trailing edge of the pulse. The temporal duration of the GDD and TOD pulses can be defined using equations 4.2 and 4.3 respectively. To estimate the width of the majority of the intensity when these √ two parameters are combined, the intensity variance ∆RM S = 2 < t2 > − < t >2 was calculated. Various pulse parameters for 33 fs FTL pulses shaped with several combinations of GDD and TOD used are given in table 4.3. 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses GVD [fs2 ] 0 780 1040 1040 1040 TOD [fs3 ] 0 13000 27000 13000 40500 ∆RM S [fs] 33 126 100 110 100 134 Relative Intensity 1 0.31 0.33 0.32 0.24 Table 4.3: Calculated relative peak intensity and pulse duration (∆RM S = √ 2 < t2 > − < t >2 ) for the various GVD and TOD combinations used in the experiment. The same yield analysis described above (and presented in figure 4.14) was performed for this data and is shown in the crossword (color map) in figure 4.22. The white squares contain no data and the color bar indicates the strength of the dissociation yield for that particular pulse shape relative to the FTL. The photodissociation yield for the one-photon transitions located close to the Condon point (v≥9) with KER ≥ 0.75 eV are orange, (representing a value of 1), for all pulse shapes (see figure 4.22 (a)). Thus indicating, as described before, that these transitions are independent of pulse shape. However, dissociation via bond-softening from the vibrational levels (v≤6) with KER ≤ 0.41 eV exhibit a multi-colored dependence, see figure 4.22 (b). The two optimal pulse 000 shapes for this process were found to be (a) φ = -13500 fs3 and (b) a combination of 00 000 φ = 780 fs2 and φ = -13500 fs3 (see figure 4.24). Both these pulses are of comparable peak intensity and are asymmetric on the rising edge. This again implies that alignment is the dominant mechanism responsible for the dissociation enhancement. It is known that the sufficiently intense sequence of pre-pulses for TOD in pulse (a) can effectively induce non-adiabatic alignment and consequently enhance the dissociation (see section 4.1.5). Pulse (b) however contains no pre-pulses and a rising edge of comparatively lesser intensity. The enhancement could therefore be due to either adiabatic alignment or the temporal order of the frequencies on the rising edge of the pulse. Since the ‘blue’ frequencies are situated on the leading edge they can displace the avoided crossing to lower RAC which creates a lower potential barrier, and increases dissociation (see section 4.1.3). Both pulse shapes appear to be equally efficient, so determining the dominant mechanism requires further theoretical support. 4.1 Photodissociation of Hydrogenic Ions Using Shaped Laser Pulses 135 Figure 4.22: Crossword for the dissociation probabilities of H+ 2 for pulses shaped with various combinations of GDD and TOD parameters normalized to the yield of the FTL pulse for (a) One-photon transitions (KER ≤ 0.75 eV) and (b) bond-softening driven dissociation (KER ≤ 0.41 eV). The colourmap gives an indication of the dissociation yield relative to the FTL pulse. The white squares contain no data. 4.2 Conclusions and Future Outlook 136 Furthermore, the complicated interplay between these two mechanisms is highlighted in figure 4.24. These two pulses are shaped with the same magnitude but opposite 00 000 00 000 sign with (a) φ = -780 fs2 and φ = -13500 fs3 and (b) φ = 780 fs2 and φ = 13500 fs3 . Where we would expect pulse (a) to be more efficient due to the asymmetry on the rising edge, this is not the case. Since it is the ‘blue’ frequencies which lead the pulse, this suggests that it is the temporal order of the frequencies which dominate for these parameters. It follows from the above discussion that although the mechanisms underlying the enhanced dissociation yield for GVD and TOD independently has been unveiled, it is difficult to predict which will dominate for certain conditions. Figure 4.23: Crossword illustrating the relative dissociation probabilities for bond000 3 softening of the vibrational level v≤6 of H+ 2 for pulses shaped with φ = -13500 fs 00 000 and (b) φ = 780 fs2 and φ = -13500 fs3 . The simulated Polarization-Gating FROG traces of these pulses are shown in (a)(ii) and (b)(ii) and the corresponding temporal profiles are outlined by the red line in (a)(iii) and (b)(iii) respectively. The blue line in figures (a)(iii) and (b)(iii) show the original 33 fs FTL pulse normalized to 1. 4.2 Conclusions and Future Outlook In summary, a three-dimensional (3D) momentum imaging technique was used to measure the KER and angular distributions for the dissociation fragments of hydrogenic 4.2 Conclusions and Future Outlook 137 Figure 4.24: Crossword illustrating the relative dissociation probabilities for bond00 softening of the vibrational level v≤6 of H+ 2 for pulses000shaped with (a)(i) φ = -780 000 00 fs2 and φ = -13500 fs3 and (b)(i) φ = 780 fs2 and φ = 13500 fs3 . The simulated Polarization-Gating FROG traces of these pulses are shown in (a)(ii) and (b)(ii) and the corresponding temporal profiles are outlined by the red line in (a)(iii) and (b)(iii) respectively. The blue line in figures (a)(iii) and (b)(iii) show the original 33 fs FTL pulse normalized to 1. molecules using analytically shaped pulses. The pulses were shaped using the quadratic 00 000 (ϕ ) and cubic (ϕ ) spectral phase terms of the Taylor expansion given in equation 2.12 and are referred to as GDD and TOD respectively. The magnitude and sign of these terms was then altered in a systematic manner and any differences in the spectra observed and a yield analysis relative to the FTL performed. The scientific findings reported in this thesis conclude that the KER structure of the resonant transitions can be manipulated using the temporal order of the frequencies of the pulse. This manifests itself as either the resonant peak (v=9 for H+ 2 ) splitting into two distinct features or an apparent increase in the energy resolution for negative and positive TOD pulses respectively. The ability to increase the dissociation yield of the low lying vibrational levels of H+ 2 using pulses shaped with GDD and TOD was experimentally demonstrated. It was proposed that the 20% enhancement in bondsoftening events observed for positive GDD was a consequence of the position and 4.2 Conclusions and Future Outlook 138 size of the avoided crossing being dynamically shifted in accordance with the direction and rate of the frequency sweep. Since the blue detuned frequencies are located on the rising edge of the pulse for positive GDD, the avoided crossing will be induced at smaller internuclear separtations and hence the potential barrier will be lower, leading to an increase in dissociation. Similarly, a remarkable 50% increase in the bond-softening events observed for negative TOD pulses is attributed to the asymmetry of the temporal profile. The series of pre-pulses created are intense enough to induce non-adiabatic alignment, which subsequently leads to an increase in the dissociation rate. It is also interesting to note that the relative dissociation enhancement observed for negative TOD is stronger for lower intensities. Although both the instantaneous frequency and the temporal profile of the pulse can be used as a tool to control the dissociation of hydrogenic molecules, it is unclear which mechanism dominates when both GDD and TOD are combined. Further experiments using a single stronger pre-pulse, instead of the series of pulses or an asymmetric temporal profile which contains no frequency sweep could be used to elucidate the conditions under which control mechanism prevails. There was no experimental evidence found in this data to support the concept that the frequency sweep of a GDD pulse could be used as a tool to control the branching ratio of HD+ . The difference in energy between the dissociation limits of HD+ are very small (3.7 meV) and such a control scheme could potentially be demonstrated more clearly using a molecule with a larger energy separation between the two dissociation channels. TOD lacks the flexibility of an accurately time sequence of pulses tailored to kick molecules at specific revival times and is therefore not the best pulse shaping tool for optimizing molecular alignment. However, it does provide a robust, coherent control tool which should not be highly isotope dependent and moreover is both experimentally and theoretically accessible. The underlying mechanisms induced by GDD and TOD have been identified and can act as a platform to understanding the interaction of these shaped pulses with more complex molecules. Perhaps when interacted with polyatomic molecules alignment along a preferred direction of orientation will occur due to a specific property of the molecule and subsequently enhance the dissociation of 4.2 Conclusions and Future Outlook 139 a specific product. In a recent study, the most efficient temporal profile for silicon wafer micro machining was found to be asymmetric [117]. It follows that the combination of GDD and TOD could be explored in a similar manner. Furthermore, such pulse shapes could be used to explore laser ablation of metals and areas of semi-conductor science. Consequently, subject to further extensive investigation, such pulse shapes could potentially find a use in industrial and healthcare applications. Chapter 5 Laser Induced Fragmentation of CD+ The electron density of heteronuclear molecules is shared unequally between the two nuclei. Therefore the inversion of charge symmetry that exists between the electronic states of homonuclear molecules is violated and electronic transitions which would nominally be forbidden are allowed. Heteronuclear molecules contain a complex plethora of electronic states and an abundance of possible fragmentation pathways. Owing to the different electron affinity and ionization energy of the two constituent atoms, the energy required to reach these pathways can differ greatly. It is therefore interesting to explore how the branching ratio of these channels change as a function of laser pulse parameters. However, measuring the molecular fragmentation of mass asymmetric molecules using the 3D momentum imaging technique is experimentally challenging. For this reason, a longitudinal and trasverse field imaging (LATFI) technique has been developed by the group of Prof. Ben-Itzhak at the J.R.M laboratory, at Kansas State University, USA (KSU). As part of a collaboration it was used to study the laser-induced fragmentation of the highly asymmetric CD+ molecular ion. The principles of the LATFI method and the various fragmentation channels and branching ratios measured for CD+ are outlined in this chapter. 5.1 Laser-Induced Fragmentation of CD+ 5.1 141 Laser-Induced Fragmentation of CD+ The development of short pulse lasers and the study of ultrafast dynamical processes has recieved tremendous attention in recent years, leading to a new realm of physics coined ‘femtochemistry’. Following technological advancements, unprecidented progress in this emergent field has been achieved with the ‘holy grail’ being able to control and direct chemical reactions of complex molecules. However, interpreting the laser-induced dynamics of molecular ions with a more complicated internal structure than that provided by the two dominant potential energy curves (PEC) of H+ 2 can prove problematic. The various theoretical and experimental challenges associated with studing more complex molecules are daunting, but have inspired scientists to explore this regime more extensively. In this work CD+ , a multielectron heteronuclear diatomic, has been chosen as an exemplar molecule for extending the study of homonuclear simple molecules. The multitude of potential energy curves (PEC) of differing spin degeneracy, symmetry and quantum number, as denoted by the term symbols (see section 1.3.3), of CD+ are shown in figure 5.1. Interpreting the laser-induced dissociation pathways of such a molecule is an important step towards understanding more complex systems. In the prescence of the laser field, the spin-orbit coupling between the singlet and triplet states is negligible. Hence from the optical transition rules (see section 1.1) these states can be treated independently as the required spin flip of an electron is highly improbable. Provided reliable PES of a molecule can be obtained and ideally an estimation of the initial vibrational population, the Floquet picture (see section 1.4.2) can be used to gain an understanding of possible dissociation pathways. Previous studies have reported the use of angular distributions to limit possible dissociation pathways of O+ 2 [118], the and kinetic energy release (KER) to find the initial and final states in double ionisation [119]. Furthermore, in a more comprehensive study all of the information was combined using a 3D momentum imaging technique and used to determine the laser-induced dissociation pathways of O+ 2 [120]. From an intense field perspective, CD+ is a particularly interesting target to probe 5.1 Laser-Induced Fragmentation of CD+ 142 Figure 5.1: The potential energy curves of CD+ where the singlet curves are represented by the red lines and the triplet represented by the black. Where the Π states are dashed and the Σ states are solid. as the two lowest dissociation limits C(3 P) + D+ and C+ + D(1 S) are separated by an energy of 2.4 eV which is accessible by either one 785 nm or two 381 nm photons. Investigating the intensity and wavelength (381 nm and 785 nm) dependence of the possible dissociation pathways may elucidate dynamics associated with each spin state. Furthermore, the branching ratio of these two dissociation channels is particularly interesting and timely given the long out-standing goal of achieving and controlling chemical reactions. Previous studies show a strong intensity dependence in the branching ratio of for dissociation of ND+ and DCL+ [121, 122]. However the energy separation of the two studied dissociation limits was less than one photon in both cases. Thus CD+ presents unique characteristics for further study. Although tremendous progress has been made in understanding laser-induced dynamics, measuring the dissociation products unambiguously has in the past proven difficult. The advantage of using a coincidence 3D-imaging technique (as described in sectio [REF]) is that the initial velocity of the target allows both neutral and charged fragments to be detected. Furthermore, the longitudinal electric field accelerates the 5.2 Longitudinal and Transverse Field Imaging Technique 143 charged fragment with respect to its neutral counterpart creating a temporal separation. Thus when measured in coincidence the fragmentation channels can be clearly separated. However, the large mass ratio (12:2) of CD+ makes measuring both of the dissociation channels simultaneously using this method impractical. The slow fragment can get blocked by the Faraday cup positioned to collect the primary ion beam. Alternatively, the fast fragment may travel outside the detector face. In these first measurements of intense field dissociation of CD+ , a piecewise measurement procedure to determine the dissociation branching ratio was developed. 5.2 Longitudinal and Transverse Field Imaging Technique As part of a collaboration with the J. R. M Laboratory at Kansas State University (KSU), USA, the fragmentation of the CD+ molecular ion was studied in the lab of Prof. Itzik Ben-Itzhak. The 3D momentum imaging setup used at KSU is very similar to that at WIS and described in detail in chapter 3 and shown schematically in figure 3.1. It is operated as a longitudinal field imaging (LFI) technique in the format described. In these experiments the dissociation velocity of the fragments are measured and the kinetic energy release (KER) and angle θ between the laser polarization and the molecular dissociation axis calculated. The molecular breakup is symmetrical about the azimuthal angle φ and this can be used to reconstruct losses in measured data if required. The limitations imposed by operating the technique in this mode is the loss of low energy fragments caused by the Faraday cup which collects the primary ion beam. To overcome this restriction the group at KSU developed the longitudinal and transverse field imaging (LATFI) technique. This method requires an additional deflector strategically positioned between the spectrometer and the detector (see figure 5.2 (a)). The fragments are then deflected by the transverse static E-field separating low energy KER fragments from the primary ion beam. The fragments are then spatially separated on the detector according to their energy to charge ratio. This technique enables fragments with zero KER to be measured. Figure 5.2 (b) shows a schematic diagram of the relative spatial separations for the 5.2 Longitudinal and Transverse Field Imaging Technique 144 centre of mass of the CD+ fragments on the detector for various theoretical deflector voltage conditions. For the traditional longitudinal field imaging (LFI) technique where zero voltage is applied to the deflector (see Figure 5.2 (b)(i)) there are substantial losses due to the Faraday cup. For the low voltage conditions (see Figure 5.2 (b)(ii)) the primary ion beam is deflected and collected by an off-axis Faraday cup. The neutrals strike the detector where the primary ion beam would have been in the absence of the deflector. The centre-of-mass (COM) of only two CD+ fragments (D+ and C2+ ) can be cleanly measured under these conditions as the COM of the C+ is lost to the Faraday cup. For the high voltage settings (see Figure 5.2 (b)(iii)) the COM for the C+ is sufficiently deflected away from the Faraday cup. However, as a consequence the COM of the D+ is deflected completely off the detector and theoretically only about half of the C2+ distribution can be measured. A quantitative illustration of where the COM of each fragment lies relative to each other with respect to the deflected CD+ ion beam is shown in figure 5.3. If the primary ion beam is deflected by 4-5 mm from the centre of the detector then correspondingly the COM of the D+ fragments will be located at a distance 28-35 mm from the centre (green line). A separation distance of 2-2.5 mm between the C+ and the Faraday cup centre (light blue line) is necessary to ensure the C+ COM can be measured. The primary beam should therfore be deflected by 12-15 mm. The voltages required to achieve these relative distances can then be determined theoretically. Although the size of the detector is effectively reduced by the deflection of the CD+ beam, some additional detection space can be obtained by initially dog-legging the CD+ beam. To determine the dissociation branching ratios a piecewise method was used. The deflector voltage settings were optimized to measure each dissociation channel independently. Ideally, under all voltage conditions a common fragmentation channel would be measured and used to normalize the data sets to each other. In principal, the charge asymmetric dissociation (CAD) channel C2+ +D would have been an ideal candidate for this as it suffers from minimal geometric losses in both cases. However, in practice this CAD channel was not observed for the 391 nm pulses and appears only in the high intensity 795 nm measurements. 5.2 Longitudinal and Transverse Field Imaging Technique 145 Figure 5.2: (a) A schematic diagram of the longitudinal and transverse field imaging (LATFI) setup developed at Kansas State University (KSU), USA. The neutrals are unaffected by the presence of the static transverse electric field and the charged species are deflected according to their Energy/Mass ratio. (b) The relative positions of the CD+ fragments on the detector are illustrated for (i) 0 V (ii) low voltage and (iii) high voltage settings applied to the transverse deflector respectively. 5.3 Results and Discussion 146 Figure 5.3: The relative positions of the fragments on the detector with respect to the deflected primary CD+ ion beam. To further minimize geometric losses the polarization of the laser can be used to increase the angular acceptance of the detector. Also the spectrometer voltage and ion beam energy can be altered. The physical size of the momentum distribution of the fragments on the detector can be reduced be decreasing the travel time to the detector. Although this may retain a greater contribution of the lighter fragments to the confines of the detector, the distribution of heavier fragments are more likely to be lost in the Faraday cup. Another reason for using a higher ion beam energy is to ensure that all fragments are efficiently detected. For a CD+ ion beam energy of 21 keV and spectrometer voltage 1400 V the D neutral fragment was detected with an energy 2.8 keV. It was experimentally verified that the measurement was obtained on the platau of the efficiency curve. For the same experimental conditions the brancing ratio of the two dissociation channels was found to be independent of MCP voltage. 5.3 Results and Discussion 147 Figure 5.4: Coincidence time of flight spectra showing the laser-induced fragmenation channels measured for the interaction of 30 fs pulses at 785 nm and I0 = 1 ×1016 W cm−2 with CD+ . 5.3 Results and Discussion The 30 fs and 55 fs pulses at a central wavelength 785 nm and 391 nm respectively, were generated by the PULSAR laser (10 kHz) at KSU. The laser beam was then focused to achieve an intensity range 1.3 × 1013 − 1.6 × 1016 W cm−2 and 2.2 × 1011 − 2.5 × 1015 W cm−2 respectively. The coincident time of flight spectra for the laser induced fragmentation of CD+ is shown in figure 5.4. There were five fragmentation channels identified, two dissociation channnels, two charge symmetric ionization channels and a charge asymmetric dissociation (CAD) channel. 5.3.1 Dissociation Channels The potential energy curves for the field free electronic states for CD+ are given in figure 5.1. In the experiement CD+ is produced from methane by electron impact in an electron cyclotron resonance (ECR) ion source. The length of the C-D bond in methane (1.085 a.u.) is the same as that of the ground state of the CD+ molecular ion (1.09 a.u.). A Franck-Condon projection of the the ground state wave function 5.3 Results and Discussion 148 onto the CD+ states can then be used to estimate the vibrational population. The population is likely to be relatively low-lying in a mixture of both the singlet X1 Σ+ and triplet a3 Π states, as there are no optical transitions between the two. However, the exact ratio of the population between the two states is unknown. To determine the most probable dissociation pathways the dressed states Floquet approach (see section 1.4.2) is employed to visualise the dissociation routes (see figure 5.5). The excitations between states are indicated as vertical transitions, resonant with an integer number of photons (nh̄ω). Thus all the electronic curves are shifted downwards (or upwards) in energy by the number of absorbed (or emitted) photons. For example X1 Σ+ − 2ω indicates that the X1 Σ+ state has been shifted down in energy by two photons. Each state then repeats itself periodically with an energy separation equal to multiples of photon energy. The singlet and triplet CD+ potential energy curves (PEC) dressed with 785 nm photons are shown in figure 5.5 (a) and (b) respectively. When considering the second harmonic (391 nm photons) one may still refer to 5.5, but should note that only transitions involving an even number of photons at 785 nm are applicable. From the abundance of potential energy curves for CD+ determining the possible dissociation pathways seems an insurmountable task. However, there are four recommended guidelines to follow in the process of elimination [120]. The transition probability stipulates that the most plausible pathway requires the fewest number of photons. This means that contributions from highlying excited states are unlikely to make major contributions. The angular distribution of the fragments are a signature of the type and number of contributions from transitions favoring internuclear alignment parallel (∆Λ=0, leading to a cosn θ distribution) and perpendicular (∆Λ=±1, giving sinn θ distribution) to the laser polarization (see section 1.4.7). Furthermore, the position and shape of the KER spectra can be used to gain an insight into the general shape of the potential barrier over which the vibrational wavepacket dissociates and hence the PECs comprising it. Each curve crossing should be considered in the adiabatic picture as an avoided crossing. The electric field must be turned on for long enough to allow sufficient time for the wavepacket to traverse the series of potential energy 5.3 Results and Discussion 149 Figure 5.5: The potential energy curves of CD+ dressed with 795 nm photons for the (a) singlet and (b) triplet states. 5.3 Results and Discussion 150 Figure 5.6: Angular distributions (where only half of the distribution is measured and the other half reflected) and KER spectra for the C+ + D dissociation channel of CD+ for pulses of (a) 30 fs at 795 nm and I0 = 3.2 ×1015 W cm−2 and (b) 50 fs at 391 nm and I0 = 2.5 ×1015 W cm−2 . curves and escape. In addition to all of the above, the molecular dipole selection rules must be obeyed which elimnates numerous ineffectual crossings. Finally, it should be noted that an inherent problem with the volume effect (see section 2.1.1) is that any intensity dependent transitions may be obscured and sensitivity to any fine structures can become lost. C+ + D Pathway The lowest dissociation limit of CD+ is the C+ + D(1 S) channel which can dissociate only dissociate via two different pathways, from either a singlet X 1 Σ+ → A1 Π or a triplet a3 Π+ → c3 Σ+ state (see figure 5.1). The angular distributions and KER spectra for this channel at 795 and 391 nm are shown in figure 5.6 (a) and (b) respectively. The KER spectra for the 391 nm pulses demonstrates a sharp peak at ∼0.1 eV. The 5.3 Results and Discussion 151 broad angular base in the distribution shown in 5.6 (b) indicates a mixture of both parallel and perpendicular transitions. Only one possible transition from a singlet state can result in KER distribution peaked close to zero: X 1 Σ+ → A1 Π − 2h̄ω (5.1) This transition requires only one (381 nm) photon and relies on the low lying vibrational levels of the ground state being populated making it a highly feasible option. However, equation 5.1 suggests that perpendicular transitions dominate and this would result in a distribution concentrated at cosθ=0. This is conflicting with the strongly peaked contribution observed along the laser polarization, as shown in figure 5.6 (b). Nevertheless, for the intensity range studied dynamic alignment could be the reason for the observed angular distribution. If the population undergoes a resonant Raman transition to the bound part of the A1 Π state at the inner turning point where the transition moment is highest, then an aligned and highly rotationally excited molecule is more likely to dissociate. For the 785 nm pulses, the KER spectra is peaked around 0.4 eV and the angular distribution is highly aligned along the laser polarization with a sharp cos2 θ distribution. Thus suggesting that parallel transitions dominate. The same singlet state pathway and mechanism described in equation 5.1 but instead using two 785 nm photons are also relevant. From figure 5.6 it is clear that the contribution for the 785 nm is more highly aligned than for the 391 nm. This could reflect the higher intensities used for the 795 nm as this would lead to an increase in alignment. From the dressed triplet state potential energy curves in figure 5.5 (b) it is clear that the molecule can dissociate through the following pathway: a3 Π → b3 Σ− − 2h̄ω → d3 Π − 3h̄ω → a3 Π − h̄ω (5.2) However, this combination of parallel and perpendicular transitions requires six 785 nm photons. It is therfore expected to demonstrate a strong intensity dependence and considered unlikely. 5.3 Results and Discussion 152 Figure 5.7: Angular distributions and KER spectra for the D+ + C dissociation channel of CD+ for pulses of (a) 30 fs 795 nm and I0 = 3.2 ×1015 W cm−2 and (b) 50 fs at 391 nm and I0 = 2.5 ×1015 W cm−2 . D+ + C Pathway The alternative dissociation channels D+ + C are situated at an energy of 2.4 eV, 3.65 eV and 5 eV above the C+ + D(1 S) dissociation limit. The KER spectra and angular distributions shown in figure 5.7 peak around 0.5 eV and do not span as low as zero KER, contrary to the observation for the C+ + D channel. The angular distributions are peaked along the direction of the polarization for both wavelengths and their similarity in structure suggest similar dissociation pathways. There are no feasible pathways for the 381 nm photons starting from the triplet states as the corresponding dissociation limit in the dressed state picture is shifted to an energy greater than 1 eV from the bottom of the a3 Π curve (see figure 5.5). From the KER spectra and consideration of the singlet dressed states, an energetically possible 5.3 Results and Discussion 153 pathway for both wavelengths is given by: X 1 Σ+ → A1 Π − 2h̄ω → 31 Σ+ − 6h̄ω (5.3) However, this involves two perpendicular transitions and is therefore expected to peak at cosθ=1 and not along the polarisation as was experimentally observed. For the 785 nm pulses an additional pathway consisting of parallel and perpendicular transition is also possible: X 1 Σ+ → A1 Π − 2h̄ω → 21 Π+ − 5h̄ω 5.3.2 (5.4) Branching Ratio Each dissociation channel was measured independently using the optimal longitudinal and transverse field imaging (LFTI) conditions to maximize the solid angle collected for each fragment (see section 5.2). The number of counts from each channel [C+ + D] and [D+ + C] was then normalize to ensure the number of molecules exposed to the laser was constant. A correction factor to account for the fraction of the solid angle efficiently collected was calculated using a moving average of the counts in the φ plane. The overall counts in an ideal flat distribution was then estimated and the percentage of the φ distribution experimentally measured was determined. The branching ratio is defined as the ratio of the number of dissociation counts from one channel to the total number of dissociation counts: Branching Ratio = [C + [C + + D] + D] + [D+ + C] (5.5) This can be used to examine the intensity dependence of the two dissociation channels. Figure 5.8 shows the branching ratio as a function of intensity for the 391 nm and the 785 nm pulses. The error bars calculated reflect the statistical errors. For the 391 nm pulses a linear increase in the C+ + D(1 S) channel with intensity was observed. This is in agreement with the one photon pathway X 1 Σ+ → A1 Π − 2h̄ω suggested as such processes are expected to scale linearly with intensity. At higher intensities 5.3 Results and Discussion 154 Figure 5.8: The branching ratio for the two dissociation channels of CD+ as a function of intensity for 391 nm (red dots) and 785 nm (black dots). The error bars calculated reflect only statistical errors. the gradual flattening of the line may be an indication of the dissocaiation dynamics being complicated by the opening of new dissocaition pathways or even the onset of ionisation. The branching ratio for the 785 nm pulses shows no significant intensity dependence over the range of intensities studied. 5.3.3 Ionization Channels The charge symmetric ionization channels are the hardest to measure. There are no optimal combination of voltage conditions that can accomodate the efficient collection of the two charged constituent fragments of CD+ . Hence substantial losses are inevitable for this data. The angular distributions for the single and double ionization channels for an intensity range 1 × 1014 Wcm−2 − 1 × 1016 Wcm−2 at 785 nm are shown in figures 5.10 and 5.11 respectively. 5.3 Results and Discussion 155 Figure 5.9: The Potential Energy Curves of CD+ including the ionisation limits. Single Ionization The single ionization channel of CD+ results in the two fragment ions C+ and D+ . The angular distribution of the kinetic energy released in the single ionization channel for 785 nm pulses at a range of intensities are shown in figure 5.10. The cut along the direction of the polarisation axis in angular distribution is an experimental artifact caused by the loss of the slow C+ to the Farady cup. The observed structure is a broad angular base with a highly aligned central feature. This is a signature of stepwise ionisation and has been reported for other diatomics [45]. Furthermore, the contribution from each feature demonstrates an intensity dependence. If ionisation proceeds in a stepwise manner, the molecule is first excited to a dissociative state and then subsequently ionized as the molecule stretches. Thus some characteristics of the dissociation channels are retained and reflected in the angular distributions as is the case here. 5.3 Results and Discussion 156 Figure 5.10: The angular distribution for the single ionisation channel C+ + D(1 S) of CD+ at 785 nm for an intensity range 1 × 10 14 Wcm−2 to 1 × 10 16 Wcm−2 . 5.3 Results and Discussion 157 Figure 5.11: The angular distribution for the double ionisation channel C2+ + D+ of CD+ at 785 nm at an intensity of (a) 2.5 × 10 14 Wcm−2 and (b) 1 × 10 16 Wcm−2 Double Ionization The multiple ionization of CD+ can lead to the fragment ions C2+ and D+ . The angular distribution of the kinetic energy released in the double ionization channel for 795 nm pulses are shown in figure 5.11. A strongly aligned contribution along the laser polarization can be observed, but it should be noted that the ability to measure low cosθ values in this KER range is limited, and the angular distribution may be broader. The peak of the KER is seen to shift from 10-14 eV with increasing intensity. The contribution from electron rescattering (see section 1.5.1) was experimentally tested by using circularly polarized light. This way the initially ionized electron cannot return to the parent ion and release another electron through ineleastic scattering. No significant reduction in the multielectron ionisation was observed for circulary polarised light. Thus ruling any significant contribution from this ionisation mechanism out. The proposed mechanism for this double ionisation channel is the stair-step ionisation process described in section 1.5.3. In this sequential ionization procedure, the molecule begins to traverse either of the two dissociation channels before being ionized. Following this, the molecule then stretches along the single ionisation channel before being ionized again. Since these series of events occur consecutively at later times in the pulse, the laser field has more time and intensity to populate a broader higher rotational state. This means that the angular distributions emerginging from higher ionization stages are likely to be more aligned. Furthermore, the increasing 5.3 Results and Discussion 158 Figure 5.12: The angular distribution and KER spectra for the charge asymmetric dissocaition channel C2+ + D of CD+ at 785 nm at an intensity of (a) 1 × 10 16 Wcm−2 . number of events acts to obscure any KER structure from the inital stages. 5.3.4 Charge Asymmetric Dissociation The loss of an electron from CD+ can also result in charge asymmetric dissociation (CAD) as described in section 1.5.2. This refers to the charge being shared unevenly between the fragments. For CD+ this can result in population dissociating into a C2+ +D state. The charge symmetric dissociation (CSD) channels are usually favored as their limits are generally energetically lower. The calculations of such highly charged CAD potential energy curves are extreemely demanding and have been attempted on several occassions with variations in the assignment of states. The PEC’s presented in figure 5.9 show the 10 eV energy separation between the single ionisation C+ (2 D)+D+ and the CAD channels. Thus contributions from the CAD channel are substantially weaker by comparison. Furthermore, the authors suggest the opening of an avoided crossing between the CSD (D+ + C+ (2 D)) and CAD channels at 10.5Å, with a 1.5 eV energy separation. The wavepacket can only feasibly reach this avoided crossing via excitation to either 5.4 Conclusions and Future Outlook 159 the 22 Σ+ or the 22 Π potential curves. From the angular distribution and KER spectra for the C2+ +D dissociation channel of CD+ (see figure 5.12) it can be concluded that the molecule must be stretched to approximately 4 - 8 a.u. in order to emerge from either of these two channels with a KER within the range 0-3 eV. According to the Landau-Zener formula, the transition probability of the wavepacket at the avoided crossing is governed by its velocity and it is this property that will determine the final fragmentation pathway of the molecule. Interestingly, the CAD channel was only observed at 785 nm (30 fs) and not 391 nm (50 fs) for equivalent intensities. Intuatively, one would expect the latter to be more probable with the higher energy photons, given the location of the dissociation limit. A possible explaination is that the temporal duration of the pulse plays an integral role in the fragmentation dynamics [123]. 5.4 Conclusions and Future Outlook the use oof short pulses.... would allow how far the molecule strtches during the stairstep ionisation process to be monitored. Research Publications Refereed Publications A. Natan, J. A. Levitt, L. Graham, O. Katz and Y. Silberberg Standoff detection via single-beam spectral notch filtered pulses. Appl. Phys. Lett. 100, 051111 (2012) J. B. Greenwood, O. Kelly, C. R. Calvert, M. J. Duffy, R. B. King, L. Belshaw, L. Graham, J. D. Alexander, I. D. Williams, W. A. Bryan, I. C. E. Turcu, C. M. Cacho and E. M Springate A comb-sampling method for enhanced mass analysis in linear electrostatic ion traps. Rev. Sci. Instrum. 82, 043103 (2011) J. D. Alexander, L. Graham, C. R. Calvert, O. Kelly, R. B. King, I. D. Williams and J. B. Greenwood Determination of absolute ion yields from a MALDI source through calibration of an image-charge detector. Meas. Sci. Technol. 21, 045802 (2010) J. D. Alexander, C. R Calvert, R. B. King, O. Kelly, L. Graham, W. A. Bryan, G. R. A. J. Nemeth, W. R. Newell, C. A. Froud, I. C. E Turcu, E. Springate, I. D. Williams and J. B. Greenwood Photodissociation of D+ 3 in an intense, femtosecond laser field. J. Phys. B. 42, 141004 (2009) 5.4 Conclusions and Future Outlook 161 P. Bruggeman, L. Graham, J. Degroote, J. Vierendeels and C. Leys Water surface deformation in strong electrical fields and its influence on electrical breakdown in a metal pin-water electrode system J. Phys. D. 40, 4779 (2007) L. Graham, P. J. Van der Burgt, J. Alexander, T. L. Merrigan, C. A. Hunniford, C. J. Latimer, I. D. Williams and R. W McCullough Fragmentation of Acetonitrile in collisions with H− and O− negative ions In Preparation M. Zohrabi, L. Graham, U. Lev, U. Ablikim, B. D. Bruner, J. J Hua, J. McKenna, K. J. Betsch, B. Jochim, A. M. Summers, B. Berry, D. Strasser, O. Heber, Y. Silberberg, D. Zajfman, K. D. Carnes, B. D. Esry and I. Ben-Itzhak Dissociation of H+ 2 by broad bandwidth laser pulses: selective molecular response causing effective bandwidth narrowing In Preparation U. Lev, L. Graham, B. D. Bruner, J. J Hua, V. S. Prabhudesai, A. Natan, C. B. Madsen, B. D. Esry, I. Ben-Itzhak, D. Schwalm, I. D. Williams, O. Heber, Y. Silberberg and D. Zajfman Quantum control of H+ 2 photodissociation using femtosecond pulses shaped with third order dispersion In Preparation Research Presentations September 2012 Quantum Control of Photodissociation Using Shaped Ultrafast Pulses Quantum Atomic, Molecular, and Plasma Physics, (QuAMP), Belfast, UK July 2011 Fragmentation of Acetonitrile in collisions with H− and O− negative ions XXVII International Conference on Photonic, Electronic and Atomic Collisions (ICPEAC), Belfast, UK January 2010 The Ionization and Fragmentation of Acetonitrile by Low-Energy NegativeIon Impact Atomic and Molecular Interactions Group (AMIG), Milton Keynes, England May 2009 Absolute Calibration of an image-charge detector 4th Annual ITS-LEIF Meeting 2008, Girona, Spain May 2009 Ultrafast Laser Driven Recombination in Excited ions Physics at EBITS and Advanced Research Lightsources (PEARL), Dublin, Ireland Dec 2008 Ultrafast Laser Driven Recombination in Excited ions High Power Laser Science Meeting, Abingdon, England (3rd poster prize) 5.4 Conclusions and Future Outlook 163 References [1] J. M. Hollas, High Resolution Spectroscopy. Wiley, 1982. [2] R. Eisberg and R. Resnick, Quantum Physics of atoms, molecules, solids, nuclei, and particles; Second Edition. John Wiley Sons,, 1985. [3] H. B. Dunford, Elements of Diatomic Molecular Spectra. [4] A. M. Clark, “Molecular orbitals inmoe,” November 2008. [5] M. Gerloch, Orbitals, Terms and States. John Wiley Sons,, 1986. [6] P. W. Atkins, Molecular Quantum Mechanics; Second Edition. Oxford University Press, 1983. [7] P. M. Morse, “Diatomic molecules according to the wave mechanics. ii. vibrational levels,” Phys. Rev., vol. 34, p. 57, 1929. [8] J. H. D. Eland, Photoelectron Spectroscopy, Second Edition. Butterworth Co, 1984. [9] A. Giusti-Suzor, F. H. Mies, L. F. DiMauro, E. Charron, and B. Yang, “Dynamics of H2 + in intense laser fields,” J. Phys. B., vol. 28, p. 309, 1995. [10] J. H. Posthumus, “The dynamics of small molecules in intense laser fields,” Rep. Prog. Phys., vol. 67, p. 623. [11] A. D. Bandrauk, Molecules in Laser Fields. Marcel Dekker, Inc., New York,, 1994. [12] D. R. Bates., “The oscillator strength of H2 + , 1sσ2pσ,” J. Chem. Phys., vol. 19, p. 1122. [13] M. Uhlmann., Dynamics of diatomic molecules in intense laser elds. Technishe Universitat Dresden. Ph.D Thesis., 2006. [14] G. Floquet Ann. Ec. Norm. Suppl., vol. 12, p. 47, 1887. References 165 [15] J. H. Shirley, “Solution of the schrödinger equation with a hamilton periodic in time,” Phys. Rev., vol. 138, p. B979, 1965. [16] A. D. Bandrauk and M. L. Sink, “Photoissociation in intense laser fields: Predissociation analogy.,” J.Chem.Phys., vol. 74, p. 1110, 1981. [17] V. Roudnev and B. D. Esry, “General theory of carrier-envelope phase effects,” Phys. Rev. Lett., vol. 99, p. 220406, 2007. [18] L. J. Frasinski, J. H. Posthumus, J. Plumridge, K. Codling, P. F. Taday, and A. J. Langley, “Manipulation of bond hardening in H2+ by chirping of intense femtosecond laser pulses,” Phys. Rev. Lett., vol. 83, p. 3625, 1999. [19] P. H. Bucksbaum, A. Zavriyev, H. G. Muller, and D. W. Schumacher, “Softening of the H2 + molecular bond in intense laser fields,” Phys. Rev. Lett., vol. 64, p. 1883, 1990. [20] A. Zavriyev, P. H. Bucksbaum, H. G. Muller, and D. W. Schumacher, “Ionization and dissociation of h2 in intense laser fields at 1.064 µm, 532 nm, and 355 nm,” Phys. Rev. A., vol. 42, p. 5500, 1990. [21] J. H. Posthumus, L. J. Frasinski, A. J. Giles, and K. Codling, “Dissociative ionization of molecules in intense laser fields: a method of predicting ion kinetic energies and appearance intensities,” J. Phys.B., vol. 28, p. L349, 1995. [22] A. Giusti-Suzor, X. He, O. Atabek, and F. H. Mies, “Above-threshold dissociation of H2 + in intense laser fields,” Phys. Rev. Lett., vol. 64, p. 515, 1990. [23] J. McKenna, A. M. Sayler, F. Anis, B. Gaire, N. G. Johnson, E. Parke, J. J. Hua, H. Mashiko, C. M. Nakamura, E. Moon, Z. Chang, K. D. Carnes, B. D. Esry, and I. Ben-Itzhak, “Enhancing high-order above-threshold dissociation of H2 + beams with few-cycle laser pulses,” Phys. Rev. Lett., vol. 100, p. 133001, 2008. [24] P. A. Orr, I. D. Williams, J. B. Greenwood, I. C. E. Turcu, W. A. Bryan, J. Pedregosa-Gutierrez, and C. W. Walter, “Above threshold dissociation of vibrationally cold HD+ molecules,” Phys. Rev. Lett., vol. 98, p. 163001, 2007. [25] J. McKenna, A. M. Sayler, F. Anis, N. G. Johnson, B. Gaire, U. Lev, M. A. Zohrabi, K. D. Carnes, B. D. Esry, and I. Ben-Itzhak, “Vibrationally cold CO2+ in intense ultrashort laser pulses,” Phys. Rev. A, vol. 81, p. 061401, 2010. [26] E. E. Aubanel, J.-M. Gauthier, and A. D. Bandrauk, “Molecular stabilization and angular distribution in photodissociation of H2 + in intense laser fields,” Phys. Rev. A., vol. 48, p. 2145, 1993. [27] A. Giusti-Suzor and F. H. Mies, “Vibrational trapping and suppression of dissociation in intense laser fields,” Phys. Rev. Lett., vol. 68, p. 3869, 1992. References 166 [28] A. Zavriyev, P. H. Bucksbaum, J. Squier, and F. Saline, “Light-induced vibrational structure in H2 + and D2 + in intense laser fields,” Phys. Rev. Lett., vol. 70, p. 1077, 1993. [29] F. Anis and B. D. Esry, “Role of nuclear rotation in dissociation of H2 + in a short laser pulse,” Phys. Rev. A, vol. 77, p. 033416, 2008. [30] M. Protopapas, C. H. Keitel, and P. L. Knight, “Atomic physics with super-high intensity lasers,” Rep. Prog. Phys., vol. 60, no. 389, 1997. [31] M. Goppert-Mayer., “Elementary processes with two quantum jumps,” Ann. der. Phys. (Leipzig)., vol. 9, p. 273. [32] M. V. Ammosov, N. B. Delone, and V. P. Krainov, “Tunnel ionisation of complex atoms and of atomic ions in an alternating electromagnetic field,” Sov. Phys. JEPT., vol. 64, p. 1191. [33] M. J. DeWitt and R. J. Levis, “Calculating the keldysh adiabaciity parameter for atomic, diatomic and polyatomic molecules,” J. Chem. Phys, vol. 108, no. 7739. [34] T. Nubbemeyer, K. Gorling, A. Saenz, U. Eichmann, and W. Sandner, “Strongfield tunneling without ionisation,” Phys. Rev. Lett., p. 233001, 2008. [35] B. Manschwetus, T. Nubbemeyer, K. Gorling, G. Steinmeyer, U. . Eichmann, H. Rottke, and W. Sandner, “Strong laser field fragmentation of H2 + : Coulomb explosion withough double ionisation,” Phys. Rev. Letter., p. 113002, 2009. [36] A. L’Huillier, L. A. Lompre, G. Mainfrey, and C. Manus, “Multiply charged ions induced by multiphoton absoorption in rare gases at 0.53 micrometers,” Phys. Rev. A., vol. 27, p. 2503, 1994. [37] D. N. Fittinghoff, P. R. Bolton, B. Chang, and K. C. Kulander, “Observation of non-sequential double ionisation of helium with optical tunneling,” Phys. Rev. Lett., vol. 69, p. 2642, 1997. [38] U. Eichmann, M. Dorr, H. Madea, W. Becker, and W. Sandner, “Collective multielectron tunneling ionisation in strong fields,” Phys. Rev. Lett., vol. 84, p. 3550, 2000. [39] B. Walker, E. Mevel, B. Yang, P. Breger, J. P. Chambaret, A. Antonetti, L. F. DiMauro, and P. Agostini, “Double ionization in the perturbative and tunneling regimes,” Phys. Rev. A, vol. 48, p. R894, 1993. [40] L. F. DiMauro and P. Agostini, “Ionisation dynamics in strong laser fields,” Adv. At. Mol. Opt. Phys.,, vol. 35, p. 79, 1995. [41] P. B. Corkum, “Plasma perspective on strong field multiphoton ionisation,” Phys. Rev. Lett., vol. 35, p. 79. References 167 [42] T. Pfeifer, C. Spielmann, and G. Gerber, “Femtosecond x-ray science,” Rep. Prog. Phys., vol. 69. [43] K. Boyer, T. S. Luk, J. C. Solem, and C. K. Rhodes, “Kinetic energy distributions of ionic fragments produced by subpicosecond multiphoton ionization of N2 ,” Phys. Rev. A, vol. 39, p. 1186, 1989. [44] R. S. Muliken, “Intensities of electronic transitions in molecular spectra ii. charge transfer spectra,” J. Chem. Phys., vol. 7, 1939. [45] B. Gaire, J. McKenna, N. G. Johnson, A. M. Sayler, E. Parke, K. D. Carnes, and I. Ben-Itzhak, “Laser-induced multiple ionization of molecular ion beams: N2 + , CO+ , NO+ and O2 + ,” Phys. Rev. A, vol. 79, p. 063414, 2009. [46] F. Légaré, I. V. Litvinyuk, P. W. Dooley, F. Quéré, A. D. Bandrauk, D. M. Villeneuve, and P. B. Corkum, “Time-resolved double ionization with few cycle laser pulses,” Phys. Rev. Lett., vol. 91, p. 093002, 2003. [47] P. B. Corkum, “Plasma perspective on strong field multiphoton ionization,” Phys. Rev. Lett., vol. 71, p. 1994, 1993. [48] T. Zuo and A. D. Bandrauk, “Charge-resonance-enhanced ionization of diatomic molecular ions by intense lasers,” Phys. Rev. A., vol. 52, p. R2511, 1995. [49] K. Codling and L. J. Frasinski, “Dissociative ionization of small molecules in intense laser fields,” J. Phys.B., vol. 26, p. 783. [50] J. H. Posthumus, A. J. Giles, M. R. Thompson, and K. Codling, “Fieldionization, coulomb explosion of diatomic molecules in intense laser fields,” J. Phys.B., vol. 29, p. 5811, 1996. [51] H. Stapelfeldt and T. Seideman, “Colloquium: Aligning molecules with strong laser pulses,” Rev. Mod. Phys, vol. 75, p. 543, 2002. [52] B. Friedrich and D. Herschbach, “Alignment and trapping of molecules in intense laser fileds,” Phys. Rev. Lett, vol. 74, p. 4623, 1995. [53] J. Ortigoso, M. Rodriguez, M. Gupta, and B. Friedrich, “Time evolution of pendular states created by the interaction of molecular polarizability with a pulsed nonresonant laser field,” J. Chem. Phys., vol. 110, p. 3870, 1999. [54] G. N. Gibson, M. Li, C. Guo, and J. Neira, “Strong-field dissociation and ionization of H2 + using ultrashort laser pulses,” Phys. Rev. Lett., vol. 79, p. 2022, 1997. [55] A. Rudenko, B. Feuerstein, K. Zrost, V. L. B. de Jesus, T. Ergler, C. Dimopoulou, C. D. Schroter, R. Moshammer, and J. Ullrich, “Fragmentation dynamics of molecular hydrogen in strong ultrashort laser pulses,” J. Phys.B, vol. 38, p. 487, 2005. References 168 [56] X. Urbain, B. Fabre, E. M. Staicu-Casagrande, N. de Ruette, V. M. Andrianarijaona, J. Jureta, J. H. Posthumus, A. Saenz, E. Baldit, and C. Cornaggia, “Intense-laser-field ionization of molecular hydrogen in the tunneling regime and its effect on the vibrational excitation of H2 + ,” Phys. Rev. Lett., vol. 92, p. 163004, 2004. [57] A. Requate, A. Becker, and F. H. M. Faisal, “s-matrix theory of inelastic vibronic ionization of molecules in intense laser fields,” Phys. Rev. A, vol. 73, p. 033406, 2006. [58] I. D. Williams, P. McKenna, B. Srigengan, I. M. G. Johnston, W. A. Bryan, J. H. Sanderson, A. El-Zein, T. R. J. Goodworth, W. R. Newell, P. F. Taday, and A. J. Langley, “Fast-beam study of H2 + ions in an intense femtosecond laser field,” J. Phys. B., vol. 33, p. 2743, 2000. [59] Z. Amitay, A. Baer, M. Dahan, J. Levin, Z. Vager, D. Zajfman, L. Knoll, M. Lange, D. Schwalm, R. Wester, A. Wolf, I. F. Schneider, and A. Suzor-Weiner, “Dissociative recombination of vibrationally excited HD+ : State-selective experimental investigation,” Phys. Rev. A., vol. 60, 1999. [60] G. H. Dunn, “Franck condon factors for the ionization of H2 + and D2 + ,” J. Chem. Phys., vol. 44, p. 2592, 1966. [61] F. von Busch and G. H. Dunn, “Photodissociation of H2 + and D2 + : Experiment,” Phys. Rev. A., vol. 5, p. 1726, 1972. [62] A. A. A. El-Zein, P. McKenna, Bryan, I. M. G. Johnston, T. R. J. Goodworth, J. H. Sanderson, I. D. Williams, W. R. Newell, P. F. Taday, E. J. Divall, and A. J. Langley, “A detailed study of multiply charged ion production within a high intensity laser focus,” Physica. Scripta., vol. T92, p. 119, 2001. [63] W. A. Bryan, S. L. Stebbings, E. M. L. English, T. R. J. Goodworth, W. R. Newell, J. McKenna, M. Suresh, B. Srigengan, I. D. Williams, I. C. E. Turcu, J. M. Smith, E. J. Divall, C. J. Hooker, and A. J. Langley, “Geometry- and diffraction-independent ionization probabilities in intense laser fields: Probing atomic ionization mechanisms with effective intensity matching,” Phys. Rev. A., vol. 73, p. 013407, 2006. [64] P. Hansch, M. A. Walker, and L. D. Van Woerkom, “Spatially dependent multiphoton multiple ionization,” Phys. Rev. A., vol. 54, p. R2559, 1996. [65] P. Wang, A. M. Sayler, K. D. Carnes, B. D. Esry, and I. Ben-Itzhak, “Disentangling the volume effect through intensity-difference spectra: Application to laser-induced dissociation of H2 + ,” Opt. Lett., vol. 30, p. 6, 2005. [66] P. Q. Wang, A. M. Sayler, K. D. Carnes, J. F. Xia, M. A. Smith, B. D. Esry, and I. Ben-Itzhak, “Dissociation of H2 + in intense femtosecond laser fields studied References 169 by coincidence three-dimensional momentum imaging,” Phys. Rev. A, vol. 74, p. 043411, 2006. [67] V. S. Prabhudesai, U. Lev, A. Natan, B. D. Bruner, A. Diner, O. Heber, D. Strasser, D. Schwalm, I. Ben-Itzhak, J. J. Hua, B. D. Esry, Y. Silberberg, and D. Zajfman, “Tracing the photodissociation probability of H2 + in intense fields using chirped laser pulses,” Phys. Rev. A., vol. 81, p. 023401, 2010. [68] J. McKenna, “Probing atomic and molecular dynamics with intense femtosecond laser pulses,” Queens University Belfast, P.hD Thesis., 2006. [69] W. C. Wiley and I. H. McLaren, “Time-of-flight mass spectrometer with improved resolution,” Rev. Sci. Instrum., vol. 26, p. 1150, 1955. [70] L. J. Frasinski, K. Codling, and P. A. Hatherly, “Covariance mapping: A correlation method applied to multiphoton muliple ionisation,” Science, vol. 246, p. 1029, 1989. [71] D. A. Carda, E. S. Wisniewski, D. Folmer, and A. W. Castleman-Jr, “The relationship between covariance and anti-covariance mapping,” International Journal of Mass Spectrometry., vol. 223, p. 355, 2003. [72] K. Sändig, H. Figger, and T. W. Hänsch, “Dissociation dynamics of H2 + in intense laser fields: Investigation of photofragments from single vibrational levels,” Phys. Rev. Lett., vol. 85, p. 4876, 2000. [73] A. Monmayrant, S. Weber, and B. Chatel, “A newcomers guide to ultrashort pulse shaping and characterization,” J. Phys. B., vol. 43, p. 34, 2010. [74] F. Träger, Springer Handbook of Lasers and Optics. Springer Link, 2007. [75] A. H. Zewail, “Laser femtochemistry,” Science., vol. 242, p. 1645, 1988. [76] H. Rabitz, “Shaped laser pulses as reagents,” Science., vol. 299. [77] A. H. Zewail, “Femtochemistry. past, present, and future,” Pure Appl. Chem., vol. 72, p. 2210, 2000. [78] C. Brif, R. Chakrabarti, and H. Rabitz, “Control of quantum phenomena: past, present and future,” New. J. Phys., vol. 12, p. 075008, 2010. [79] M. Brumer, P. W. amd Shapiro, Principles of the Quantum Control of Moecular Processes. Wiley-Interscience Hoboken NJ, USA,, 2000. [80] S. A. Rice and M. Zhao, Optical Control of Molecular Dynamics. Interscience New York USA,, 2000. Wiley- [81] S. R. Judson and H. Rabitz, “Teaching lasers to control molecules,” Phys. Rev. Lett., vol. 68, p. 1500, 1992. References 170 [82] T. C. Weinacht, J. Ahn, and P. H. Bucksbaum, “Controlling the shape of a quantum wavefunction,” nature., vol. 397, p. 233, 1998. [83] G. Vogt, G. Krampert, P. Niklaus, P. Nuernberger, , and G. Gerber, “Optimal control of photoisomerization,” Phys. Rev. Lett., vol. 94, p. 068305, 2005. [84] R. Levis and H. Rabitz, “Closing the loop on bond selective chemistry using tailored strong field laser pulses,” J. Phys. Chem., vol. 106, p. 6427, 2002. [85] R. J. Levis, G. M. Menkir, and H. Rabitz, “Selective bond dissociation and rearrangement with optimally tailored, strong-field laser pulses,” Science, vol. 292, p. 27, 2001. [86] T. Brixner, N. H. Damrauer, P. Niklaus, and G. Gerber, “Photoselective adaptive femtosecond quantum control in the liquid phase,” Nature., vol. 414, 2001. [87] G. Y. Chen, Z. W. Wang, and W. T. Hill, “Adaptive control of the CO2 bending vibration: Deciphering field-system dynamics,” Phys. Rev. A., vol. 79, p. 011401, 2009. [88] M. F. Kling, C. Siedschlag, A. J. Verhoef, J. I. Khan, M. Schultze, T. Uphues, Y. Ni, M. Uiberacker, M. Drescher, F. Krausz, and M. J. J. Vrakking, “Control of electron localization in molecular dissociation,” Science., vol. 312, p. 246, 2006. [89] A. Assion, T. Baumert, M. Bergt, T. Brixner, B. Kiefer, V. Seyfried, M. Strehle, and G. Gerber, “Control of chemical reactions by feedback-optimized phaseshaped femtosecond laser pulses,” Science., vol. 282, 1998. [90] K. O. Nielsen, “The development of magnetic ion sources for an electromagnetic isotope separator,” Nucl. Instrum., vol. 1, 1957. [91] J. J. Barroso and M. O. Terra, “Beam transport in a quadrupole electrostatic system,” Brazilian Journal of Physics., vol. 34, 2004. [92] “Electrostatic charged particle optics,” Private Communications., 2009. [93] M. Marynowski, W. Franzen, , and M. El-Batanouny, “Analysis of the properties of an electrostatic triplet quadrupole lens used as and electron beam transport device,” Rev. Sci. Instrum., vol. 65, p. 3718, 1994. [94] A. Stingl, C. Spielmann, and F. Krausz, “Generation of 11-fs pulses from a ti:sapphire laser without the use of prisms,” Opt. Lett., vol. 19, p. 3, 1994. [95] D. Strickland and G. Mourou, “Compression of amplified chirped optical pulses,” Opt. Commun., vol. 56, p. 219, 1985. [96] E. Wintner, F. Krausz, M. Lenzner, C. Spielmann, and A. J. Schmidt, “Sub20-fs, kilohertz-repetition-rate ti:sapphire amplifier,” Opt. Lett., vol. 20, p. 1397, 1995. References 171 [97] P. O. Shea, M. Kimmel, X. Gu, and R. Trebino, “Highly simplified device for ultrashort-pulse measurement,” Opt.Lett, vol. 26, p. 932, 2001. [98] K. W. DeLong, R. Trebino, and D. J. Kane, “Comparison of ultrashort-pulse frequency-resolved-optical-gating traces for three common beam geometries,” J. Opt. Soc. Am. B, vol. 11, no. 9, p. 1595, 1994. [99] M. A. Krumbugel, J. N. Sweetser, D. N. Fittinghoff, K. W. DeLong, and R. Trebino, “Ultrafast optical switching by use of fully phase-matched cascaded secondorder nonlinearities in a polarization-gate geometry,,” Opt. Lett., vol. 22, p. 245, 1997. [100] M. Zohrabi, L. Graham, U. Lev, U. Ablikim, B. D. Bruner, J. J. Hua, J. McKenna, K. J. Betsch, B. Jochim, A. M. Summers, B. Berry, D. Strasser, O. Heber, Y. Silberberg, D. Zajfman, K. D. Carnes, B. D. Esry, and I. BenItzhak, “Dissociation of H2 + by broad bandwidth laser pulses: selective molecular response causing effective bandwidth narrowing,” Phys. Rev. A, vol. to be submitted, 2013. [101] I. Ben-Itzhak, P. Q. Wang, J. F. Xia, A. M. Sayler, M. A. Smith, K. D. Carnes, and B. D. Esry, “Dissociation and ionization of H2 + by ultrashort intense laser pulses probed by coincidence 3d momentum imaging,” Phys. Rev. Lett., vol. 95, p. 073002. [102] C. Zhang and X. Miao, “Effects of the chirped pulse on dissociation and ionization processes for hydrogen molecular ions at low vibrational states,” Spectroscopy Letters, vol. 84, p. 013408, 2011. [103] Y. B. Band and P. S. Julienne, “Molecular population transfer, alignment, and orientation using chirped pulse absorption,” J. Chem. Phys., vol. 97, p. 9107, 1992. [104] I. Pastirk, E. J. Brown, Q. Zhang, and M. Dantus, “Molecular population transfer, alignment, and orientation using chirped pulse absorption,” [105] Y. Liu, Y. Liu, and Q. Gong, “Control of landau-zener transition in nai predissociation with chirped femtosecond laser pulses,” EPL., vol. 101, p. 68006, 2013. [106] U. Lev, V. Prabhudesai, A. Natan, B. Bruner, A. Diner, O. Heber, D. Strasser, D. Schwalm, I. Ben-Itzhak, J. J. Hua, B. D. Esry, Y. Silberberg, and D. Zajfman, “Photodissociation of H2 + by intense chirped pulses beyond the effect of pulse duration and peak power,” J. Phys.: Conf. Ser., vol. 194, p. 032060, 2009. [107] A. Natan, U. Lev, V. S. Prabhudesai, B. D. Bruner, D. Strasser, D. Schwalm, I. Ben-Itzhak, O. Heber, D. Zajfman, and Y. Silberberg, “Quantum control of photodissociation by manipulation of bond softening,” Phys. Rev. A, vol. 86, p. 043418, 2012. References 172 [108] J. McKenna, A. M. Sayler, B. Gaire, N. G. Johnson, M. Zohrabi, K. D. Carnes, B. D. Esry, and I. Ben-Itzhak., “Permanent dipole transitions remain elusive in HD+ strong eld dissociation,” J. Phys. B., vol. 40. [109] I. Ben-Itzhak Private Communications. [110] S. Zhang, C. Lu, J. Shi, T. Jia, Z. Wang, and Z. Sun, “Field-free molecular alignment by shaping femtosecond laser pulse with cubic phase modulation,” Phys. Rev. A., vol. 84, p. 013408. [111] J. McKenna, M. Suresh, D. S. Murphy, W. A. Bryan, L.-Y. Peng, S. L. Stebbings, E. M. L. English, J. Wood, B. Srigengan, I. C. E. Turcu, J. L. Collier, J. F. McCann, W. R. Newell, and I. D. Williams, “Intense-field dissociation dynamics of D2 + molecular ions using ultrafast laser pulses,” J. Phys. B., vol. 40. [112] R. Torres, R. de Nalda, and J. P. Marangos, “Dynamics of laser-induced molecular alignment in the impulsive and adiabatic regimes: A direct comparison,” Phys. Rev. A., vol. 72, p. 023420, 2005. [113] J. P. Cryan, P. H. Bucksbaum, H. Philip, and R. N. Coffee, “Field-free alignment in repetitively kicked nitrogen gas,” Phys. Rev. A., vol. 80, p. 063412, 2009. [114] I. S. Averbukh and R. Arvieu, “Angular focusing, squeezing, and rainbow formation in a strongly driven quantum rotor,” Phys. Rev. Lett., vol. 87, p. 163601. [115] J. J. Hua and B. D. Esry, “Isotopic pulse-length scaling of molecular dissociation in an intense laser field,” Phys. Rev. A, vol. 78, p. 055403, 2008. [116] J. J. Hua, “Isotopic effects in H2 + dynamics in an intense laser field,” Kansas State University, PhD Thesis., 2009. [117] T. C. Gunaratne, X. Zhu, V. V. Lozovoy, and M. Dantus, “Influence of the temporal shape of femtosecond pulses on silicon micromachining,” J. Appl. Phys., vol. 106, 2009. [118] A. Hishikawa, A. Iwamae, and K. Yamanouchi, “Ultrafast structural deformation of no2 in intense laser fields studied by mass-resolved momentum imaging,” J. Chem. Phys., vol. 111, p. 8871, 1999. [119] A. S. Alnaser, S. Voss, X. M. Tong, C. M. Maharjan, P. Ranitovic, B. Ulrich, T. Osipov, B. Shan, Z. Chang, and C. L. Cocke, “Effects of molecular structure on ion disintegration patterns in ionization of o2 and n2 by short laser pulses,” Phys. Rev. Lett., vol. 93, p. 113003, 2004. [120] A. M. Sayler, P. Q. Wang, K. D. Carnes, B. D. Esry, and I. Ben-Itzhak, “Determining laser-induced dissociation pathways of multielectron diatomic molecules: Application to the dissociation of o+ 2 by high-intensity ultrashort pulses,” Phys. Rev. A, vol. 75, p. 063420, 2007. References 173 [121] J. McKenna, A. M. Sayler, B. Gaire, N. G. Johnson, E. Parke, K. D. Carnes, B. D. Esry, and I. Ben-Itzhak, “Intensity dependence in the dissociation branching ratio of ND+ using intense femtosecond laser pulses,” Phys. Rev. A, vol. 77, p. 063422, 2008. [122] M. V. Korolkov and K. Weitzel, “On the control of product yields in the photofragmentation of deuteriumchlorid ions (DCL+ ),” J. Chem. Phys., vol. 123, p. 164308, 2005. [123] N. G. Kling, J. McKenna, A. M. Sayler, B. Gaire, M. Zohrabi, U. Ablikim, K. D. Carnes, and I. Ben-Itzhak, “Charge asymmetric dissociation of a co+ molecularion beam induced by strong laser fields,” Phys. Rev. A, vol. 87, p. 013418, 2013.