Photoelectron photoion coincidence spectroscopy

Last updated

Photoelectron photoion coincidence spectroscopy (PEPICO) is a combination of photoionization mass spectrometry and photoelectron spectroscopy. [1] It is largely based on the photoelectric effect. Free molecules from a gas-phase sample are ionized by incident vacuum ultraviolet (VUV) radiation. In the ensuing photoionization, a cation and a photoelectron are formed for each sample molecule. The mass of the photoion is determined by time-of-flight mass spectrometry, whereas, in current setups, photoelectrons are typically detected by velocity map imaging. Electron times-of-flight are three orders of magnitude smaller than those of ions, which allows electron detection to be used as a time stamp for the ionization event, starting the clock for the ion time-of-flight analysis. In contrast with pulsed experiments, such as REMPI, in which the light pulse must act as the time stamp, this allows to use continuous light sources, e.g. a discharge lamp or a synchrotron light source. No more than several ion–electron pairs are present simultaneously in the instrument, and the electron–ion pairs belonging to a single photoionization event can be identified and detected in delayed coincidence.

Contents

History

A painting featuring the PEPICO endstation at the Swiss Light Source. IPEPICO by Jonelle Harvey.jpg
A painting featuring the PEPICO endstation at the Swiss Light Source.

Brehm and von Puttkammer published the first PEPICO study on methane in 1967. [2] In the early works, a fixed energy light source was used, and the electron detection was carried out using retarding grids or hemispherical analyzers: the mass spectra were recorded as a function of electron energy. Tunable vacuum ultraviolet light sources were used in later setups, [3] [4] in which fixed, mostly zero kinetic energy electrons were detected, and the mass spectra were recorded as a function of photon energy. Detecting zero kinetic energy or threshold electrons in threshold photoelectron photoion coincidence spectroscopy, TPEPICO, has two major advantages. Firstly, no kinetic energy electrons are produced in energy ranges with poor Franck–Condon factors in the photoelectron spectrum, but threshold electrons can still be emitted via other ionization mechanisms. [5] Secondly, threshold electrons are stationary and can be detected with higher collection efficiencies, thereby increasing signal levels.

Threshold electron detection was first based on line-of-sight, i.e. a small positive field was applied towards the electron detector, and kinetic energy electrons with perpendicular velocities are stopped by small apertures. [6] The inherent compromise between resolution and collection efficiency was resolved by applying velocity map imaging [7] conditions. [8] Most recent setups offer meV or better (0.1 kJ mol−1) resolution both in terms of photon energy and electron kinetic energy. [9] [10]

The 5–20 eV (500–2000 kJ mol−1, λ = 250–60 nm) energy range is of prime interest in valence photoionization. Widely tunable light sources are few and far between in this energy range. The only laboratory based one is the H2 discharge lamp, which delivers quasi-continuous radiation up to 14 eV. [11] The few high resolution laser setups for this energy range are not easily tunable over several eV. Currently, VUV beamlines at third generation synchrotron light sources are the brightest and most tunable photon sources for valence ionization. The first high energy resolution PEPICO experiment at a synchrotron was the pulsed-field ionization setup at the Chemical Dynamics Beamline of the Advanced Light Source. [12]

Overview

Velocity map imaging photoelectron photoion coincidence apparatus. Electrons with different kinetic energies are shown as well as ions with a room temperature kinetic energy distribution. Scheme - Photoelectron photoion coincidence apparatus.png
Velocity map imaging photoelectron photoion coincidence apparatus. Electrons with different kinetic energies are shown as well as ions with a room temperature kinetic energy distribution.

The primary application of TPEPICO is the production of internal energy selected ions to study their unimolecular dissociation dynamics as a function of internal energy. The electrons are extracted by a continuous electric field and are velocity map imaged depending on their initial kinetic energy. Ions are accelerated in the opposite direction and their mass is determined by time-of-flight mass spectrometry. The data analysis yields dissociation thresholds, which can be used to derive new thermochemistry for the sample. [13]

The electron imager side can also be used to record photoionization cross sections, photoelectron energy and angular distributions. With the help of circularly polarized light, photoelectron circular dichroism (PECD) can be studied. [14] A thorough understanding of PECD effects could help explain the homochirality of life. [15] Flash pyrolysis can also be used to produce free radicals or intermediates, which are then characterized to complement e.g. combustion studies. [16] [17] In such cases, the photoion mass analysis is used to confirm the identity of the radical produced.

Photoelectron photoion coincidence spectroscopy can be used to shed light on reaction mechanisms, [18] and can also be generalized to study double ionization in (photoelectron) photoion photoion coincidence ((PE)PIPICO), [19] fluorescence using photoelectron photon coincidence (PEFCO), [20] or photoelectron photoelectron coincidence (PEPECO). [21] Times-of-flight of photoelectrons and photoions can be combined in a form of a map, which visualizes the dynamics of the dissociative ionization process. [22] Ion–electron velocity vector correlation functions can be obtained in double imaging setups, in which the ion detector also delivers position information. [23]

Energy selection

Potential energy diagram for dissociative photoionization. When only zero kinetic energy electrons are detected, the photon energy above the adiabatic ionization energy is converted into the internal energy of the photoion AB PEC Scheme.tif
Potential energy diagram for dissociative photoionization. When only zero kinetic energy electrons are detected, the photon energy above the adiabatic ionization energy is converted into the internal energy of the photoion AB

The relatively low intensity of the ionizing VUV radiation guarantees one-photon processes, in other words only one, fixed energy photon will be responsible for photoionization. The energy balance of photoionization comprises the internal energy and the adiabatic ionization energy of the neutral as well as the photon energy, the kinetic energy of the photoelectron and of the photoion. Because only threshold electrons are considered and the conservation of momentum holds, the last two terms vanish, and the internal energy of the photoion is known:

Scanning the photon energy corresponds to shifting the internal energy distribution of the parent ion. The parent ion sits in a potential energy well, in which the lowest energy exit channel often corresponds to the breaking of the weakest chemical bond, resulting in the formation of a fragment or daughter ion. A mass spectrum is recorded at every photon energy, and the fractional ion abundances are plotted to obtain the breakdown diagram. At low energies no parent ion is energetic enough to dissociate, and the parent ion corresponds to 100% of the ion signal. As the photon energy is increased, a certain fraction of the parent ions (in fact according to the cumulative distribution function of the neutral internal energy distribution) still has too little energy to dissociate, but some do. The parent ion fractional abundances decrease, and the daughter ion signal increases. At the dissociative photoionization threshold, E0, all parent ions, even the ones with initially 0 internal energy, can dissociate, and the daughter ion abundance reaches 100% in the breakdown diagram.

If the potential energy well of the parent ion is shallow and the complete initial thermal energy distribution is broader than the depth of the well, the breakdown diagram can also be used to determine adiabatic ionization energies. [24]

Data analysis

The data analysis becomes more demanding if there are competing parallel dissociation channels or if the dissociation at threshold is too slow to be observed on the time scale (several μs) of the experiment. In the first case, the slower dissociation channel will appear only at higher energies, an effect called competitive shift, whereas in the second, the resulting kinetic shift means that the fragmentation will only be observed at some excess energy, i.e. only when it is fast enough to take place on the experimental time scale. When several dissociation steps follow sequentially, the second step typically occurs at high excess energies: the system has much more internal energy than needed for breaking the weakest bond in the parent ion. Some of this excess energy is retained as internal energy of the fragment ion, some may be converted into the internal energy of the leaving neutral fragment (invisible to mass spectrometry) and the rest is released as kinetic energy, in that the fragments fly apart at some non-zero velocity.

More often than not, dissociative photoionization processes can be described within a statistical framework, similarly to the approach used in collision-induced dissociation experiments. If the ergodic hypothesis holds, the system will explore each region of the phase space with a probability according to its volume. A transition state (TS) can then be defined in the phase space, which connects the dissociating ion with the dissociation products, and the dissociation rates for the slow or competing dissociations can be expressed in terms of the TS phase space volume vs. the total phase space volume. The total phase space volume is calculated in a microcanonical ensemble using the known energy and the density of states of the dissociating ion. There are several approaches how to define the transition state, the most widely used being RRKM theory. The unimolecular dissociation rate curve as a function of energy, k(E), vanishes below the dissociative photoionization energy, E0. [25]

Statistical theory can also be used in the microcanonical formalism to describe the excess energy partitioning in sequential dissociation steps, as proposed by Klots [26] for a canonical ensemble. Such a statistical approach was used for more than a hundred systems to determine accurate dissociative photoionization onsets, and derive thermochemical information from them. [27]

Furthermore, algorithms based on probabilistic Bayesian analyses are known to considerably reduce systematic biases induced by false coincidences. The intensity of these false coincidences can big strong enough to appear as a separate peaks in the signal and complicate the analysis of the spectra. [28]

Thermochemical applications

Dissociative photoionization processes can be generalized as:

AB + A+ + B + e

If the enthalpies of formation of two of the three species are known, the third can be calculated with the help of the dissociative photoionization energy, E0, using Hess's law. This approach was used, for instance, to determine the enthalpy of formation of the methyl ion, CH3+, [29] which in turn was used to obtain the enthalpy of formation of iodomethane, CH3I as 15.23 kJ mol−1, with an uncertainty of only 0.3 kJ mol−1. [30]

If different sample molecules produce shared fragment ions, a complete thermochemical chain can be constructed, as was shown for some methyl trihalides, [31] where the uncertainty in e.g. the CHCl2Br, (Halon-1021) heat of formation was reduced from 20 to 2 kJ mol−1. Furthermore, dissociative photoionization energies can be combined with calculated isodesmic reaction energies to build thermochemical networks. Such an approach was used to revise primary alkylamine enthalpies of formation. [32]

See also

Related Research Articles

<span class="mw-page-title-main">Auger effect</span> Physical phenomenon

The Auger effect or Auger−Meitner effect is a physical phenomenon in which the filling of an inner-shell vacancy of an atom is accompanied by the emission of an electron from the same atom. When a core electron is removed, leaving a vacancy, an electron from a higher energy level may fall into the vacancy, resulting in a release of energy. For light atoms (Z<12), this energy is most often transferred to a valence electron which is subsequently ejected from the atom. This second ejected electron is called an Auger electron. For heavier atomic nuclei, the release of the energy in the form of an emitted photon becomes gradually more probable.

<span class="mw-page-title-main">Photoelectric effect</span> Emission of electrons when light hits a material

The photoelectric effect is the emission of electrons from a material caused by electromagnetic radiation (light). Electrons emitted in this manner are called photoelectrons. The phenomenon is studied in condensed matter physics, solid state, and quantum chemistry to draw inferences about the properties of atoms, molecules and solids. The effect has found use in electronic devices specialized for light detection and precisely timed electron emission.

<span class="mw-page-title-main">Ionization</span> Process by which atoms or molecules acquire charge by gaining or losing electrons

Ionization is the process by which an atom or a molecule acquires a negative or positive charge by gaining or losing electrons, often in conjunction with other chemical changes. The resulting electrically charged atom or molecule is called an ion. Ionization can result from the loss of an electron after collisions with subatomic particles, collisions with other atoms, molecules and ions, or through the interaction with electromagnetic radiation. Heterolytic bond cleavage and heterolytic substitution reactions can result in the formation of ion pairs. Ionization can occur through radioactive decay by the internal conversion process, in which an excited nucleus transfers its energy to one of the inner-shell electrons causing it to be ejected.

<span class="mw-page-title-main">X-ray photoelectron spectroscopy</span> Spectroscopic technique

X-ray photoelectron spectroscopy (XPS) is a surface-sensitive quantitative spectroscopic technique based on the photoelectric effect that can identify the elements that exist within a material or are covering its surface, as well as their chemical state, and the overall electronic structure and density of the electronic states in the material. XPS is a powerful measurement technique because it not only shows what elements are present, but also what other elements they are bonded to. The technique can be used in line profiling of the elemental composition across the surface, or in depth profiling when paired with ion-beam etching. It is often applied to study chemical processes in the materials in their as-received state or after cleavage, scraping, exposure to heat, reactive gasses or solutions, ultraviolet light, or during ion implantation.

<span class="mw-page-title-main">Mass spectrometry</span> Analytical technique based on determining mass to charge ratio of ions

Mass spectrometry (MS), also called mass spec, is an analytical technique that is used to measure the mass-to-charge ratio of ions. The results are presented as a mass spectrum, a plot of intensity as a function of the mass-to-charge ratio. Mass spectrometry is used in many different fields and is applied to pure samples as well as complex mixtures.

<span class="mw-page-title-main">Photoemission spectroscopy</span> Examining a substance by measuring electrons emitted in the photoelectric effect

Photoemission spectroscopy (PES), also known as photoelectron spectroscopy, refers to energy measurement of electrons emitted from solids, gases or liquids by the photoelectric effect, in order to determine the binding energies of electrons in the substance. The term refers to various techniques, depending on whether the ionization energy is provided by X-ray, XUV or UV photons. Regardless of the incident photon beam, however, all photoelectron spectroscopy revolves around the general theme of surface analysis by measuring the ejected electrons.

<span class="mw-page-title-main">Photoionization</span> Ion formation via a photon interacting with a molecule or atom

Photoionization is the physical process in which an ion is formed from the interaction of a photon with an atom or molecule.

Infrared multiple photon dissociation (IRMPD) is a technique used in mass spectrometry to fragment molecules in the gas phase usually for structural analysis of the original (parent) molecule.

<span class="mw-page-title-main">Secondary electrons</span> Electrons generated as ionization products

Secondary electrons are electrons generated as ionization products. They are called 'secondary' because they are generated by other radiation. This radiation can be in the form of ions, electrons, or photons with sufficiently high energy, i.e. exceeding the ionization potential. Photoelectrons can be considered an example of secondary electrons where the primary radiation are photons; in some discussions photoelectrons with higher energy (>50 eV) are still considered "primary" while the electrons freed by the photoelectrons are "secondary".

<span class="mw-page-title-main">Extreme ultraviolet</span> Ultraviolet light with a wavelength of 10–121nm

Extreme ultraviolet radiation or high-energy ultraviolet radiation is electromagnetic radiation in the part of the electromagnetic spectrum spanning wavelengths shorter that the hydrogen Lyman-alpha line from 121 nm down to the X-ray band of 10 nm, and therefore having photons with energies from 10.26 eV up to 124.24 eV. EUV is naturally generated by the solar corona and artificially by plasma, high harmonic generation sources and synchrotron light sources. Since UVC extends to 100 nm, there is some overlap in the terms.

Rydberg ionization spectroscopy is a spectroscopy technique in which multiple photons are absorbed by an atom causing the removal of an electron to form an ion.

This page deals with the electron affinity as a property of isolated atoms or molecules. Solid state electron affinities are not listed here.

Ultraviolet photoelectron spectroscopy (UPS) refers to the measurement of kinetic energy spectra of photoelectrons emitted by molecules which have absorbed ultraviolet photons, in order to determine molecular orbital energies in the valence region.

<span class="mw-page-title-main">Time-of-flight mass spectrometry</span> Method of mass spectrometry

Time-of-flight mass spectrometry (TOFMS) is a method of mass spectrometry in which an ion's mass-to-charge ratio is determined by a time of flight measurement. Ions are accelerated by an electric field of known strength. This acceleration results in an ion having the same kinetic energy as any other ion that has the same charge. The velocity of the ion depends on the mass-to-charge ratio. The time that it subsequently takes for the ion to reach a detector at a known distance is measured. This time will depend on the velocity of the ion, and therefore is a measure of its mass-to-charge ratio. From this ratio and known experimental parameters, one can identify the ion.

Photofragment ion imaging or, more generally, Product Imaging is an experimental technique for making measurements of the velocity of product molecules or particles following a chemical reaction or the photodissociation of a parent molecule. The method uses a two-dimensional detector, usually a microchannel plate, to record the arrival positions of state-selected ions created by resonantly enhanced multi-photon ionization (REMPI). The first experiment using photofragment ion imaging was performed by David W Chandler and Paul L Houston in 1987 on the phototodissociation dynamics of methyl iodide (iodomethane, CH3I).

Appearance energy is the minimum energy that must be supplied to a gas phase atom or molecule in order to produce an ion. In mass spectrometry, it is accounted as the voltage to correspond for electron ionization. This is the minimum electron energy that produces an ion. In photoionization, it is the minimum photon energy of a photon that produces some ion signal. For example, the indene bromide ion (IndBr+) only loses bromine at an incident photon energy of 10.2 eV, so the product, indenyl, has an appearance energy of 10.2 eV.

Photoelectrochemical processes are processes in photoelectrochemistry; they usually involve transforming light into other forms of energy. These processes apply to photochemistry, optically pumped lasers, sensitized solar cells, luminescence, and photochromism.

<span class="mw-page-title-main">Atmospheric-pressure photoionization</span> Soft ionization method

Atmospheric pressure photoionization (APPI) is a soft ionization method used in mass spectrometry (MS) usually coupled to liquid chromatography (LC). Molecules are ionized using a vacuum ultraviolet (VUV) light source operating at atmospheric pressure, either by direct absorption followed by electron ejection or through ionization of a dopant molecule that leads to chemical ionization of target molecules. The sample is usually a solvent spray that is vaporized by nebulization and heat. The benefit of APPI is that it ionizes molecules across a broad range of polarity and is particularly useful for ionization of low polarity molecules for which other popular ionization methods such as electrospray ionization (ESI) and atmospheric pressure chemical ionization (APCI) are less suitable. It is also less prone to ion suppression and matrix effects compared to ESI and APCI and typically has a wide linear dynamic range. The application of APPI with LC/MS is commonly used for analysis of petroleum compounds, pesticides, steroids, and drug metabolites lacking polar functional groups and is being extensively deployed for ambient ionization particularly for explosives detection in security applications.

<span class="mw-page-title-main">Resonance ionization</span> Process to excite an atom beyond its ionization potential to form an ion

Resonance ionization is a process in optical physics used to excite a specific atom beyond its ionization potential to form an ion using a beam of photons irradiated from a pulsed laser light. In resonance ionization, the absorption or emission properties of the emitted photons are not considered, rather only the resulting excited ions are mass-selected, detected and measured. Depending on the laser light source used, one electron can be removed from each atom so that resonance ionization produces an efficient selectivity in two ways: elemental selectivity in ionization and isotopic selectivity in measurement.

References

  1. Baer, Tomas; Booze, Jon; Weitzel, Karl-Michael (February 1991). "Photoelectron Photoion Coincidence Studies of Ion Dissociation Dynamics". In Ng, Cheuk-Yiu (ed.). Vacuum Ultraviolet Photoionization and Photodissociation of Molecules and Clusters. World Scientific Pub Co Inc. pp. 259–296. ISBN   981-02-0430-2.
  2. Brehm, B.; von Puttkammer, E. (1967). "Koinzidensmessung von Photoionen und Photoelektronen bei Methan". Zeitschrift für Naturforschung A. 22 (1): 8. Bibcode:1967ZNatA..22....8B. doi: 10.1515/zna-1967-0103 .
  3. Stockbauer, R. (1973). "Threshold electron-photoion coincidence mass spectrometric study of CH4, CD4, C2H6, and C2D6". Journal of Chemical Physics. 58 (9): 3800–3815. Bibcode:1973JChPh..58.3800S. doi:10.1063/1.1679733.
  4. Werner, AS.; Baer, T. (1975). "Absolute unimolecular decay rates of energy selected C4H6+ metastable ions". Journal of Chemical Physics. 62 (7): 2900–2910. Bibcode:1975JChPh..62.2900W. doi:10.1063/1.430828.
  5. Guyon, P. M.; Baer, Tomas; Nenner, Irene (1983). "Interactions between neutral dissociation and ionization continua in N2O". The Journal of Chemical Physics. 78 (6): 3665. Bibcode:1983JChPh..78.3665G. doi:10.1063/1.445141.
  6. Baer, T.; Peatman, W. B.; Schlag, E. W. (1969). "Photoionization resonance studies with a steradiancy analyzer. II. The photoionization of CH3I". Chemical Physics Letters. 4 (5): 243. Bibcode:1969CPL.....4..243B. doi:10.1016/0009-2614(69)80174-0.
  7. Eppink, A. T. J. B.; Parker, D. H. (1997). "Velocity map imaging of ions and electrons using electrostatic lenses: Application in photoelectron and photofragment ion imaging of molecular oxygen". Review of Scientific Instruments. 68 (9): 3477. Bibcode:1997RScI...68.3477E. doi:10.1063/1.1148310.
  8. SztáRay, B. L.; Baer, T. (2003). "Suppression of hot electrons in threshold photoelectron photoion coincidence spectroscopy using velocity focusing optics". Review of Scientific Instruments. 74 (8): 3763. Bibcode:2003RScI...74.3763S. doi:10.1063/1.1593788.
  9. Garcia, G. A.; Soldi-Lose, H. L. S.; Nahon, L. (2009). "A versatile electron-ion coincidence spectrometer for photoelectron momentum imaging and threshold spectroscopy on mass selected ions using synchrotron radiation". Review of Scientific Instruments. 80 (2): 023102–023102–12. Bibcode:2009RScI...80b3102G. doi:10.1063/1.3079331. PMID   19256635.
  10. Bodi, A.; Johnson, M.; Gerber, T.; Gengeliczki, Z.; SztáRay, B. L.; Baer, T. (2009). "Imaging photoelectron photoion coincidence spectroscopy with velocity focusing electron optics". Review of Scientific Instruments. 80 (3): 034101–034101–7. Bibcode:2009RScI...80c4101B. doi:10.1063/1.3082016. PMID   19334934. S2CID   3920794.
  11. Paresce, F.; Kumar, S.; Bowyer, C. S. (1971). "Continuous Discharge Line Source for the Extreme Ultraviolet". Applied Optics. 10 (8): 1904–1908. Bibcode:1971ApOpt..10.1904P. doi:10.1364/AO.10.001904. PMID   20111225.
  12. Jarvis, G. K.; Weitzel, K. M.; Malow, M.; Baer, T.; Song, Y.; Ng, C. Y. (1999). "High-resolution pulsed field ionization photoelectron-photoion coincidence spectroscopy using synchrotron radiation". Review of Scientific Instruments. 70 (10): 3892. Bibcode:1999RScI...70.3892J. doi:10.1063/1.1150009.
  13. Baer, T.; Sztáray, B. L.; Kercher, J. P.; Lago, A. F.; Bödi, A.; Skull, C.; Palathinkal, D. (2005). "Threshold photoelectron photoion coincidence studies of parallel and sequential dissociation reactions". Physical Chemistry Chemical Physics. 7 (7): 1507–1513. Bibcode:2005PCCP....7.1507B. doi:10.1039/b502051d. PMID   19787975.
  14. Garcia, G. A.; Nahon, L.; Harding, C. J.; Powis, I. (2008). "Chiral signatures in angle-resolved valence photoelectron spectroscopy of pure glycidol enantiomers". Physical Chemistry Chemical Physics. 10 (12): 1628–1639. Bibcode:2008PCCP...10.1628G. doi:10.1039/b714095a. PMID   18338063.
  15. Nahon, L.; Garcia, G. A.; Harding, C. J.; Mikajlo, E.; Powis, I. (2006). "Determination of chiral asymmetries in the valence photoionization of camphor enantiomers by photoelectron imaging using tunable circularly polarized light". The Journal of Chemical Physics. 125 (11): 114309. Bibcode:2006JChPh.125k4309N. doi:10.1063/1.2336432. PMID   16999476.
  16. Fischer, I.; Schüßler, T.; Deyerl, H. J. R.; Elhanine, M.; Alcaraz, C. (2007). "Photoionization and dissociative photoionization of the allyl radical, C3H5". International Journal of Mass Spectrometry. 261 (2–3): 227. Bibcode:2007IJMSp.261..227F. doi:10.1016/j.ijms.2006.09.023.
  17. Steinbauer, M.; Hemberger, P.; Fischer, I.; Bodi, A. (2011). "Photoionization of C7H6 and C7H5: Observation of the Fulvenallenyl Radical". ChemPhysChem. 12 (10): 1795–1797. doi:10.1002/cphc.201000892. PMID   21132691.
  18. Ferrier, B.; Boulanger, A. M.; Holland, D.; Shaw, D.; Mayer, P. (2009). "Nitro–nitrite isomerization and transition state switching in the dissociation of ionized nitromethane: A threshold photoelectron–photoion coincidence spectroscopy study". European Journal of Mass Spectrometry. 15 (5): 157–66. doi:10.1255/ejms.943. PMID   19423901. S2CID   37022546.
  19. Eland, J. H. D. (1987). "The dynamics of three-body dissociations of dications studied by the triple coincidence technique PEPIPICO". Molecular Physics. 61 (3): 725–745. Bibcode:1987MolPh..61..725E. doi:10.1080/00268978700101421.
  20. Maier, J. P.; Thommen, F. (1980). "Fluorescence quantum yields and cascade-free lifetimes of state selected CO2+, COS+, CS2+ and N2O+ determined by photoelectron—photon coincidence spectroscopy". Chemical Physics. 51 (3): 319. Bibcode:1980CP.....51..319M. doi:10.1016/0301-0104(80)80106-6.
  21. Eland, J. H. D. (2003). "Complete double photoionisation spectra of small molecules from TOF-PEPECO measurements". Chemical Physics. 294 (2): 171–201. Bibcode:2003CP....294..171E. doi:10.1016/j.chemphys.2003.08.001.
  22. L J Frasinski, M Stankiewicz, K J Randall, P A Hatherly and K Codling "Dissociative photoionisation of molecules probed by triple coincidence; double time-of-flight techniques" J. Phys. B: At. Mol. Phys.19 L819–L824 (1986) open access
  23. Lebech, M.; Houver, J. C.; Dowek, D. (2002). "Ion–electron velocity vector correlations in dissociative photoionization of simple molecules using electrostatic lenses". Review of Scientific Instruments. 73 (4): 1866. Bibcode:2002RScI...73.1866L. doi:10.1063/1.1458063.
  24. Bodi, A.; Kvaran, Á. S.; Sztáray, B. L. (2011). "Thermochemistry of Halomethanes CFnBr4–n(n = 0–3) Based on iPEPICO Experiments and Quantum Chemical Computations". The Journal of Physical Chemistry A. 115 (46): 13443–13451. Bibcode:2011JPCA..11513443B. doi:10.1021/jp208018r. PMID   21985477.
  25. Baer, Tomas; Hase, William L. (1996). Unimolecular Reaction Dynamics: Theory and Experiments. Oxford University Press. ISBN   0-19-507494-7.
  26. Klots, C. E. (1973). "Thermochemical and kinetic information from metastable decompositions of ions". The Journal of Chemical Physics. 58 (12): 5364–5367. Bibcode:1973JChPh..58.5364K. doi:10.1063/1.1679153.
  27. Sztáray, B.; Bodi, A.; Baer, T. (2010). "Modeling unimolecular reactions in photoelectron photoion coincidence experiments". Journal of Mass Spectrometry. 45 (11): 1233–1245. Bibcode:2010JMSp...45.1233S. doi:10.1002/jms.1813. PMID   20872904.
  28. Heim, Pascal; Rumetshofer, Michael; Ranftl, Sascha; Thaler, Bernhard; Ernst, Wolfgang; Koch, Markus; Linden, Wolfgang (2019-01-19). "Bayesian Analysis of Femtosecond Pump-Probe Photoelectron-Photoion Coincidence Spectra with Fluctuating Laser Intensities". Entropy. 21 (1): 93. arXiv: 1901.06933 . Bibcode:2019Entrp..21...93H. doi: 10.3390/e21010093 . ISSN   1099-4300. PMC   7514205 . PMID   33266809.
  29. Weitzel, K. M.; Malow, M.; Jarvis, G. K.; Baer, T.; Song, Y.; Ng, C. Y. (1999). "High-resolution pulsed field ionization photoelectron–photoion coincidence study of CH4: Accurate 0 K dissociation threshold for CH3+". The Journal of Chemical Physics. 111 (18): 8267. Bibcode:1999JChPh.111.8267W. doi:10.1063/1.480169. S2CID   95900468.
  30. Bodi, A.; Shuman, N. S.; Baer, T. (2009). "On the ionization and dissociative photoionization of iodomethane: A definitive experimental enthalpy of formation of CH3I". Physical Chemistry Chemical Physics . Royal Society of Chemistry. 11 (46): 11013–11021. Bibcode:2009PCCP...1111013B. doi:10.1039/b915400k. PMID   19924337.
  31. Shuman, N. S.; Zhao, L. Y.; Boles, M.; Baer, T.; SztáRay, B. L. (2008). "Heats of Formation of HCCl3, HCCl2Br, HCClBr2, HCBr3, and Their Fragment Ions Studied by Threshold Photoelectron Photoion Coincidence". The Journal of Physical Chemistry A. 112 (42): 10533–10538. Bibcode:2008JPCA..11210533S. doi:10.1021/jp8056459. PMID   18823098.
  32. Bodi, A.; Kercher, J. P.; Bond, C.; Meteesatien, P.; Sztáray, B. L.; Baer, T. (2006). "Photoion Photoelectron Coincidence Spectroscopy of Primary Amines RCH2NH2 (R = H, CH3, C2H5, C3H7,i-C3H7): Alkylamine and Alkyl Radical Heats of Formation by Isodesmic Reaction Networks". The Journal of Physical Chemistry A. 110 (50): 13425–13433. Bibcode:2006JPCA..11013425B. doi:10.1021/jp064739s. PMID   17165868.