Gas electron diffraction

Last updated

Gas electron diffraction (GED) is one of the applications of electron diffraction techniques. [1] The target of this method is the determination of the structure of gaseous molecules, i.e., the geometrical arrangement of the atoms from which a molecule is built up. GED is one of two experimental methods (besides microwave spectroscopy) to determine the structure of free molecules, undistorted by intermolecular forces, which are omnipresent in the solid and liquid state. The determination of accurate molecular structures [2] by GED studies is fundamental for an understanding of structural chemistry. [3] [1]

Contents

Introduction

Diffraction occurs because the wavelength of electrons accelerated by a potential of a few thousand volts is of the same order of magnitude as internuclear distances in molecules. The principle is the same as that of other electron diffraction methods such as LEED and RHEED, but the obtainable diffraction pattern is considerably weaker than those of LEED and RHEED because the density of the target is about one thousand times smaller. Since the orientation of the target molecules relative to the electron beams is random, the internuclear distance information obtained is one-dimensional. Thus only relatively simple molecules can be completely structurally characterized by electron diffraction in the gas phase. It is possible to combine information obtained from other sources, such as rotational spectra, NMR spectroscopy or high-quality quantum-mechanical calculations with electron diffraction data, if the latter are not sufficient to determine the molecule's structure completely.

The total scattering intensity in GED is given as a function of the momentum transfer, which is defined as the difference between the wave vector of the incident electron beam and that of the scattered electron beam and has the reciprocal dimension of length. [4] The total scattering intensity is composed of two parts: the atomic scattering intensity and the molecular scattering intensity. The former decreases monotonically and contains no information about the molecular structure. The latter has sinusoidal modulations as a result of the interference of the scattering spherical waves generated by the scattering from the atoms included in the target molecule. The interferences reflect the distributions of the atoms composing the molecules, so the molecular structure is determined from this part.

Figure 2: Diffraction pattern of gaseous benzene GED C6H6 diff pattern.jpg
Figure 2: Diffraction pattern of gaseous benzene

Experiment

Scheme 1: Schematic drawing of an electron diffraction apparatus GED scheme 1.jpg
Scheme 1: Schematic drawing of an electron diffraction apparatus
Scheme 2: Data reduction process from the concentric scattering pattern to the molecular scattering intensity curve Data reduction.jpg
Scheme 2: Data reduction process from the concentric scattering pattern to the molecular scattering intensity curve

Figure 1 shows a drawing and a photograph of an electron diffraction apparatus. Scheme 1 shows the schematic procedure of an electron diffraction experiment. A fast electron beam is generated in an electron gun, enters a diffraction chamber typically at a vacuum of 10−7 mbar. The electron beam hits a perpendicular stream of a gaseous sample effusing from a nozzle of a small diameter (typically 0.2 mm). At this point, the electrons are scattered. Most of the sample is immediately condensed and frozen onto the surface of a cold trap held at -196 °C (liquid nitrogen). The scattered electrons are detected on the surface of a suitable detector in a well-defined distance to the point of scattering.

Figure 1: Gas-diffraction apparatus at the University of Bielefeld, Germany GED Apparatus.jpg
Figure 1: Gas-diffraction apparatus at the University of Bielefeld, Germany
Figure 3: Scheme of a rotating sector, placement of the rotating sector within a GED apparatus and two examples of diffraction pattrens recorded with and without rotating sector. Rotating sector.jpg
Figure 3: Scheme of a rotating sector, placement of the rotating sector within a GED apparatus and two examples of diffraction pattrens recorded with and without rotating sector.

The scattering pattern consists of diffuse concentric rings (see Figure 2). The steep decent of intensity can be compensated for by passing the electrons through a fast rotation sector (Figure 3). This is cut in a way, that electrons with small scattering angles are more shadowed than those at wider scattering angles. The detector can be a photographic plate, an electron imaging plate (usual technique today) or other position sensitive devices such as hybrid pixel detectors (future technique).

The intensities generated from reading out the plates or processing intensity data from other detectors are then corrected for the sector effect. They are initially a function of distance between primary beam position and intensity, and then converted into a function of scattering angle. The so-called atomic intensity and the experimental background are subtracted to give the final experimental molecular scattering intensities as a function of s (the change of momentum).

These data are then processed by suitable fitting software like UNEX for refining a suitable model for the compound and to yield precise structural information in terms of bond lengths, angles and torsional angles.

Theory

Scheme 2: Schematic scattering process of an electron passing a positively charged atomic nucleus GED scattering.jpg
Scheme 2: Schematic scattering process of an electron passing a positively charged atomic nucleus
Firgure 4. Electron wave scattered at a pair of atomic nuclei at different distances Electron waves.jpg
Firgure 4. Electron wave scattered at a pair of atomic nuclei at different distances

GED can be described by scattering theory. The outcome if applied to gases with randomly oriented molecules is provided here in short: [5] [4]

Scattering occurs at each individual atom (), but also at pairs (also called molecular scattering) (), or triples (), of atoms.

is the scattering variable or change of electron momentum, and its absolute value is defined as

with being the electron wavelength defined above, and being the scattering angle.

The above-mentioned contributions of scattering add up to the total scattering

where is the experimental background intensity, which is needed to describe the experiment completely.

The contribution of individual atom scattering is called atomic scattering and easy to calculate:

with , being the distance between the point of scattering and the detector, being the intensity of the primary electron beam, and being the scattering amplitude of the i-th atom. In essence, this is a summation over the scattering contributions of all atoms independent of the molecular structure. is the main contribution and easily obtained if the atomic composition of the gas (sum formula) is known.

The most interesting contribution is the molecular scattering, because it contains information about the distance between all pairs of atoms in a molecule (bonded or non-bonded):

with being the parameter of main interest: the atomic distance between two atoms, being the mean square amplitude of vibration between the two atoms, the anharmonicity constant (correcting the vibration description for deviations from a purely harmonic model), and is a phase factor, which becomes important if a pair of atoms with very different nuclear charge is involved.

The first part is similar to the atomic scattering, but contains two scattering factors of the involved atoms. Summation is performed over all atom pairs.

is negligible in most cases and not described here in more detail. is mostly determined by fitting and subtracting smooth functions to account for the background contribution.

So it is the molecular scattering intensity that is of interest, and this is obtained by calculation all other contributions and subtracting them from the experimentally measured total scattering function.

Results

Figure 5: Examples of molecular intensity curves (left) and their Fourier transforms, the radial distribution curves of P4 and P3As. Examples P4 P3As.jpg
Figure 5: Examples of molecular intensity curves (left) and their Fourier transforms, the radial distribution curves of P4 and P3As.

Figure 5 shows two typical examples of results. The molecular scattering intensity curves are used to refine a structural model by means of a least squares fitting program. This yield precise structural information. The Fourier transformation of the molecular scattering intensity curves gives the radial distribution curves (RDC). These represent the probability to find a certain distance between two nuclei of a molecule. The curves below the RDC represent the diffrerence between the experiment and the model, i.e. the quality of fit.

The very simple example in Figure 5 shows the results for evaporated white phosphorus, P4. It is a perfectly tetrahedral molecule and has thus only one P-P distance. This makes the molecular scattering intensity curve a very simple one; a sine curve which is damped due to molecular vibration. The radial distribution curve (RDC) shows a maximum at 2.1994 Å with a least-squares error of 0.0003 Å, represented as 2.1994(3) Å. The width of the peak represents the molecular vibration and is the result of Fourier transformation of the damping part. This peak width means that the P-P distance varies by this vibration within a certain range given as a vibrational amplitude u, in this example uT(P‒P) = 0.0560(5) Å.

The slightly more complicated molecule P3As has two different distances P-P and P-As. Because their contributions overlap in the RDC, the peak is broader (also seen in a more rapid damping in the molecular scattering). The determination of these two independent parameters is more difficult and results in less precise parameter values than for P4.

Some selected other examples of important contributions to the structural chemistry of molecules are provided here:

Related Research Articles

<span class="mw-page-title-main">Diatomic molecule</span> Molecule composed of any two atoms

Diatomic molecules are molecules composed of only two atoms, of the same or different chemical elements. If a diatomic molecule consists of two atoms of the same element, such as hydrogen or oxygen, then it is said to be homonuclear. Otherwise, if a diatomic molecule consists of two different atoms, such as carbon monoxide or nitric oxide, the molecule is said to be heteronuclear. The bond in a homonuclear diatomic molecule is non-polar.

<span class="mw-page-title-main">Infrared spectroscopy</span> Measurement of infrared radiations interaction with matter

Infrared spectroscopy is the measurement of the interaction of infrared radiation with matter by absorption, emission, or reflection. It is used to study and identify chemical substances or functional groups in solid, liquid, or gaseous forms. It can be used to characterize new materials or identify and verify known and unknown samples. The method or technique of infrared spectroscopy is conducted with an instrument called an infrared spectrometer which produces an infrared spectrum. An IR spectrum can be visualized in a graph of infrared light absorbance on the vertical axis vs. frequency, wavenumber or wavelength on the horizontal axis. Typical units of wavenumber used in IR spectra are reciprocal centimeters, with the symbol cm−1. Units of IR wavelength are commonly given in micrometers, symbol μm, which are related to the wavenumber in a reciprocal way. A common laboratory instrument that uses this technique is a Fourier transform infrared (FTIR) spectrometer. Two-dimensional IR is also possible as discussed below.

<span class="mw-page-title-main">X-ray crystallography</span> Technique used for determining crystal structures and identifying mineral compounds

X-ray crystallography is the experimental science determining the atomic and molecular structure of a crystal, in which the crystalline structure causes a beam of incident X-rays to diffract into many specific directions. By measuring the angles and intensities of these diffracted beams, a crystallographer can produce a three-dimensional picture of the density of electrons within the crystal. From this electron density, the mean positions of the atoms in the crystal can be determined, as well as their chemical bonds, their crystallographic disorder, and various other information.

<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.

In physics, mean free path is the average distance over which a moving particle travels before substantially changing its direction or energy, typically as a result of one or more successive collisions with other particles.

<span class="mw-page-title-main">Electron diffraction</span> Bending of electron beams due to electrostatic interactions with matter

Electron diffraction refers to changes in the direction of electron beams due to interactions with atoms. Close to the atoms the changes are described as Fresnel diffraction; far away they are called Fraunhofer diffraction. The resulting map of the directions of the electrons far from the sample is called a diffraction pattern, see for instance Figure 1. These patterns are similar to x-ray and neutron diffraction patterns, and are used to study the atomic structure of gases, liquids, surfaces and bulk solids. Electron diffraction also plays a major role in the contrast of images in electron microscopes.

<span class="mw-page-title-main">Synchrotron light source</span> Particle accelerator designed to produce intense x-ray beams

A synchrotron light source is a source of electromagnetic radiation (EM) usually produced by a storage ring, for scientific and technical purposes. First observed in synchrotrons, synchrotron light is now produced by storage rings and other specialized particle accelerators, typically accelerating electrons. Once the high-energy electron beam has been generated, it is directed into auxiliary components such as bending magnets and insertion devices in storage rings and free electron lasers. These supply the strong magnetic fields perpendicular to the beam that are needed to convert high energy electrons into photons.

<span class="mw-page-title-main">Chemical structure</span> Organized way in which molecules are ordered and sorted

A chemical structure of a molecule is a spatial arrangement of its atoms and their chemical bonds. Its determination includes a chemist's specifying the molecular geometry and, when feasible and necessary, the electronic structure of the target molecule or other solid. Molecular geometry refers to the spatial arrangement of atoms in a molecule and the chemical bonds that hold the atoms together and can be represented using structural formulae and by molecular models; complete electronic structure descriptions include specifying the occupation of a molecule's molecular orbitals. Structure determination can be applied to a range of targets from very simple molecules to very complex ones.

<span class="mw-page-title-main">Raman scattering</span> Inelastic scattering of photons

Raman scattering or the Raman effect is the inelastic scattering of photons by matter, meaning that there is both an exchange of energy and a change in the light's direction. Typically this effect involves vibrational energy being gained by a molecule as incident photons from a visible laser are shifted to lower energy. This is called normal Stokes Raman scattering. The effect is exploited by chemists and physicists to gain information about materials for a variety of purposes by performing various forms of Raman spectroscopy. Many other variants of Raman spectroscopy allow rotational energy to be examined and electronic energy levels may be examined if an X-ray source is used in addition to other possibilities. More complex techniques involving pulsed lasers, multiple laser beams and so on are known.

<span class="mw-page-title-main">Molecular physics</span> Study of the physical and chemical properties of molecules

Molecular physics is the study of the physical properties of molecules and molecular dynamics. The field overlaps significantly with physical chemistry, chemical physics, and quantum chemistry. It is often considered as a sub-field of atomic, molecular, and optical physics. Research groups studying molecular physics are typically designated as one of these other fields. Molecular physics addresses phenomena due to both molecular structure and individual atomic processes within molecules. Like atomic physics, it relies on a combination of classical and quantum mechanics to describe interactions between electromagnetic radiation and matter. Experiments in the field often rely heavily on techniques borrowed from atomic physics, such as spectroscopy and scattering.

<span class="mw-page-title-main">Molecular geometry</span> Study of the 3D shapes of molecules

Molecular geometry is the three-dimensional arrangement of the atoms that constitute a molecule. It includes the general shape of the molecule as well as bond lengths, bond angles, torsional angles and any other geometrical parameters that determine the position of each atom.

In physics, the phase problem is the problem of loss of information concerning the phase that can occur when making a physical measurement. The name comes from the field of X-ray crystallography, where the phase problem has to be solved for the determination of a structure from diffraction data. The phase problem is also met in the fields of imaging and signal processing. Various approaches of phase retrieval have been developed over the years.

<span class="mw-page-title-main">Powder diffraction</span>

Powder diffraction is a scientific technique using X-ray, neutron, or electron diffraction on powder or microcrystalline samples for structural characterization of materials. An instrument dedicated to performing such powder measurements is called a powder diffractometer.

<span class="mw-page-title-main">Low-energy electron diffraction</span> Technique for the determination of the surface structure of single-crystalline materials

Low-energy electron diffraction (LEED) is a technique for the determination of the surface structure of single-crystalline materials by bombardment with a collimated beam of low-energy electrons (30–200 eV) and observation of diffracted electrons as spots on a fluorescent screen.

In condensed matter physics and crystallography, the static structure factor is a mathematical description of how a material scatters incident radiation. The structure factor is a critical tool in the interpretation of scattering patterns obtained in X-ray, electron and neutron diffraction experiments.

Helium atom scattering (HAS) is a surface analysis technique used in materials science. It provides information about the surface structure and lattice dynamics of a material by measuring the diffracted atoms from a monochromatic helium beam incident on the sample.

<span class="mw-page-title-main">Low-energy ion scattering</span>

Low-energy ion scattering spectroscopy (LEIS), sometimes referred to simply as ion scattering spectroscopy (ISS), is a surface-sensitive analytical technique used to characterize the chemical and structural makeup of materials. LEIS involves directing a stream of charged particles known as ions at a surface and making observations of the positions, velocities, and energies of the ions that have interacted with the surface. Data that is thus collected can be used to deduce information about the material such as the relative positions of atoms in a surface lattice and the elemental identity of those atoms. LEIS is closely related to both medium-energy ion scattering (MEIS) and high-energy ion scattering, differing primarily in the energy range of the ion beam used to probe the surface. While much of the information collected using LEIS can be obtained using other surface science techniques, LEIS is unique in its sensitivity to both structure and composition of surfaces. Additionally, LEIS is one of a very few surface-sensitive techniques capable of directly observing hydrogen atoms, an aspect that may make it an increasingly more important technique as the hydrogen economy is being explored.

In X-ray absorption spectroscopy, the K-edge is a sudden increase in x-ray absorption occurring when the energy of the X-rays is just above the binding energy of the innermost electron shell of the atoms interacting with the photons. The term is based on X-ray notation, where the innermost electron shell is known as the K-shell. Physically, this sudden increase in attenuation is caused by the photoelectric absorption of the photons. For this interaction to occur, the photons must have more energy than the binding energy of the K-shell electrons (K-edge). A photon having an energy just above the binding energy of the electron is therefore more likely to be absorbed than a photon having an energy just below this binding energy or significantly above it.

Structural chemistry is a part of chemistry and deals with spatial structures of molecules and solids.

In crystallography, direct methods is a set of techniques used for structure determination using diffraction data and a priori information. It is a solution to the crystallographic phase problem, where phase information is lost during a diffraction measurement. Direct methods provides a method of estimating the phase information by establishing statistical relationships between the recorded amplitude information and phases of strong reflections.

References

  1. 1 2 Rankin, David W. H. (2 January 2013). Structural methods in molecular inorganic chemistry. Morrison, Carole A., 1972-, Mitzel, Norbert W., 1966-. Chichester, West Sussex, United Kingdom. ISBN   978-1-118-46288-1. OCLC   810442747.{{cite book}}: CS1 maint: location missing publisher (link)
  2. Accurate molecular structures : their determination and importance. Domenicano, Aldo., Hargittai, István. [Chester, England]: International Union of Crystallography. 1992. ISBN   0-19-855556-3. OCLC   26264763.{{cite book}}: CS1 maint: others (link)
  3. Wells, A. F. (Alexander Frank), 1912- (12 July 2012). Structural inorganic chemistry (Fifth ed.). Oxford. ISBN   978-0-19-965763-6. OCLC   801026482.{{cite book}}: CS1 maint: location missing publisher (link) CS1 maint: multiple names: authors list (link)
  4. 1 2 Bonham, R.A. (1974). High Energy Electron Scattering. Van Nostrand Reinhold.
  5. Hargittai, I. (1988). Stereochemical Applications of Gas‐Phase Electron Diffraction, Part A: The Electron Diffraction Technique. Weinheim: VCH Verlagsgesellschaft. ISBN   0-89573-337-4.
  6. Hedberg, Kenneth; Schomaker, Verner (April 1951). "A Reinvestigation of the Structures of Diborane and Ethane by Electron Diffraction 1,2". Journal of the American Chemical Society. 73 (4): 1482–1487. doi:10.1021/ja01148a022. ISSN   0002-7863.
  7. Hedberg, Kenneth (1955-12-01). "The Molecular Structure of Trisilylamine (SiH3)3N1,2". Journal of the American Chemical Society. 77 (24): 6491–6492. doi:10.1021/ja01629a015. ISSN   0002-7863.
  8. Cossairt, Brandi M.; Cummins, Christopher C.; Head, Ashley R.; Lichtenberger, Dennis L.; Berger, Raphael J. F.; Hayes, Stuart A.; Mitzel, Norbert W.; Wu, Gang (2010-06-23). "On the Molecular and Electronic Structures of AsP3 and P4". Journal of the American Chemical Society. 132 (24): 8459–8465. doi:10.1021/ja102580d. ISSN   0002-7863. PMID   20515032.
  9. Hedberg, K.; Hedberg, L.; Bethune, D. S.; Brown, C. A.; Dorn, H. C.; Johnson, R. D.; De Vries, M. (1991-10-18). "Bond Lengths in Free Molecules of Buckminsterfullerene, C60, from Gas-Phase Electron Diffraction". Science. 254 (5030): 410–412. Bibcode:1991Sci...254..410H. doi:10.1126/science.254.5030.410. ISSN   0036-8075. PMID   17742230. S2CID   25860557.
  10. Hedberg, Kenneth; Hedberg, Lise; Bühl, Michael; Bethune, Donald S.; Brown, C. A.; Johnson, Robert D. (1997-06-01). "Molecular Structure of Free Molecules of the Fullerene C70 from Gas-Phase Electron Diffraction". Journal of the American Chemical Society. 119 (23): 5314–5320. doi:10.1021/ja970110e. ISSN   0002-7863.
  11. Vishnevskiy, Yury V.; Tikhonov, Denis S.; Schwabedissen, Jan; Stammler, Hans-Georg; Moll, Richard; Krumm, Burkhard; Klapötke, Thomas M.; Mitzel, Norbert W. (2017-08-01). "Tetranitromethane: A Nightmare of Molecular Flexibility in the Gaseous and Solid States". Angewandte Chemie International Edition. 56 (32): 9619–9623. doi:10.1002/anie.201704396. PMID   28557111.
  12. Mitzel, Norbert W.; Brown, Daniel H.; Parsons, Simon; Brain, Paul T.; Pulham, Colin R.; Rankin, David W. H. (1998). "Differences Between Gas-Phase and Solid-State Molecular Structures of the Simplest Phosphonium Ylide, Me3P=CH2". Angewandte Chemie International Edition. 37 (12): 1670–1672. doi:10.1002/(SICI)1521-3773(19980703)37:12<1670::AID-ANIE1670>3.0.CO;2-S. ISSN   1521-3773. PMID   29711513.
  13. Mitzel, Norbert W.; Smart, Bruce A.; Dreihäupl, Karl-Heinz; Rankin, David W. H.; Schmidbaur, Hubert (January 1996). "Low Symmetry in P(NR 2 ) 3 Skeletons and Related Fragments: An Inherent Phenomenon". Journal of the American Chemical Society. 118 (50): 12673–12682. doi:10.1021/ja9621861. ISSN   0002-7863.
  14. Fokin, Andrey A.; Zhuk, Tatyana S.; Blomeyer, Sebastian; Pérez, Cristóbal; Chernish, Lesya V.; Pashenko, Alexander E.; Antony, Jens; Vishnevskiy, Yury V.; Berger, Raphael J. F.; Grimme, Stefan; Logemann, Christian (2017-11-22). "Intramolecular London Dispersion Interaction Effects on Gas-Phase and Solid-State Structures of Diamondoid Dimers". Journal of the American Chemical Society. 139 (46): 16696–16707. doi:10.1021/jacs.7b07884. ISSN   0002-7863. PMID   29037036.
  15. Kveseth, Kari (August 2019). "The story of gas-phase electron diffraction (GED) in Norway". Structural Chemistry. 30 (4): 1505–1516. doi:10.1007/s11224-019-01309-w. ISSN   1040-0400. S2CID   146084935.