Jump to content
 







Main menu
   


Navigation  



Main page
Contents
Current events
Random article
About Wikipedia
Contact us
Donate
 




Contribute  



Help
Learn to edit
Community portal
Recent changes
Upload file
 








Search  

































Create account

Log in
 









Create account
 Log in
 




Pages for logged out editors learn more  



Contributions
Talk
 



















Contents

   



(Top)
 


1 Introduction  





2 Experiment  





3 Theory  





4 Results  





5 Links  





6 References  














Gas electron diffraction






Deutsch
Русский
 

Edit links
 









Article
Talk
 

















Read
Edit
View history
 








Tools
   


Actions  



Read
Edit
View history
 




General  



What links here
Related changes
Upload file
Special pages
Permanent link
Page information
Cite this page
Get shortened URL
Download QR code
Wikidata item
 




Print/export  



Download as PDF
Printable version
 
















Appearance
   

 






From Wikipedia, the free encyclopedia
 


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]

Introduction

[edit]

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 dimensionoflength.[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

Experiment

[edit]
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

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
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.
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

[edit]
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

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

[edit]
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:

[edit]

References

[edit]
  1. ^ a b 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)
  • ^ 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)
  • ^ 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) CS1 maint: numeric names: authors list (link)
  • ^ a b Bonham, R.A. (1974). High Energy Electron Scattering. Van Nostrand Reinhold.
  • ^ Hargittai, I. (1988). Stereochemical Applications of Gas‐Phase Electron Diffraction, Part A: The Electron Diffraction Technique. Weinheim: VCH Verlagsgesellschaft. ISBN 0-89573-337-4.
  • ^ 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.
  • ^ 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.
  • ^ 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.
  • ^ 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.
  • ^ 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.
  • ^ 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.
  • ^ 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.
  • ^ 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.
  • ^ 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.
  • ^ 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.

  • Retrieved from "https://en.wikipedia.org/w/index.php?title=Gas_electron_diffraction&oldid=1234646797"

    Category: 
    Diffraction
    Hidden categories: 
    CS1 maint: location missing publisher
    CS1 maint: others
    CS1 maint: multiple names: authors list
    CS1 maint: numeric names: authors list
    Articles with short description
    Short description matches Wikidata
    All articles with dead external links
    Articles with dead external links from December 2022
    Articles with permanently dead external links
     



    This page was last edited on 15 July 2024, at 12:15 (UTC).

    Text is available under the Creative Commons Attribution-ShareAlike License 4.0; additional terms may apply. By using this site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of the Wikimedia Foundation, Inc., a non-profit organization.



    Privacy policy

    About Wikipedia

    Disclaimers

    Contact Wikipedia

    Code of Conduct

    Developers

    Statistics

    Cookie statement

    Mobile view



    Wikimedia Foundation
    Powered by MediaWiki