International
Tables for
Crystallography
Volume C
Mathematical, physical and chemical tables
Edited by E. Prince

International Tables for Crystallography (2006). Vol. C. ch. 4.3, pp. 416-419

Section 4.3.7. Measurement of structure factors and determination of crystal thickness by electron diffraction

J. Gjønnese and J. W. Steedsm

4.3.7. Measurement of structure factors and determination of crystal thickness by electron diffraction

| top | pdf |

Current advances in quantitative electron diffraction are connected with improved experimental facilities, notably the combination of convergent-beam electron diffraction (CBED) with new detection systems. This is reflected in extended applications of electron diffraction intensities to problems in crystallography, ranging from valence-electron distributions in crystals with small unit cells to structure determination of biological molecules in membranes. The experimental procedures can be seen in relation to the two main principles for measurement of diffracted intensities from crystals:

  • rocking curves, i.e. intensity profiles measured as function of deviation, sg, from the Bragg condition, and

  • integrated intensities, which form the well known basis for X-ray and neutron diffraction determination of crystal structure.

Integrated intensities are not easily defined in the most common type of electron-diffraction pattern, viz the selected-area (SAD) spot pattern. This is due to the combination of dynamical scattering and the orientation and thickness variations usually present within the typically micrometre-size illuminated area. This combination leads to spot pattern intensities that are poorly defined averages over complicated scattering functions of many structure factors. Convergent-beam electron diffraction is a better alternative for intensity measurements, especially for inorganic structures with small-to-moderate unit cells. In CBED, a fine beam is focused within an area of a few hundred ångströms, with a divergence of the order of a tenth of a degree. The diffraction pattern then appears in the form of discs, which are essentially two-dimensional rocking curves from a small illuminated area, within which thickness and orientation can be regarded as constant. These intensity distributions are obtained under well defined conditions and are well suited for comparison with theoretical calculations. The intensity can be recorded either photographically, or with other parallel recording systems, viz YAG screen/CCD camera (Krivanek, Mooney, Fan, Leber & Meyer, 1991[link]) or image plates (Mori, Oikawa & Harada, 1990[link]) – or sequentially by a scanning system. The inelastic background can be removed by an energy filter (Krahl, Pätzold & Swoboda, 1990[link]; Krivanek, Gubbens, Dellby & Meyer, 1991[link]). Detailed intensity profiles in one or two dimensions can then be measured with high precision for low-order reflections from simple structures. But there are limitations also with the CBED technique: the crystal should be fairly perfect within the illuminated area and the unit cell relatively small, so that overlap between discs can be avoided. The current development of electron diffraction is therefore characterized by a wide range of techniques, which extend from the traditional spot pattern to two-dimensional, filtered rocking curves, adapted to the structure problems under study and the specimens that are available.

Spot-pattern intensities are best for thin samples of crystals with light atoms, especially organic and biological materials. Dorset and co-workers (Dorset, Jap, Ho & Glaeser, 1979[link]; Dorset, 1991[link]) have shown how conventional crystallographic techniques (`direct phasing') can be applied in ab initio structure determination of thin organic crystals from spot intensities in projections. Two main complications were treated by them: bending of the crystal and dynamical scattering. Thin crystals will frequently be bent; this will give some integration of the reflection, but may also produce a slight distortion of the structure, as pointed out by Cowley (1961[link]), who proposed a correction formula. The thickness range for which a kinematical approach to intensities is valid was estimated theoretically by Dorset et al. (1979[link]). For organic crystals, they quoted a few hundred ångströms as a limit for kinematical scattering in dense projections at 100 kV.

Radiation damage is a problem, but with low-dose and cryo-techniques, electron-microscopy methods can be applied to many organic crystals, as shown by several recent investigations. Voigt-Martin, Yan, Gilmore, Shankland & Bricogne (1994[link]) collected electron-diffraction intensities from a beam-sensitive dione and constructed a 1.4 Å Fourier map by a direct method based on maximum entropy. Large numbers of electron-diffraction intensities have been collected from biological molecules crystallized in membranes. The structure amplitudes can be combined with phases extracted from high-resolution micrographs, following Henderson Unwin's (1975[link]) early work. Kühlbrandt, Wang & Fujiyoshi (1994[link]) collected about 18 000 amplitudes and 15 000 phases for a protein complex in an electron cryomicroscope operating at 4.2 K (Fujiyoshi et al., 1991[link]). Using these data, they determined the structure from a three-dimensional Fourier map calculated to 3.4 Å resolution. The assumption of kinematical scattering in such studies has been investigated by Spargo (1994[link]), who found the amplitudes to be kinematic within 4% but with somewhat larger deviations for phases.

For inorganic structures, spot-pattern intensities are less useful because of the stronger dynamical interactions, especially in dense zones. Nevertheless, it may be possible to derive a structure and refine parameters from spot-pattern intensities. Andersson (1975[link]) used experimental intensities from selected projections for comparison with dynamical calculations, including an empirical correction factor for orientation spread, in a structure determination of V14O8. Recently, Zou, Sukharev & Hovmöller (1993[link]) combined spot-pattern intensities read from film by the program ELD with image processing of high-resolution micrographs for structure determination of a complex perovskite.

A considerable improvement over the spot pattern has been obtained by the elegant double-precession technique devised by Vincent & Midgley (1994[link]). They programmed scanning coils above and below the specimen in the electron microscope so as to achieve simultaneous precession of the focused incident beam and the diffraction pattern around the optical axis. The net effect is equivalent to a precession of the specimen with a stationary incident beam. Integrated intensities can be obtained from reflections out to a Bragg angle [\theta] equal to the precession angle [\varphi] for the zeroth Laue zone. In addition, reflections in the first and second Laue zones appear as broad concentric rings. Dynamical effects are reduced appreciably by this procedure, especially in the non-zero Laue zones. The experimental integrated intensities, Ig, must be multiplied with a geometrical factor analogous to the Lorentz factor in X-ray diffraction, viz [I_g=I_G^{\rm exp}\sin \varepsilon; \quad \cos \varepsilon ={{(g^2 -2nkh)}\over {2k\phi g}}, \eqno (4.3.7.1)]where nh is the reciprocal spacing between the zeroth and nth layers. The intensities can be used for structure determination by procedures taken over from X-ray crystallography, e.g. the conditional Patterson projections that are used by the Bristol group (Vincent, Bird & Steeds, 1984[link]). The precession method may be seen as intermediate between the spot pattern and the CBED technique. Another intermediate approach was proposed by Goodman (1976[link]) and used later by Olsen, Goodman & Whitfield (1985[link]) in the structure determination of a series of selenides. CBED patterns from thin crystals were taken in dense zones; intensities were measured at corresponding points in the discs, e.g. at the zone-axis position. Structure parameters were determined by fitting the observed intensities to dynamical calculations.

Higher precision and more direct comparisons with dynamical scattering calculations are achieved by measurements of intensity distributions within the CBED discs, i.e. one- or two-dimensional rocking curves. An up-to-date review of these techniques is found in the recent book by Spence & Zuo (1992[link]), where all aspects of the CBED technique, theory and applications are covered, including determination of lattice constants and strains, crystal symmetry, and fault vectors of defects. Refinement of structure factors in crystals with small unit cells are treated in detail. For determination of bond charges, the structure factors (Fourier potentials) should be determined to an accuracy of a few tenths of a percent; calculations must then be based on many-beam dynamical scattering theory, see Chapter 8.8[link] . Removal of the inelastic background by an energy filter will improve the data considerably; analytical expressions for the inelastic background including multiple-scattering contributions may be an alternative (Marthinsen, Holmestad & Høier, 1994[link]).

Early CBED applications to the determination of structure factors were based on features that can be related to dynamical effects in the two-beam case. Although insufficient for most accurate analyses, the two-beam expression for the intensity profile may be a useful guide. In its standard form, [I_g(s)={{(U_g/k)^2}\over {s_g^2+(U_g/k)^2}}\;\sin ^2\left [\pi t\sqrt {s_g^2+(U_g/k)^2}\right] ,\eqno (4.3.7.2)]where Ug and sg are Fourier potential and excitation error for the reflection g, k wave number and t thickness. The expression can be rewritten in terms of the eigenvalues γ(i, j) that correspond to the two Bloch-wave branches, i, j: [I_g^{i,\,j}(s_g)={{(U_g/k)^2}\over {(\gamma ^{(i)}-\gamma ^{(\,j)})}^2}\,\sin ^2[\pi t(\gamma ^{(i)}-\gamma ^{(\,j)})],\eqno (4.3.7.3)]where [\gamma ^{i,\,j}=\textstyle {1\over 2}\left [s_g^2\pm \sqrt {s_g^2+(U_g/k)^2}\;\right] .]Note that the minimum separation between the branches i, j or the gap at the dispersion surface is [(\gamma ^{(\,j)}-\gamma ^{(i)})_{\rm min}=U_g/k=1/\xi _g,\eqno (4.3.7.4)]where [\xi]g is an extinction distance. The two-beam form is often found to be a good approximation to an intensity profile Ig(sg) even when other beams are excited, provided an effective potential [U_g^{\rm eff}], which corresponds to the gap at the dispersion surface, is substituted for Ug. This is suggested by many features in CBED and Kikuchi patterns and borne out by detailed calculations, see e.g. Høier (1972[link]). Approximate expressions for [U_g^{\rm eff}] have been developed along different lines; the best known is the Bethe potential [U_g^{\rm eff} =U_g - \sum _h {{U_{g-h}U_h}\over {2ks_h}}.\eqno (4.3.7.5)]Other perturbation approaches are based on scattering between Bloch waves, in analogy with the `interband scattering' introduced by Howie (1963[link]) for diffuse scattering; the term `Bloch-wave hybridization' was introduced by Buxton (1976[link]). Exact treatment of symmetrical few-beam cases is possible (see Fukuhara, 1966[link]; Kogiso & Takahashi, 1977[link]). The three-beam case (Kambe, 1957[link]; Gjønnes & Høier, 1971[link]) is described in detail in the book by Spence & Zuo (1992[link]).

Many intensity features can be related to the structure of the dispersion surface, as represented by the function γ(kx, ky). The gap [equation (4.3.7.4)[link]] is an important parameter, as in the four-beam symmetrical case in Fig. 4.3.7.1[link] . Intensity measurements along one dimension can then be referred to three groups, according to the width of the gap, viz:

  • small gap – integrated intensity;

    [Figure 4.3.7.1]

    Figure 4.3.7.1| top | pdf |

    (a) Dispersion-surface section for the symmetric four-beam case (0, g, g + h, m), γk is a function of kx, referred to (b), where kx = ky = 0 corresponds to the exact Bragg condition for all three reflections. The two gaps appear at sg = ±(UhUm)/k with widths (Ug ± Ug + h)/k.

  • large gap – rocking curve, thickness fringes;

  • zero gap – critical effects.

A small gap at the dispersion surface implies that the two-beam-like rocking curve above approaches a kinematical form and can be represented by an integrated intensity. Within a certain thickness range, this intensity may be proportional to [|U^{\rm eff}_g|^2], with an angular width inversely proportional to gt. Several schemes have been proposed for measurement of relative integrated intensities for reflections in the outer, high-angle region, where the lines are narrow and can be easily separated from the background. Steeds (1984[link]) proposed use of the HOLZ (high-order Laue-zone) lines, which appear in CBED patterns taken with the central disc at the zone-axis position. Along a ring that defines the first-order Laue zone (FOLZ), reflections appear as segments that can be associated with scattering from strongly excited Bloch waves in the central ZOLZ part into the FOLZ reflections. Vincent, Bird & Steeds (1984[link]) proposed an intensity expression [I_g^{(\,j)}\propto |\varepsilon ^{(\,j)}\beta _g^{(\,j)}|^2\exp (-2\mu t){{1 - \exp [-2(\mu ^{(\,j)}-\mu)t]}\over {2(\mu ^{(\,j)}-\mu)}}\eqno (4.3.7.6)]for integrated intensity for a line segment associated with scattering from (or into) the ZOLZ Bloch wave j. [\varepsilon^{(\,j)}] is here the excitation coefficient and β(j) the matrix element for scattering between the Bloch wave j and the plane wave g. μ(j) and μ are absorption coefficients for the Bloch wave and plane wave, respectively; t is the thickness. From measurements of a number of such FOLZ (or SOLZ) reflections, they were able to carry out ab initio structure determinations using so-called conditional Patterson projections and coordinate refinement. Tanaka & Tsuda (1990[link]) have refined atomic positions from zone-axis HOLZ intensities. Ratios between HOLZ intensities have been used for determination of the Debye–Waller factor (Holmestad, Weickenmeier, Zuo, Spence & Horita, 1993[link]).

Another CBED approach to integrated intensities is due to Taftø & Metzger (1985[link]). They measured a set of high-order reflections along a systematic row with a wide-aperture CBED tilted off symmetrical incidence. A number of high-order reflections are then simultaneously excited in a range where the reflections are narrow and do not overlap. Gjønnes & Bøe (1994[link]) and Ma, Rømming, Lebech & Gjønnes (1992[link]) applied the technique to the refinement of coordinates and thermal parameters in high-Tc superconductors and intermetallic compounds. The validity and limitation of the kinematical approximation and dynamic potentials in this case has been discussed by Gjønnes & Bøe (1994[link]).

Zero gap at the dispersion surface corresponds to zero effective Fourier potential or, to be more exact, an accidental degeneracy, γ(i) = γ(j), in the Bloch-wave solution. This is the basis for the critical-voltage method first shown by Watanabe, Uyeda & Fukuhara (1969[link]). From vanishing contrast of the Kikuchi line corresponding to a second-order reflection 2g, they determined a relation between the structure factors Ug and U2g. Gjønnes & Høier (1971[link]) derived the condition for the accidental degeneracy in the general centrosymmetrical three-beam case 0,g,h, expressed in terms of the excitation errors sg,h and Fourier potentials Ug,h,g−h, viz [2ks_g={{U_g(U_h^2 - U_{g-h}^2)m}\over {U_hU_{g-h}m_0}}; \quad 2ks_h={{U_h(U_g^2 - U_{g-h}^2)m}\over {U_gU_{g-h}m_0}};\eqno (4.3.7.7)]where m and m0 are the relativistic and rest mass of the incident electron. Experimentally, this condition is obtained at a particular voltage and diffraction condition as vanishing line contrast of a Kikuchi or Kossel line – or as a reversal of a contrast feature. The second-order critical-voltage effect is then obtained as a special case, e.g. by the mass ratio: [(m/m_0)_{\rm crit}={{U_{2h}h^2}\over {U_h^2 - U_{2h}^2}}. \eqno (4.3.7.8)]Measurements have been carried out for a number of elements and alloy phases; see the review by Fox & Fisher (1988[link]) and later work on alloys by Fox & Tabbernor (1991[link]). Zone-axis critical voltages have been used by Matsuhata & Steeds (1987[link]). For analytical expressions and experimental determination of non-systematic critical voltages, see Matsuhata & Gjønnes (1994[link]).

Large gaps at the dispersion surface are associated with strong inner reflections – and a strong dynamical effect of two-beam-like character. The absolute magnitude of the gap – or its inverse, the extinction distance – can be obtained in different ways. Early measurements were based on the split of diffraction spots from a wedge, see Lehmpfuhl (1974[link]), or the corresponding fringe periods measured in bright- and dark-field micrographs (Ando, Ichimiya & Uyeda, 1974[link]). The most precise and applicable large-gap methods are based on the refinement of the fringe pattern in CBED discs from strong reflections, as developed by Goodman & Lehmpfuhl (1967[link]) and Voss, Lehmpfuhl & Smith (1980[link]). In recent years, this technique has been developed to high perfection by means of filtered CBED patterns, see Spence & Zuo (1992[link]) and papers referred to therein. See also Chapter 8.8[link] .

The gap at the dispersion surface can also be obtained directly from the split observed at the crossing of a weak Kikuchi line with a strong band. Gjønnes & Høier (1971[link]) showed how this can be used to determine strong low-order reflections. High voltage may improve the accuracy (Terasaki, Watanabe & Gjønnes, 1979[link]). The sensitivity of the intersecting Kikuchi-line (IKL) method was further increased by the use of CBED instead of Kikuchi patterns (Matsuhata, Tomokiyo, Watanabe & Eguchi, 1984[link]; Taftø & Gjønnes, 1985[link]). In a recent development, Høier, Bakken, Marthinsen & Holmestad (1993[link]) have measured the intensity distribution in the CBED discs around such intersections and have refined the main structure factors involved.

Two-dimensional rocking curves collected by CBED patterns around the axis of a dense zone are complicated by extensive many-beam dynamical interactions. The Bristol–Bath group (Saunders, Bird, Midgley & Vincent, 1994[link]) claim that the strong dynamic effects can be exploited to yield high sensitivity in refinement of low-order structure factors. They have also developed procedures for ab initio structure determination based on zone-axis patterns (Bird & Saunders, 1992[link]), see Chapter 8.8[link] .

Determination of phase invariants. It has been known for some time (e.g. Kambe, 1957[link]) that the dynamical three-beam case contains information about phase. As in the X-ray case, measurement of dynamical effects can be used to determine the value of triplets (Zuo, Høier & Spence, 1989[link]) and to determine phase angles to better than one tenth of a degree (Zuo, Spence, Downs & Mayer, 1993[link]) which is far better than any X-ray method. Bird (1990[link]) has pointed out that the phase of the absorption potential may differ from the phase of the real potential.

Thickness is an important parameter in electron-diffraction experiments. In structure-factor determination based on CBED patterns, thickness is often included in the refinement. Thickness can also be determined directly from profiles connected with large gaps at the dispersion surface (Goodman & Lehmpfuhl, 1967[link]; Blake, Jostsons, Kelly & Napier, 1978[link]; Glazer, Ramesh, Hilton, & Sarikaya, 1985[link]). The method is based on the outer part of the fringe profile, which is not so sensitive to the structure factor. The intensity minimum of the ith fringe in the diffracted disc occurs at a position corresponding to the excitation error si and expressed as [(s_i^2+1/\varepsilon _g^2)t^2=n_i^2,\eqno (4.3.7.9)]where ni is a small integer describing the order of the minimum. This equation can be arranged in two ways for graphic determination of thickness. The commonest method appears to be to plot (si/ni)2 against 1/ni2and then determine the thickness from the intersection with the ordinate axis (Kelly, Jostsons, Blake & Napier, 1975[link]). Glazer et al. (1985[link]) claim that the method originally proposed by Ackermann (1948[link]), where [s_i^2] is plotted against ni and the thickness is taken from the slope, is more accurate. In both cases, the outer part of the rocking curve is emphasized; exact knowledge of the gap is not necessary for a good determination of thickness, provided the assumption of a two-beam-like rocking curve is valid.

References

First citation Ackermann, I. (1948). Observations on the dynamical interference phenomena in convergent electron beams. II. Ann. Phys. (Leipzig), 2, 41–54.Google Scholar
First citation Andersson, B. (1975). Structure analysis of the γ-phase in the vanadium oxide system by electron diffraction studies. Acta Cryst. A31, 63–70.Google Scholar
First citation Ando, Y., Ichimiya, A. & Uyeda, R. (1974). A determination of values and signs of the 111 and 222 structure factors of silicon. Acta Cryst. A30, 600–601.Google Scholar
First citation Bird, D. M. (1990). Absorption in high-energy electron diffraction from non-centrosymmetrical crystals. Acta Cryst. A46, 208–214.Google Scholar
First citation Bird, D. M. & Saunders, M. (1992). Inversion of convergent-beam electron diffraction patterns. Acta Cryst. A48, 555–562.Google Scholar
First citation Blake, R. G., Jostsons, A., Kelly, P. M. & Napier, J. G. (1978). The determination of extinction distances and anomalous absorption coefficients by scanning electron microscopy. Philos. Mag. A37, 1–16.Google Scholar
First citation Buxton, B. F. (1976). Bloch waves in high order Laue zone effects in high energy electron diffraction. Proc. R. Soc. London Ser. A, 300, 335–361.Google Scholar
First citation Cowley, J. M. (1961). Diffraction intensities from bent crystals. Acta Cryst. 14, 920–926.Google Scholar
First citation Dorset, D. L. (1991). Is electron crystallography possible? The direct determination of organic crystal structures. Ultramicroscopy, 38, 23–40.Google Scholar
First citation Dorset, D. L., Jap, B. K., Ho, M. M. & Glaeser, R. M. (1979). Direct phasing of electron diffraction data from organic crystals: the effect of n-beam dynamical scattering. Acta Cryst. A35, 1001–1009.Google Scholar
First citation Fox, A. G. & Fisher, R. M. (1988). A summary of low-angle X-ray atomic scattering factors measured by the critical voltage effect in high energy electron diffraction. Aust. J. Phys. 41, 461–468.Google Scholar
First citation Fox, A. G. & Tabbernor, M. A. (1991). The bonding charge density of [\beta]′NiAl′′. Acta Metall. 39, 669–678.Google Scholar
First citation Fujiyoshi, Y., Mizusaki, T., Morikawa, K., Yamagishi, H., Aoki, Y., Kihara, H. & Harada, Y. (1991). Development of a superfluid helium stage for high-resolution electron microscopy. Ultramicroscopy, 38, 241–251.Google Scholar
First citation Fukuhara, A. (1966). Many-ray approximation in the dynamical theory of electron diffraction. J. Phys. Soc. Jpn, 21, 2645–2662.Google Scholar
First citation Gjønnes, J. & Høier, R. (1971). The application of non-systematic many-beam dynamic effects to structure-factor determination. Acta Cryst. A27, 313–316.Google Scholar
First citation Gjønnes, K. & Bøe, N. (1994). Refinement of temperature factors and charge distributions in YBa2Cu3O7 and YBa2(Cu,Co)3O7 from CBED intensities. Micron Microsc. Acta, 25, 29–44.Google Scholar
First citation Glazer, J., Ramesh, R., Hilton, M. R. & Sarikaya, M. (1985). Comparison of convergent beam electron diffraction methods for determination of foil thickness. Philos. Mag. A52, 59–63.Google Scholar
First citation Goodman, P. (1976). Examination of the graphite structure using convergent-beam electron diffraction. Acta Cryst. A32, 793–798.Google Scholar
First citation Goodman, P. & Lehmpfuhl, G. (1967). Electron diffraction study of MgO h00 systematic interactions. Acta Cryst. 22, 14–24.Google Scholar
First citation Henderson, R. & Unwin, P. N. T. (1975). Three-dimensional model of purple membrane obtained by electron microscopy. Nature, 257, 28–32.Google Scholar
First citation Høier, R. (1972). Displaced lines in Kikuchi patterns. Phys. Status Solidi A, 11, 597–610.Google Scholar
First citation Høier, R., Bakken, L. N., Marthinsen, K. & Holmestad, R. (1993). Structure factor determination in non-centrosymmetrical crystals by a two-dimensional CBED-based multi-parameter refinement method. Ultramicroscopy, 49, 159–170.Google Scholar
First citation Holmestad, R., Weickenmeier, A. L., Zuo, J. M., Spence, J. C. H. & Horita, Z. (1993). Debye–Waller factor measurement in TiAl from HOLZ reflections. Electron microscopy and analysis 1993, pp. 141–144. Bristol: IOP Publishing.Google Scholar
First citation Howie, A. (1963). Inelastic scattering of electrons by crystals. I. The theory of small angle inelastic scattering. Proc. R. Soc. London Ser. A, 271, 268–287.Google Scholar
First citation Kambe, K. (1957). Study of simultaneous reflections in electron diffraction by crystals. J. Phys. Soc. Jpn, 12, 13–36.Google Scholar
First citation Kelly, P. M., Jostsons, A., Blake, R. G. & Napier, J. G. (1975). The determination of foil thickness by scanning transmission electron microscopy. Phys. Status Solidi A, 31, 771–780.Google Scholar
First citation Kogiso, M. & Takahashi, H. (1977). Group-theoretical method in the many-beam theory in electron diffraction. J. Phys. Soc. Jpn, 42, 223–229.Google Scholar
First citation Krahl, D., Pätzold, H. & Swoboda, M. (1990). An aberration-minimized imaging energy filter of simple design. Proceedings of 12th International Conference on Electron Microscopy 1990, Vol. 2, pp. 60–61.Google Scholar
First citation Krivanek, O. L., Gubbens, A. J., Dellby, N. & Meyer, C. E. (1991). Design and first applications of a post-column imaging filter. Microsc. Microanal. Microstruct. (France), 3, 187–199.Google Scholar
First citation Krivanek, O. L., Mooney, P. E., Fan, G. Y. Leber, M. L. & Meyer, C. E. (1991). Slow-scan CCD cameras for transmission electron microscopy. Electron microscopy and analysis 1991, pp. 523–526. Bristol: IOP Publishing.Google Scholar
First citation Kühlbrandt, W., Wang, D. N. & Fujiyoshi, Y. (1994). Atomic model of plant light-harvesting complex by electron crystallography. Nature (London), 367, 614–621.Google Scholar
First citation Lehmpfuhl, G. (1974). Dynamical interaction of electron waves in a perfect single crystal. Z. Naturforsch. Teil A, 27, 424–433.Google Scholar
First citation Ma, Y., Rømming, C., Lebech, B. & Gjønnes, J. (1992). Structure refinement of Al3Zr using single-crystal X-ray diffraction, powder neutron diffraction and CBED. Acta Cryst. B48, 11–16.Google Scholar
First citation Marthinsen, K., Holmestad, R. & Høier, R. (1994). Analytical filtering of low-angle inelastic scattering contributions to CBED contrast. Ultramicroscopy, 55, 268–275.Google Scholar
First citation Matsuhata, H. & Gjønnes, J. (1994). Bloch-wave degeneracies and non-systematic critical voltage: a method for structure-factor determination. Acta Cryst. A50, 107–115.Google Scholar
First citation Matsuhata, H. & Steeds, J. W. (1987). Observation of accidental Bloch-wave degeneracies of zone-axis critical voltage. Philos. Mag. B55, 39–54.Google Scholar
First citation Matsuhata, H., Tomokiyo, Y., Watanabe, H. & Eguchi, T. (1984). Determination of the structure factors of Cu and Cu3Au by the intersecting Kikuchi-line method. Acta Cryst. B40, 544–549.Google Scholar
First citation Mori, N., Oikawa, T & Harada, Y. (1990). Development of the imaging plate for the transmission electron microscope and its characteristics. J. Electron Microsc. (Japan), 39, 433–436.Google Scholar
First citation Olsen, A., Goodman, P. & Whitfield, H. (1985). Tl3SbS3, Tl3SbSe3, Tl3Sb3−xSex and Tl3SbyAs1−ySe3. J. Solid State Chem. 60, 305–315.Google Scholar
First citation Saunders, M., Bird, D. M., Midgley, P. A. & Vincent, R. (1994). Structure factor refinement by zone-axis CBED pattern matching. Proceedings of 13th International Congress on Electron Microscopy, Paris, France, 17–22 July 1994, Vol. 1, pp. 847–848.Google Scholar
First citation Spargo, A. E. C. (1994). Electron crystallography and crystal structure. Proceedings of 13th International Congress on Electron Microscopy, Paris, France 17–22 July 1994, Vol. 1, pp. 959–960.Google Scholar
First citation Spence, J. C. H. & Zuo, J. M. (1992). Electron microdiffraction. New York: Plenum.Google Scholar
First citation Steeds, J. W. (1984). Further development in the analysis of convergent beam electron diffraction (CBED) data. EMAG 1983. Inst. Phys. Conf. Ser. No. 69, pp. 31–36.Google Scholar
First citation Taftø, J. & Gjønnes, J. (1985). The intersecting Kikuchi line technique: critical voltage at any voltage. Ultramicroscopy, 17, 329–334.Google Scholar
First citation Taftø, J. & Metzger, T. H. (1985). Large-angle convergent-beam electron diffraction; a simple technique for the study of modulated structures with application to V2D. J. Appl. Cryst. 18, 110–113.Google Scholar
First citation Tanaka, M. & Tsuda, K. (1990). Determination of positional parameters by convergent-beam electron diffraction. Proceedings of 12th International Congress on Electron Microscopy 1990, Vol. 2, pp. 518–519.Google Scholar
First citation Terasaki, O., Watanabe, D. & Gjønnes, J. (1979). Determination of crystal structure factor of Si by the intersecting-Kikuchi-line method. Acta Cryst. A35, 895–900.Google Scholar
First citation Vincent, R., Bird, D. M. & Steeds, J. W. (1984). Structure of AuGeAs determined by convergent beam electron diffraction. II. Refinement of structural parameters. Philos. Mag. A50, 765–786.Google Scholar
First citation Vincent, R. & Midgley, P. A. (1994). Double conical beam-rocking system for measurement of integrated electron diffraction intensities. Ultramicroscopy, 53, 271–284.Google Scholar
First citation Voigt-Martin, I. G., Yan, D. H., Gilmore, C. J., Shankland, K. & Bricogne, G. (1994). The use of maximum entropy and likelihood ranking to determine the crystal structure of 4-[4′-(N-dimethylamino)benzylidene]pyrazolidine-3,5-dione at 1.4 Å resolution from electron diffraction and high-resolution electron microscopy image data. Ultramicroscopy, 56, 271–288.Google Scholar
First citation Voss, R., Lehmpfuhl, G. & Smith, D. J. (1980). Influence of doping on the crystal potential of silicon investigated by the convergent beam electron diffraction technique. Z. Naturforsch. Teil A, 35, 973–984.Google Scholar
First citation Watanabe, D., Uyeda, R. & Fukuhara, A. (1969). Determination of the atomic form factor by high-voltage electron diffraction. Acta Cryst. A25, 138–140.Google Scholar
First citation Zou, X. D., Sukharev, Y. & Hovmöller, S. (1993). ELD – a computer program for extracting intensities from electron diffraction patterns. Ultramicroscopy, 49, 147–158.Google Scholar
First citation Zuo, J. M., Høier, R. & Spence, J. C. H. (1989). Three-beam and many-beam theory in electron diffraction and its use for structure-factor phase determination in non-centrosymmetrical crystal structures. Acta Cryst. A45, 839–851.Google Scholar
First citation Zuo, J. M., Spence, J. C. H., Downs, J. & Mayer, J. (1993). Measurement of individual structure-factor phases with tenth-degree accuracy: the 00.2 reflection in BeO studied with electron and X-ray diffraction. Acta Cryst. A49, 422–429.Google Scholar








































to end of page
to top of page