This content has been downloaded from IOPscience. Please scroll down to see the full text. Download details: IP Address: 130.92.9.55 This content was downloaded on 05/09/2014 at 14:38 Please note that terms and conditions apply. Electron interferometry and interference electron microscopy View the table of contents for this issue, or go to the journal homepage for more 1981 J. Phys. E: Sci. Instrum. 14 649 (http://iopscience.iop.org/0022-3735/14/6/001) Home Search Collections Journals About Contact us My IOPscience J. Phys. E : Sci. Instrum., Vol. 14, 1981. Printed in Great Britain REVIEW ARTICLE Electron interferometry and interference electron microscopy G F Missiroli, G Pozzi and U Valdri! Istituto di Fisica and GNSM-CNR, University of Bologna, Via Irnerio 46, 40126 Bologna, Italy Abstract A state-of-the-art review of electron interferometry and interference electron microscopy is given. The various types of interferometry device, interferometers and interference microscopes, which have been proposed and/or constructed are reviewed and commented upon. The electron biprism, by far the most successful interferometry device, is treated in some detail from both the experimental and theoretical (geometric and wave optics) points of view. The applications of electron interferometry are presented with particular reference to off-axis electron holography. Finally the future perspectives are indicated. 1 Introduction Electron interferometry is defined here as that branch of electron optics dealing with interference phenomena produced with the aid of suitable devices under instrumental and specimen conditions controlled by the experimenter. Its origin dates back to 1953 when the first interference pattern was obtained (Marton et a1 1953), but it was only with the introduction of the electrostatic biprism and with the assessment that the coherence of the electron wave is preserved when it passes through a thin specimen that electron interferometry was established as a practical proposition (Mollenstedt and Duker 1956, Mollenstedt and Keller 1957, Faget and Fert 1957a, b). Since then work has been carried out more or less erratically except in a very few laboratories. There has recently been a revival in the applications of electron interferometry and further development is expected in view of the availability of field emission sources, which have reasonably good stability, high brightness and high coherence. It is therefore felt that a review of the subject would be appropriate and timely. There have been several reviews of interferometry which have concentrated on the efforts of certain groups or institutions: Faget and Fert (1957b), Faget (1961) and Fert (1961, 1962) refer especially to work carried out in the Laboratoire d’Optique Electronique at Toulouse in a discussion of the general background, the instrumentation and applications while Mollenstedt (1960), Mollenstedt and Lenz (1962) and Wahl (1970~)have given brief reports on activity at the University of Tubingen; a more recent summary of this work has been given by Mollenstedt and Lichte (1979). Also, Hibi and Yada (1976) have given a review of some of the literature up to 1973, with special emphasis on the contribution of Japanese workers to this field. However, an up-dated, impartial and comprehensive review is not available, and it is one of the aims of the present article to fill this gap. As the nomenclature of electron interferometry instrumentation is still the reserve of specialists and often reflects personal taste, a few definitions are given here (and, when required, in the following pages), in order to avoid misunderstanding. The device used to produce two or more coherent waves, essential for the occurrence of interference, will be called an interferometry device. The instrument used to perform interferometry with no imaging facilities for the specimen will be designated an electron interferometer. This is usually an electron optical bench, capable of great electrical and mechanical flexibility, in particular for specimen preparation, treatment and handling. Since the specimen is out of focus, only simple types of specimen of known geometry can be used. The name (electron) interference microscope refers to a machine where it is possible to obtain an interference pattern by coherently superimposing different regions of a focused (or nearly focused) image of the specimen. This instrument is usually a conventional electron microscope equipped with an interferometry device and capable of forming an image of the specimen, in addition to displaying the interference pattern. In the following, the applications, capability and future perspectives of electron interferometry and interference electron microscopy will be presented and discussed. Section 2 contains a critical review of the various interferometry devices developed so far and is followed by $03 and 4, respectively covering the theoretical and practical features of the biprism, by far the most successful device, while 05 deals with applications of electron interferometers and electron interference microscopes. 2 Types of interferometry device This section will review the basic ideas and the various types of interferometry device prompted by them and will also make a critical appraisal of their performance and fields of application. The close similarity between light optics and electron optics is used to classify interferometry devices into two categories : (i) division of amplitude (e.g. Michelson’s interferometer in light optics) and (ii) division of wavefront (e.g. Fresnel biprism). It will be seen that the electrostatic biprism, which belongs to the second category, is by far the most popular device in electron optics, although, significantly, the other category is more popular in light optics. 0022-3735/81/060649+23 $01.50 0 1981 The Institute of Physics 3* 649 G F Missiroli, G Pozzi and U ValdrP 2.1 Division of amplitude 2.1.1 Three-crystal interferometry device. Historically, the first interferometry device to be conceived (Marton 1952) and actually built (Marton et a1 1953, 1954, Marton 1954) was in some ways the equivalent of the optical interferometer of Mach and Zehnder. It is based on the initial splitting (e.g. rays 1 and 2 in figure l), produced through Bragg reflection by a I’ Figure 1 Schematic representation of a Marton-type amplitude splitting electron interferometry device. I, incident beam; CI,CZ,CS,thin single crystals of identical material and orientation, accurately aligned; AI, AB,beam limiting apertures; E, exit beam formed by two interfering Bragg-diffracted beams. single crystal film CI, of a fairly coherent electron beam I, on the further splitting (Le., Bragg reflection, e.g. rays 3 and 4) produced by a second crystal C2,and on the recombination (emerging ray E) caused by a third crystal C3,again by Bragg reflection. The three crystals must satisfy several stringent requirements, namely they must have the same crystal structure and must be flat, parallel to each other and of the same orientation and register. The required geometry is obtained by providing a suitable support which allows movement along, and rotation around, the optical axis ( z ) for the first crystal C1,z rotation for C Zand z rotation and tilt for C3. A set of interference fringes is obtained on a photographic plate, placed in a plane which is conjugate, through a lens, with the exit surface of the third crystal. The lens is necessary in order to suitably magnify the fringe pattern; apertures A1 and A2 must be used in order to avoid the presence of unwanted beams. If an object is interposed in the path of ray 1 or 2, the set of fringes is altered (see 893.1.2 and 5.1). The distance between two consecutive copper crystals grown epitaxially to a thickness of 10nm was 35 mm, bringing the maximum separation of the split beams 1 and 2 to 0.7 mm. The Cu films were 3 mm in diameter and the accelerating voltage was60kV. Over 1200micrographs had to be taken, each with an exposure time exceeding 6 min, in order to establish beyond any doubt the occurrence of interference fringes. An elementary theory of the three-crystal interferometry device was developed by Simpson (1954) and good correlation was found with experimental data. It is clear that this method suffers from various inconveniences: (i) difficultandcrucialalignment of thethreecrystals; (ii) low intensity due to the various splittings of the incoming beam and also to absorption effects (in fact the alignment had to be performed by means of micrographs because the intensity of the fringes was below the threshold for their direct observation on the fluorescent screen); (iii) lack of versatility. An attempt to increase the intensity of the interference pattern (theoretically by two orders of magnitude) by replacing the second crystal with a lens also produced disappointing results (Simpson 1956), probably because of the aberrations introduced by the lens. For the above reasons the experiment has never been repeated, and its interest in electron microscopy is purely historical and academic. The principle of this method was, however, applied successfully ten years later in x-ray and neutron interferometry, where the crystal requirements are less stringent than those for electron interferometry (for a review, see Bonse and Graeff 1977). 2.1.2 Other methods. Other interference methods, based on amplitude splitting by crystals, have been proposed or developed; all of them use convergent-beam techniques. Li (1978) has proposed a modified version of the threecrystal Marton interferometer which is based on Fraunhofer diffraction rather than Fresnel diffraction. It should provide some advantages, but the basic criticism of the Marton interferometer still holds and, to our knowledge, experiments have never been attempted. The formation of interference fringes, by overlapping thin crystals in convergent-beam electron microscopy, in both the central beam and the diffracted orders, has been described by Dowell and Goodman (1973) and interpreted by Dowell (1977). Similar results have been obtained from two suitably orientated thin crystals separated by a gap in convergent-beam electron diffraction patterns (Rackham et al 1977, Buxton et a1 1978). In this case lattice parameters can be measured with an accuracy of 0.2% and it is claimed that double-crystal devices can be produced routinely by bombarding the edge of a crystal with ions until a gap is formed. Because of the small gap (10 pm wide at the most) between the crystals, the use of the system as a proper interferometry device is doubtful. Interference effects in the Fraunhofer diffraction pattern have also been observed by Berndt and Doll (1978). They obtained the beam splitting by amplitude division with a diffraction grating of 1 pm spacing, placed in the condenser aperture plane; the (periodic) object was a replica grating of 0.463 pm spacing, placed in the normal specimen plane. The coherent superposition of neighbouring reflections in the diffraction plane can be used to determine the phase of structure factors, as first suggested by Hegerl and Hoppe (1970), who introduced the terminology of ‘ptychography’ for this method. They had in mind the use of an electrostatic biprism to obtain wavefront division (see §2.2.5), but no further attention was given to the subject; on the other hand, Berndt and Doll (1976, 1978) have proposed the use of amplitude division obtained by a grating and have analysed, both theoretically and experimentally by, first, light-optics simulation and, more recently, by electron optics, the feasibility of the method. It should be noted in passing that lattice fringes are also formed by interference between the primary beam and the Bragg-reflected beam(s), as first observed by Menter (1956) (see also 95.8). Similarly, Fresnel fringes visible at the specimen edges are the result of interference between the direct beam proceeding through the vacuum and the beam scattered from the edge. 2.2 Dioision of wauefront 2.2.1 Electrostatic biprism and related deoices. The electrostatic field, produced by a straight, charged wire W placed between two earthed plates, splits the wavefront of incoming 650 Electron interferometry 8’ I I I Figure 2 Schematic electron ray path in an electron microscope equipped with a convergent biprism. G, electron gun; S’, primary electron source; C, condenser lens system; S, demagnified effective source; €3, biprism; W, biprism wire; OP, observation plane; P, projector lens system; VP, final viewing plane conjugate to the observation plane. electrons from the source S to produce two virtual, coherent electron sources SI and S Z (figure 2; see also figure 5). Interference takes place in the region where the two split beams overlap. The principles of this method are analogous to those of the Fresnel biprism in light optics and this interferometry device is therefore called an electron biprism. Figure 2 also shows diagrammatically the arrangement used by Mollenstedt and Duker (1955) in their early and successful work carried out in a conventional electron microscope of the electrostatic type. The cross-over S’ of the electron gun G is demagnified into the effective source S by the condenser system C in order to increase the lateral coherence of S, while a second lens system P below the biprism B is used to obtain a magnified image of the unlocalised fringe pattern present in the region of beam overlapping. This represents the type of arrangement commonly used in electron interferometry. Let the observation plane OP, conjugate to the viewing screen VP (photographic plate), be below the biprism wire W and the effective source S be above the wire as depicted in figure 2. When no voltage is applied to the wire, its shadow is observed together with a system of diffraction fringes (see figure 6(a)). By increasing the positive voltage applied (convergent biprism), the shadow can be seen to shrink and a system of real interference fringes appears, if the coherence condition (equation (Il), 043.1.1 and 5.6) is satisfied (see figures 6(b) and (c). By applying a negative voltage (divergent biprism), the two electron beams in the region below the wire diverge and the width of the region not illuminated by the beam increases. A virtual interference field is produced above the wire; sections of this field can be imaged by a suitable excitation of lens system P. It should be noted that in the case of a convergent biprism, the lens system P is used to form an enlarged image of the fringe pattern in the VP, while in the case of the divergent biprism, P has the additional, essential function of producing a real image of the virtual fringe system ($3.2.2). One advantage of the biprism is that the fringe spacing can be varied over a wide range to suit the observation conditions, simply by changing the voltage applied to the wire (53.1.1, equation (8)). Furthermore, there is no intensity splitting and, since the electron beam travels all the time in free space, unlike in the three-crystal method, the beam intensity is not reduced. However, impracticable exposure times were required even in such favourable conditions, so that Mollenstedt and Diiker were forced to adopt electron optical devices of cylindrical symmetry (linear illumination and cylindrical lenses). With this approach they were able to observe, with the naked eye, interference fringes projected directly on to a fluorescent screen (Diiker 1955, Mollenstedt and Duker 1956). Since then the experimental conditions and the construction of the biprism have been improved, as we will see in 44. More details concerning the properties and the operation of the biprism will be given in 53. Interferometry devices which are based on the biprism will be outlined in 5§2.2.1.1,2.2.1.2, 2.2.4 and 2.2.5. 2.2.1.1 Wire cascade biprisms. Several wires at suitable potentials are placed one above the other. The aim is to increase the separation between the two split beams, without breaking the spatial coherence condition, in order to insert an object between them. Mollenstedt and Bayh (1961) have described a system formed by three biprisms in cascade (at distances of 470 mm and 50 mm) to obtain both the separation of 60 pm required for experiments on the Aharonov-Bohm effect ($5.3) and a fringe system directly visible on the fluorescent screen. The separation depends on the voltage applied to the first wire (divergent biprism) and on the distance between the first and second wires (convergent biprism). However, as well as increasing the separation of the two interfering beams, the first two biprisms produce an increase in the distance between the two virtual sources so that the fringe system cannot be resolved directly on the fluorescent screen ($3.1.1, equation (8)). Hence the necessity of using a third wire (divergent biprism) to restore visibility. In a modified version, Schaal et a1 (1966/7) have replaced the second biprism with an electrostatic lens to obtain a beam separation of up to 120 pm. 2.2.1.2 Multiple beam devices. Multiple beam interference has attracted the attention of some experimenters because it has, at least in theory, several advantages: (i) high sensitivity to phase changes, (ii) high intensity maxima (N-beam interference should produce maxima tA’2 times those of a twobeam pattern) and (iii) narrow maximum width. Buhl (1961a) was the first to use three-beam interferometry, while Anaskin and Stoyanova (1967, 1968d, e) and Anaskin et a1 (1968b) have performed more extensive experiments, using up to five beams, with a parallel array of biprisms obtained by mounting several wires on a hollow, optically polished support. In order to change the distance between the virtual electron sources (43.1.1), the plane formed by the wires could be tilted with 651 G F Missiroli, G Pozzi and U Valdrb respect to the optical axis of the electron microscope where the interferometry device was mounted. However, these devices suffer from several disadvantages : the parallelism of the wires is strictly stipulated, the device is sensitive both to irregularities and differences in the wire diameter, and to unequal voltage, and an accurate orientation of the wires with the optical axis is required in order to produce true N-beam interference. The theoretical expectations of the performance of these devices have been confirmed satisfactorily. However, the finite size of the sources produces fringe intensities which are smaller than the calculated values; in the case of three-beam interference, the intensity with respect to the two-beam case is 1.4 instead of the theoretical value of 4 =2.25 (Anaskin and Stoyanova 1967). 2.2.1.3 Charged film edge. Interference patterns have been obtained near the edge of a charged thin film (placed between two earthed grids) as a result of the superposition of the beam transmitted through the specimen and of the beam travelling in free space near the film edge (Anaskin et al 1968a). This second beam is deflected by the electrostatic field created by the applied voltage. The experiments were performed in an electron microscope by suitable defocusing on thin (30-50 nm) carbon and collodion films. The interference pattern is visible near the edge of the film and, in the case of positively charged films, appears against the background of its low magnification shadow image (650 x ). The character of the fringe pattern is similar to that of a conventional biprism. The authors have suggested that this method could be applied in electron holography; however. the lack of control on the geometry of the edge may pose severe limitations. Similar effects are produced by the electrostatic field created by charged specimens under beam irradiation. 2.2.2 Magnetic biprisms. A magnetic field is used to split the wavefront. This field may be produced by electromagnets (e.g. quadrupoles, Krimmel 1960) or by the inner magnetisation of a thin ferromagnetic film in the region where the direction of magnetisation changes (Boersch et a1 1960, 1962b). In the latter case the method has been used only to study the distribution of the magnetisation within the domain walls (for a recent review, see Jakubovics 1976). In fact, as an interferometry device, it suffers from inconveniences similar to those of the Marton interferometer: firstly, loss of intensity and coherence in the beam when passing through the magnetic material and, secondly, a lack of versatility. The magnetic quadrupole is, on the contrary, free from the above disadvantages and also from the spurious effects introduced by the shadow of the wire in the biprism; furthermore it produces achromaticfringes. However, no applications have been reported in the literature, probably because the interpretation of the interference effects is very complicated (Krimmel 1961) and therefore it is difficult to separate these effects from those to be investigated in a specimen. 2.2.3 Young interferometers. Faget and Fert (1956, 1957b) were the first to show the occurrence of fringes from the interference of two electron waves transmitted by two circular holes (Young fringes). In order to obtain the required coherence with their thermo-ionic electron source, the beam was so defocused that the resulting intensity at the image plane was too low, the exposure times were too long and the disturbing effects were too great; in fact their results are not comparable with the calculated fringe pattern. The experiment has recently been repeated for the sake of completeness by Ohtsuki and Zeitler (1977) with the aid of a field emission source, and impressive results have been obtained. In all the above experiments, it is the accidental occurrence of holes at a suitable distance in a carbon or gold film which provides the secondary sources. On the contrary, Mollenstedt and Jonsson (1959), Jonsson (1961) and Jonsson et al(l974) managed to prepare multiple (up to five) slits (50 pm long, 0.3 pm wide, with a spacing of 1 pm) in controlled conditions, and were able to produce Young fringes in agreement with the theory, to a first approximation. Deviations of the patterns from the theoretical predictions are attributed to the finite width of the slits. It should be emphasised that in this experiment, contrary to the light optics analogue, the coherent illumination of the slits could be an a priori problem. In fact, the dimensions of the slits are over lo6times greater than the electron wavelength (see also §5.6), while in light optics the slits can easily be made even smaller than the wavelength of light. Jonsson (1961) demonstrated that the region of lateral coherence ($5.6) was up to 60 pm across the slits. 2.2.4 Electron mirror interferometer. The electrons emitted from a source S and accelerated by the potential V of anode A1 are split into two partial beams 1 and 2 by the divergent biprism DB (figure 3); they enter a magnetic prism M of the Figure 3 Schematic set-up of the electron mirror interference microscope. S, electron source; DB, divergent biprism; 1, 2, split beams; M, magnetic prism; AI, gun anode; Az, mirror anode; 0, specimen; I, intermediate image plane; CB,convergent biprism; LI, electrostatic lens; Lz,intermediate lens. Castaing and Henry (1964) type, which deflects them by 90’ towards an electron mirror. The latter is formed by an anode A2 (held at the same potential V as AI), and by a (bulk) specimen 0,kept at a negative potential Vo( I Vo I > 1 VI ). The electrons of the two partial beams are decelerated in the essentially homogeneous mirror field and return at the 652 Electron interferometry distance of closest approach (plane P) in front of the specimen. While being reflected, the phase of the electron wave is influenced by the local electrostatic field, which in turn is modulated by the topography of the specimen. The reflected electrons are accelerated by A2 towards M and are bent by another 90" (not indicated in figure 3 for clarity). The setting is such that the focal plane of the electrostatic lens L1 is coincident with the plane of the real stigmatic point of the magnetic prism, and the image plane is coincident with the plane passing through the virtual achromatic and stigmatic point. The combined effect of L I and M produces a virtual intermediate image (one for each of beams 1 and 2) of the specimen in the intermediate image plane I. At the exit from M, the partial beams are deflected towards each other by a convergent biprism CB, so that the two virtual images are superimposed in plane I to form a virtual interference field. A set of lenses L P placed below CB provides a magnified real image in the viewing plane of the interference pattern, which results from the difference in the phase modulation of the two partial beams 1 and 2 (Lenz 1972). This electron mirror interferometer has been developed by a German team (Lichte et a1 1972, Lichte 1977, Lichte and Mollenstedt 1979) with a 25 kV field emission source of 20 nm diameter and a divergence of 2 x rad (full angle). The two partial beams have a width of about 5 pm at the specimen and are separated by about 10 pm. The device is the analogue of the Michelson interferometer in light optics (Lichte et a1 1972), with the difference that here parallel (instead of perpendicular) mirrors are used. Details of its applications in the study of surface roughness and microfields are given in $5.5, and of the Doppler effect in $5.7. The electron mirror interference microscope has also been used (Mollenstedt and Lichte 1978a, 1979) to obtain interference with an arrangement equivalent to the double mirror of Young and Fresnel in light optics. The double-mirror electrode is a conducting echelette grating, held at a potential slightly negative with respect to the cathode, with a lattice constant of 1.6 pm and groove angles of 5" and 85". Close to the grating (at about 1 pm), the reflecting potential varies with the grating periodicity, as do the distance of closest approach and the phase modulation of the reflected electron wave. The incident wave is consequently reflected as if it had impinged on a double mirror with a width equal to the grating spacing, i.e. as if two partial waves were originated by two coherent virtual sources. 2.2.5 Wavefront and amplitude division interferometer. A combination of coherent amplitude division produced by a single-crystal film (according to Marton) with wavefront splitting produced by an electron biprism (according to Mollenstedt and Diiker) has been used by Matteucci and Pozzi (1980). The diffraction spots formed in the back focal plane of the objective lens when a thin single crystal is irradiated by a parallel beam are used as multiple coherent sources of electrons to illuminate the biprism. Experiments have been performed in the simplest case of two Bragg spots symmetrically placed with respect to the biprism wire. Four virtual sources are created and a system of lattice fringes is observed in the image plane (conjugate to the specimen plane) with two shadows of the biprism wire superimposed. The crystal has the double function of producing an amplitude splitting of the beam and of being the specimen to be studied. This arrangement should be suitable for producing off-axis image holograms of specimens placed over the crystal and it might allow the study of crystallographic defects present in the crystal itself. However, further experimental work and a theoretical treatment (lacking at the present) is necessary in order to ascertain if relevant information about the object can actually be obtained. Kunath (1978) has proposed (see also Hegerl and Hoppe 1970 and $2.1.2) the insertion of the biprism at the level of the condenser aperture in order to obtain two diffraction patterns from a crystalline specimen and to produce interference between two spots of different diffraction order (superimposition is obtained by adjusting the biprism voltage) for the determination of the phase of the structure factors. 3 Fundamentals of biprism interferometry The basic theoretical details related to the physics of the electron biprism and to its application in electron interferometry and interference microscopy will be presented in this section. 3.1 The electron biprism and interferometer Since most of the properties of the electron biprism can simply be derived by particle physics considerations, we will follow this approach as far as possible before introducing the wave mechanics treatment. 3.1.1 Geometrical treatment. A biprism is formed by a long, conducting wire of radius r charged at a voltage VB and placed symmetrically between two plates which are kept at ground potential as shown in figure 4, where the axis of the wire is taken in the y direction (normal to the foil) and the electrons are assumed to travel parallel to the z axis. X Figure 4 The electric field distribution near the wire of radius r of a biprism is well approximated by the field of a cylindrical condenser of external radius R and inner radius r. xi is the impact parameter of the electron e. The properties of the electric field generated by the biprism have been studied by Mollenstedt and Diiker (1956) by means of an electrolytic tank in a model one hundred times larger than the actual biprism. The latter was formed by a wire (6 mm long, radius 1 pm) placed at a distance of 2 mm from the two plates (5 mm long, 2 mm high). It was found that the two-dimensional ( x , z ) field distribution so obtained in the region around the wire (where essentially the electron deflections take place) very closely resembles that of a cylindrical condenser with an outer electrode radius R which is of the same order of magnitude as but smaller than the distance between the two plates. A good approximation of the potential distribution V ( x ,y ) in a point P(x, z ) of the biprism is therefore obtained by 653 G F Missiroli, G Pozzi and U Valdrd assuming that for the screened condenser field (Komrska 1971): V ( x ,z ) =0 for R2<(x2+z2). (1) An electron of mass m and velocity uo travelling along z at a distance xi from the wire suffers an angular deflection /3 given by (weak-field approximation) (2) If the wire is positively charged (converging biprism) the electron is deflected towards the wire. In practice, the superposition between the electrons which have been deflected by the wire in opposite direction occurs only for those electrons which travel with an impact parameter xi of the order of several micrometres. Since the value of R is of a few millimetres, R S x i and equation (2) can be written as (3) 01 is the absolute value of p, whose sign depends on that of the impact parameter xi. A typical value for a: is lO-4rad. It is important to note that in this approximation the angular deflection is independent of the impact parameter. The predictions of equation (3)have been confirmed experimentally (Mollenstedt and Duker 1956) up to deflections of 5 x rad for 44 kV electrons with V B=90 V and xi = 100 pm. The latter two values are greater than those used in practice. The independence of iy. from xi and the absence of distortion has also been demonstrated theoretically by Septier (1959), who derived an analytical expression for the potential distribution for the case of a biprism with a wire of radius zero placed symmetrically between two earthed planes. Experiments by the same author, performed with 30 kV electrons in an enlarged model where the ratio between the wire diameter and the plate distance was 1:20, have confirmed the validity of the theoretical results up to deflections of 4 x 10-2 rad. As a consequence of the above property, if the biprism is illuminated by a monochromatic electron point source S on the z axis (figure 5), the charged wire will split the electron beam into two partial beams which will appear to originate from two virtual, coherent point sources SI and S Z . They are placed symmetrically with respect to the real source S, and lie in a plane perpendicular to z and containing S at a distance a from the x axis; their separation is d= 2 I aa 1 , (4) Typical values for d are 10-50 pm. In the region below the wire, where electrons originating from S I and S 2 come together (overlapping region, hatched in figure 5 ) , a non-localised interference pattern will be produced. The intersection of the overlapping region with a plane OP (obsercationplane) normal to the z axis at a distance b from the wire and conjugate to the final viewing plane (fluorescent screen of photographic plate) will be called the interference field. The width of the interference field W normal to the wire is given by Figure 5 Basic parameters of an electron biprism illuminated by a monochromatic electron point source S placed on the z axis. SI, S Z ,virtual sources at distance d ; 01, deflection angle; 11, I n , trajectories of two electrons passing through points Pol, P0z in the specimen plane SP and deflected at P in the observation plane OP (instead of passing through PI and P2) when the biprism is activated; D, interference distance; W, width of the interference field; RB, reference beam for an off-axis specimen (on the RHS). It is seen that the interference field will exist ( W >0) if a (;SI>r. Inequality (6) is an algebraic relation where a and b are conventionally taken to be positive for the case sketched in figure 5 (source S above the wire and observation plane under the wire), and negative in the opposite case, while the sign of the deflection a: is the same as that of the applied voltage (VB> 0, convergent biprism; V B<0, divergent biprism). The electrons passing through the same given point P (figure 5) of the observation plane OP when the biprism is activated would have arrived at points PI and PZ in the absence of the biprism; the distance D between these points, termed interference distance (see also §3.2),is given by D=2labl. (7) Equation (5) shows that W would coincide with D for a wire of infinitely small radius. In the interference field, a set of straight fringes running parallel to the wire may be observed under suitable conditions to be discussed below (see figure 6). The fringe spacing, s, by analogy with the Fresnel biprism, is given by s= A [ (a+b)/dI (8) where X is the electron wavelength. Since s is of the order of 0.1 pm, a lens system providing a suitable magnification must be arranged for seeing and recording the fringes on the viewing plane. The number of fringes N contained in W is given, from 654 Electron interferometry equations (5) and (8), by N = W/s= I( a + h)/ahI(A Ih/sI -2r). (9) For the fringes to be visible, their contrast should be (10) where I m a x and I m i n are the intensities at the maximum and at the minimum respectively (see also $5.6), it can be shown that C is very good, in fact not less than 0.9 (Francon 1956; the different definition of contrast here used is irrelevant to this result) if the transverse source dimension 6 (in the x direction) fulfils the condition (1 1) This condition states that the angular size of the source must be smaller than the angular size of the fringes when both are seen from the biprism wire. It is equivalent to the well known lateral coherence condition since equation (1 1) can be written as suitably high. By defining the contrast of the fringes as C = (Imnx-Imin)/(lmax+Imin) 6< Im/4h 1 . D