Vous êtes sur la page 1sur 8

Ultramicroscopy 111 (2011) 1461–1468

Contents lists available at ScienceDirect

Ultramicroscopy
journal homepage: www.elsevier.com/locate/ultramic

Theory of free electron vortices


P. Schattschneider a,b,, J. Verbeeck c
a
Institute for Solid State Physics, Vienna University of Technology, A-1040 Wien, Austria
b
Service Centre for Electron Microscopy, Vienna University of Technology, A-1040 Wien, Austria
c
EMAT, University of Antwerp, B-2020 Antwerp, Belgium

a r t i c l e i n f o abstract

Article history: The recent creation of electron vortex beams and their first practical application motivates a better
Received 16 February 2011 understanding of their properties. Here, we develop the theory of free electron vortices with quantized
Received in revised form angular momentum, based on solutions of the Schrödinger equation for cylindrical boundary condi-
4 July 2011
tions. The principle of transformation of a plane wave into vortices with quantized angular momentum,
Accepted 14 July 2011
their paraxial propagation through round magnetic lenses, and the effect of partial coherence are
Available online 23 July 2011
discussed.
Keywords: & 2011 Elsevier B.V. Open access under CC BY-NC-ND license.
TEM
Coherence
Angular momentum
Vortex beams

1. Introduction than optical ones because the electron is described by a scalar


wave equation1 (the Schrödinger equation) whereas optical
Electron vortices are free electrons carrying orbital angular vortices are vector fields. But for unpolarized light the description
momentum. They are characterized by a spiraling wavefront with of vortex modes can be simplified, and in the paraxial approx-
a screw dislocation along the propagation axis. With the publica- imation they have in fact been described by a Schrödinger type
tion of two seminal papers on the creation of electron vortices in equation [11]. One of the surprising and original features of
the electron microscope [1,2] the matter was transformed from a electron vortices is that, according to their rotational component
theoretical possibility [3] to reality. The first practical application of probability current they carry also a magnetic moment, even
was a filter for magnetic transitions in the ferromagnetic 3d for beams without spin polarization. This, together with the much
metals, thus facilitating experiments in the field of energy loss smaller scale of electron vortices compared to light, makes them
magnetic chiral dichroism (EMCD) [4]. The potential of vortices is so attractive.
much wider, ranging from probing chiral structures to manipula- We present here a theory of electron vortices based on
tion of nanoparticles, clusters and molecules, exploiting the solutions of the Schrödinger equation. We assume cylindrical
magnetic interaction [5]. geometry, starting with the general case allowing external scalar
For a review of vortex applications in optics, see [6]. The and vector potentials. Then we concentrate on potential-free
theory of field vortices was first described in the 1970s [7] in the systems, having in mind vortices propagating freely in space.2
context of dislocations in sound waves; it took almost one decade It turns out that the solutions fall into families with quantized
before such vortices in the optical frequency range were pro- angular momentum ‘ m along the optic axis.
duced [8]. It was then realized that optical vortices were related
to angular momentum of light, which already Poynting in 1909
speculated about [9]. The two formerly rather separated fields of 2. The Schrödinger equation in cylindrical geometry
research began to merge. The first experimental demonstration of
laser light carrying quantized orbital angular momentum came in We assume an electron propagating along the z axis of the
1992 [10]. microscope and defined within a circle of radius rM (the aperture).
In the context of orbital angular momentum optical vortices There may or may not be a magnetic field present in the space
are similar to electron vortices. In a sense, the latter are simpler where the free electron propagates. We look for the general

1
For standard electron microscope probes, which are not spin polarized.
 Corresponding author at: Institute for Solid State Physics, Vienna University of 2
The magnetic field of the lenses in the electron microscope is weak enough
Technology, A-1040 Wien, Austria. such that its effect can be compensated by assuming a slowly rotating reference
E-mail address: schattschneider@ifp.tuwien.ac.at (P. Schattschneider). frame during propagation through the lens [12].

0304-3991& 2011 Elsevier B.V. Open access under CC BY-NC-ND license.


doi:10.1016/j.ultramic.2011.07.004
1462 P. Schattschneider, J. Verbeeck / Ultramicroscopy 111 (2011) 1461–1468

solution of the Schrödinger equation. A natural coordinate system 3. Electrons in field free space
is cylindrical for this geometry. The Hamilton operator in SI
units is We consider the case of an electron limited by a circular
aperture. We also assume that no potentials exist. The case of
1 ^ 2 þeV^ , non-vanishing magnetic flux will be considered elsewhere.
H¼ ^ AÞ
ðpe ð1Þ
2M

where M is the electron mass, e its charge, A^ and V^ are the vector 3.1. Boundary conditions
and scalar potential operators, and p^ is the momentum operator.
The first term can be written as The boundary condition at the rim of an aperture of radius rM
centered at the origin is
2 2
p^ eðp^ A^ þ A^ pÞ
^ þe2 A^ : RðrM Þ ¼ 0: ð8Þ

With rðA^ cÞ ¼ crA^ þ A^ rc and (in Coulomb gauge) r  A^ ¼ 0, We look for solutions inside the aperture consistent with Eq. (8)
so that and match them with the trivial solution outside the aperture
Rðr Z rM Þ ¼ 0. Since we deal with free electrons there are no scalar
1 2 2 or vector potentials. Expressing the energy E of the electron by the
H¼ ðp^ 2eAp^ þ e2 A^ Þ þeV^ : ð2Þ
2M wave number of its plane wave solution

We write the time-independent Schrödinger equation ‘ 2 k20


E¼ , ð9Þ
2M
H9cS ¼ E9cS
multiplying Eq. (5) with r2, filling in solutions Eqs. (6) and (7), and
in cylindrical coordinates ðr, j,zÞ. In these coordinates the nabla rearranging terms we obtain the differential equation
operator reads R00 R0
r2 þ r þ r 2 k2r m2 ¼ 0 ð10Þ
  R R
@ @ @
r¼ , , : for the radial part of the wave function. We have used the
@r r@j @z
abbreviation
Now that p^ ¼ i‘ r we have qffiffiffiffiffiffiffiffiffiffiffiffiffiffi
kr ¼ k20 k2z : ð11Þ
 
@ Af @ @
A^ p^ ¼ i‘ Ar þ þAz , The substitution x ¼ rkr and yðxÞ :¼ Rðx=kr Þ yields
@r r @j @z
y00 y0
  x2 þ x þ x2 m2 ¼ 0: ð12Þ
@2 @ @2 @2 y y
2 2
p^ ¼ ‘ þ þ þ :
@r 2 @r r 2 @j2 @z2 This is Bessel’s differential equation. The solutions are Bessel
functions of the first kind Jm and the second kind Ym with index m.
For the special case of a constant magnetic field along z and a The Ym are divergent at x ¼0 and can be excluded for round
vanishing electrostatic potential V ¼0 (which will be important as apertures centered on the optic axis. The radial solutions
an approximation to magnetic lenses in a follow-up paper) we
Jm ðxÞ ¼ Jm ðrkr Þ ð13Þ
can conveniently use a gauge where Ar ¼ Az ¼ 0, so
are indexed according to quantum numbers kr and m. The factor
A2 ¼ A2f :
Z(z) depends on the quantum number kz. The energy spectrum is
The Schrödinger equation with Hamiltonian Eq. (1) reads given by Eq. (9) as
"  #
‘
2
@2 @2 @2 2eAf @

eAf 2 ‘2
@ Ekr ,kz ¼ ðk2r þ k2z Þ 40:
þ þ þ i  c ¼ Ec: ð3Þ 2M
2M @r 2 r@r r 2 @j2 @z2 ‘ r @j ‘
Note that the energy is always positive because the particle is
Separation of variables free. States are degenerate in the magnetic quantum number m.
Note that without magnetic field m is integer.3 It is important to
cðr, j,zÞ ¼ RðrÞYðjÞZðzÞ ð4Þ
note that the parameter kr is real since kz r k0 .
inserted into Eq. (3) transforms the partial derivatives into simple
derivatives 3.2. Special case
 
R00 R0 2eAf Y 0 Y 00 Z 00 eAf 2 2ME The special case kr ¼0 gives a particular solution
þ i þ 2 þ  ¼ 2 : ð5Þ
R rR ‘ rY r Y Z ‘ ‘ RðrÞpr 7 m :
The separation of variables is consistent with the fact that the In order to avoid singularities at r ¼0 the exponent must be
Hamiltonian of a rotationally invariant system (free space or a positive; it must also vanish at the rim of the beam defining
radially symmetric vector potential) commutes with L^ z and with P^ z ; aperture with radius r ¼ rM which is impossible for m a0. For
as a consequence, the eigenfunctions of P^ z : m¼0 the solutions read (going back to Eq. (12) for this special
Zpeizkz ð6Þ case)
Rpðr 2 rM
2
Þ:
and of L^ z :
That is to say that kr ¼0 does not allow solutions with m a0, nor a
Ypeimj , ð7Þ
constant intensity within the aperture That means that m a0
where m is integer because it must be 2p-periodic in the azimuth
due to periodic boundary conditions are also eigen functions of the 3
Non-zero magnetic fields define a complete set of Bessel functions with
Hamiltonian. non-integer index.
P. Schattschneider, J. Verbeeck / Ultramicroscopy 111 (2011) 1461–1468 1463

solutions inevitably have /kr S 4 0 and /kz S ok0 . In the next 2.0
section we discuss properties of beams with discrete angular
momentum, including /L^ z S ¼ 0.
1.5
3.3. Solutions with quantized angular momentum

Int. [arb.u.]
The general solution of the present problem for kr 4 0, Eq. (4), is 1.0
cðr, j,zÞ ¼ eizkz eimj Jm ðrkr Þ:
It is convenient to classify the solutions according to the eigenvalue
0.5
m of Eq. (7), as
X
cm ðr, j,zÞ ¼ eizkz eimj cðkr ÞJm ðrkr Þ, ð14Þ
kr
0.0
with arbitrary coefficients cðkr Þ. The reason is that cm is an 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
eigenfunction of L^ z with eigenvalue ‘ m. To see this, one may r [mu]
calculate the expectation value directly: the angular momentum of Fig. 1. Plane wave filling a round aperture of 1 mrad. First 10 basis functions.
the electron is

/L^ z S ¼ /c9L^ z 9cS: 2.0


In cylindrical coordinates the angular momentum operator reads
@
L^ z ¼ i‘ : 1.5
@j^

For solutions belonging to the Jm family, Eq. (14) we find Int. [arb.u.]
Z Z
@ 1.0
/L^ z S ¼ i‘ cn c dV ¼ ‘ m r dV ¼ ‘ m: ð15Þ
@j
The boundary condition Eq. (8) reduces the allowed values of kr to a
discrete set 0.5

kr A flnm =rM ,n A Ng,


where the lnm are the zeros of the Bessel function Jm. For m¼0 a 0.0
frequent situation is given by a wave function that is constant 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
within the round aperture and vanishes outside, c0 ¼ Pðr=rM Þ r [mu]
where P is the radial step function
( Fig. 2. Amplitude of a helical wave with m¼ 1 filling a round aperture of 1 mrad.
1, x r 1, First 10 basis functions.
PðxÞ ¼
0, x 4 1:

Expanding c0 into eigenfunctions of this geometry we obtain The result is given in Fig. 2. Contrary to the m¼ 0 case, we see here
X the amplitude dropping to zero at the origin despite the fact that
Pðr=rM Þ ¼ cn J0 ðln r=rM Þ, ð16Þ we assumed a constant amplitude. Expansion to higher order will
with make the slope steeper, eventually leaving a ‘pinhole’ at the
origin. This is a consequence of the central vortex line where
2 the phase is undefined, also known as topological charge.
cn0 ¼ ,
ln J1 ðln0 Þ
which by definition of the zeros ln0 vanishes at the rim r ¼ rM .
This gives a set of 4. Focussed vortices
 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 
kz,n ¼ k20 ðln0 =rM Þ2 ,n A N : The solutions discussed in the previous section had constant
amplitude over the aperture. Can we focus such extended beams?
As a demonstration example we assume an aperture of 1mm in The case m¼0 is trivial—it is exactly what happens when one
Fig. 1; the first 10 coefficients give already a good approximation. illuminates an aperture in the electron microscope with a plane
This solution is of type J0; from Eq. (7), its angular momentum is wave. An ideal lens will create an Airy disk in the diffraction (back
/L^ z S ¼ 0. focal) plane. The action of the lens is described by a Fourier
The relative deviations of kz from k0 for a 200 kV incident beam transform.
are of the order of 10  10 which translates into 10  5 eV, so the We can perform such transforms for beams with any quantum
splitting of the kz levels is unobservable in practice. The set of number m a 0. According to the Fourier–Bessel transform of a
different kr creates the oscillations within the aperture, visible in function F ¼ f ðrÞeimj with azimuth angle j,
Fig. 1. In combination with the set of kz they propagate the Z
im imj rM
solution into z direction. F~ ðqÞ ¼ e f ðrÞJm ðqrÞr dr,
For m¼1, the solutions are Bessel functions J1. By analogy to 2p 0

the case m¼0 we find the expansion for a wave with constant the Fourier transform of cm is
amplitude in the aperture in the family Jm, cm ¼ eikz z eimj Pðr=rM Þ: Z
X im imj rM
cm ðr, j,zÞ ¼ eikz z eimj cnm J1 ðln r=rM Þ: c~ m ðqÞ ¼ e Jm ðqrÞr dr: ð17Þ
2p 0
1464 P. Schattschneider, J. Verbeeck / Ultramicroscopy 111 (2011) 1461–1468

1.0 as can be verified by inserting the cn. Eventually,


2 X
rM
c~ ðqÞ ¼ i2m eimj cnm :Jm
2
þ 1 ðln,m Þ: ð20Þ
0.8 8p2 mn

Note that each m has different zeroes ln,m .


0.6

0.4
5. Creation of vortex electrons

0.2 Vortex electrons can be created with spiral phase plates [1] or
with holographic masks [2]. We illustrate the working principle
underlying the creation of electron vortices, a plane wave passing
0.0
through a holographic mask. Since the subject of this paper is the
0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14 theory of electron vortices and not any technical aspect we
angle [microrad] restrict the discussion to the elementary form of a fork dislocation
Fig. 3. Radial intensity distribution scaled to the maximum value in the back focal
in the following.4
(i.e. diffraction) plane of a lens for vortex waves filling a round aperture of 50 mm A plane wave is by definition infinitely extended and can
diameter, with m ¼0 (full line) m¼ 1 (dashed), m¼ 2 (dash-dotted), and m ¼3 therefore not be a solution of the present case. We ask for
(dotted). Cs ¼0, 200 kV. Abscissa: scattering angle in mrad. Zero denotes the vortex solutions that match a plane wave within the aperture and vanish
centers.
outside. I.e. we simulate a plane wave eikz z impinging onto a round
aperture with absorptive structure given by the transmission
We may calculate the far field solution via Eq. (17), equivalent to function
the diffraction pattern produced by an ideal lens. The intensity
2
Im ðqÞ ¼ 9cm 9 is isotropic. The radial intensity profiles are shown in T ¼ 12Pðr=rM Þð1cosðkx rjÞÞ, ð21Þ
Fig. 3. Only the m¼0 beam has a maximum at the center; it is the
where j is the azimuthal angle in the aperture plane, and
well known Airy disk. The vortex beams are characterized by a
kx ¼ 2p=d with the lattice distance d. The exit wave function is
ringlike structure with a maximum position depending on the
helicity m. Note that contrary to intuition, according to Plancheret’s c ¼ eikz z T ¼ eikz z ð12Pðr=rM Þ þ eikx r Pðr=rM Þej þ eikx r Pðr=rM Þej Þ,
theorem, the total intensity is equal for each m provided the cm are
ð22Þ
equally normalized. The helicity of a beam can then be determined
by measuring the intensity as a function of momentum q ¼ 9q9. where kz is the wave number of the fast electron that propagates
This is the basis for further analysis of vortex beams. It should be in the z direction. Matching this wave function to the boundary
noted that the focussed vortex has the same angular momentum as conditions of cylindrical geometry means a decomposition of c
the original one since the azimuthal part is still of type eimj . More into eigenfunctions of a round aperture. We have already encoun-
generally, after passage through the mask, the electron remains in tered the first term in brackets—expression 16 for a plane wave.
the same eigenstate of the angular momentum operator because The other terms belong to the family of vortex solutions with
the angular momentum is a constant of motion in free space. m¼ 71,

cm ðrÞpPðr=rM Þeimj : ð23Þ


4.1. Arbitrary wave forms
The phase factor e 7 ikx x þ kz z describes propagation of the partial
Any wave function can be expanded into the eigenfunctions of waves under the angle 7arctanðkx =kz Þ with respect to the optic
the cylindrical geometry described here, axis, in direction kx, caused by diffraction on the periodic mask.5
The transmission function of the holographic mask is shown in
X
cðrÞ ¼ fm ðrÞeimj : Fig. 4.
m The diffraction pattern of the holographic mask is shown in
Fig. 5. The phase of the wave function is coded as hue, the range
Its FT is
½p, p corresponding to colours continuously changing from blue
X im Z 1 X im to red (rainbow chart). The central Airy disk and its first two
c~ ðqÞ ¼ eimj fm ðrÞJm ðqrÞr dr ¼ eimj f~ m ðqÞ: ð18Þ concentric rings are well visible. The m ¼ 71 side bands show the
m
2 p 0 m
2p
characteristic volcano shape of vortices.6 Whereas the central Airy
For amplitudes fm(r) defined within a finite aperture one may use disk has a cylinder symmetric phase, the vortices show azimuthal
a series expansion similar to Eq. (16). phase variation over 2p. Compare also the radial phase jumps at
X the central pattern (blue to red to blue) with the more gradual
fm ðrÞ ¼ cnm Jm ðln,m r=rM Þ, phase variations at the vortices. A trace through the diffraction
n
pattern is shown in Fig. 6. Here the typical volcano profile of the
with side bands is clearly visible.
Z rM
2
cnm ¼ 2 J2
fm ðrÞJm ðln,m r=rM Þr dr: 4
rM m þ 1 ðln,m Þ 0 At the time of writing spiral phase plates have been demonstrated, but are
less reliable. Also real fork holographic masks are different from the chosen
Evaluating the transform Eq. (17) at positions q ¼ ln,m =rM , one can demonstration example inasmuch as they create higher order side bands.
5
One can compensate this sideways deviation for one of the side bands by
apply the orthogonality relations for Bessel functions. The result is imposing an initial phase 8 kx x onto the incident plane wave.
6
r 2 im X Real marks have sharp edges. These edges give rise to higher harmonics in
f~ m ðln,m =rM Þ ¼ M 2
cnm Jm þ 1 ðln,m Þ ð19Þ the Fourier transform, and those harmonics will create vortices with higher
4p n quantum number m. We will not deal with this case here.
P. Schattschneider, J. Verbeeck / Ultramicroscopy 111 (2011) 1461–1468 1465

with the wave front aberration w and the aperture defining


function P:
~ ðHÞ ¼ PðY=YM ÞeiwðHÞ :
W
Note that H is the variable after Fourier transform from the q
plane. The beam defining aperture appears under a half angle YM
(i.e. the half convergence angle seen from a point in the q plane).
Taking into account only defocus df and spherical aberration Cs
the phase imposed by the lens is radially symmetric, e.g. [13],
 
k C
wðYÞ ¼ z df Y2 þ s Y4 :
2 2

Note that the scattering angle H relates to the coordinate r in the


aperture as H ¼ r=f for a lens of focal length f (provided that
Y 5 1). Eq. (17) for the perfect lens is now replaced by
Z rM
im imj
Fig. 4. Transmission function Eq. (21) for creating vortex beams. The aperture has
c~ m ðqÞ ¼ e eiwðr=f Þ Jm ðqrÞr dr, ð24Þ
2p 0
5 mm diametre. Scale in mm.
pffiffiffiffiffiffiffiffi
which we evaluate at the Scherzer defocus df ¼  Cs l. The
azimuth angle j is now in the diffraction plane where the vortex
is focussed. In Figs. 7–9 are shown the radial traces through the
intensity patterns for different aperture/mask diameters. For an
ideal lens the profiles would be identical with the profile over the
right or left vortex in Fig. 5. The angle is measured with respect to
the center of the vortex (which is at 5 mrad in Fig. 5).
The maxima of the m 40 vortices move to higher radii with
increasing aberration. For the largest mask the change is well
visible; less for the 50 mm mask, and negligible for the small
Fig. 5. Diffraction pattern of the mask in Fig. 4 obtained by Fourier transform, mask. The side maxima in the profiles increase with m and with
showing the central Airy disk and two focussed vortices. The Bragg angle at 200 kV the aberration constant. Interestingly, the central zero for the
is 2:5 mrad. The phase is coded as hue. The helical structure of the vortices is well
visible (Color coding: Rainbow chart from 0 to 2p). (For interpretation of the
vortices with 9m9 40 remains always zero independent of the
references to color in this figure legend, the reader is referred to the web version of aberration. The vortex diameter scales with the reciprocal aper-
this article.) ture size. Even for the smallest mask it is in the range of mrad.
Such structures are difficult to observe.7

7. Incoherent illumination
Int. [arb.u.]

In reality, a perfectly coherent probe does not exist. Rather we


have illumination with extended sources, which reduces the
spatial (lateral) coherence of the beam. The grade of coherence
is given by the angle subtended by the source on the aperture, i.e.
the convergence angle. An exact treatment of partial coherence
would involve calculation of the density matrix of the electron.
Although only diagonal terms of this matrix are observable, the
−5 0 5
off-diagonal terms which are closely related to the mixed
angle [microrad]
dynamic form factor [14,15] may lead to visible effects when
Fig. 6. Horizontal trace through the diffraction pattern shows the volcano-like the electron interacts with matter [16–18]. The incoherent super-
profiles of the vortices. Scale in mrad. position of waves coming from different points in the source
reduces the size of the coherence patch on the aperture. This gives
rise to a broadening of the diffraction spots. The resulting vortex
6. Spherical aberration profiles are calculated from a convolution of Eq. (17) with the
incoherent intensity distribution of the source, projected on the
The focussed vortices of Fig. 5 with wave functions c ~ given by
diffraction plane. We assume here a Gaussian source shape
Eq. (17) are in the far field, or – equivalently – in the diffraction
2
=2s2
(back focal) plane of an ideal lens. Spherical aberration and rs ¼ eq :
defocus distort this ideal wave function via the wave transfer
With the charge density in the back focal plane from a perfectly
function W describing the wave front aberrations of the lens as
coherent (point) source
f~ ðqÞ ¼ c~ ðqÞ  WðqÞ:
r ¼ 9c~ m ðqÞ92 ,
This can be written as a product of the Fourier transformed
functions
7
Extremely large camera lengths are necessary. This can be achieved in the
~ ðHÞ,
fðHÞ ¼ cðHÞW low-angle diffraction mode.
1466 P. Schattschneider, J. Verbeeck / Ultramicroscopy 111 (2011) 1461–1468

1.0 0.30
m=0 m=1
0.8 0.25
0.20
0.6
0.15
0.4
0.10
0.2 0.05
0.0 0.00
0.00 0.05 0.10 0.15 0.20 0.00 0.05 0.10 0.15 0.20
0.20 m=2 0.15 m=3

0.15
0.10
0.10
0.05
0.05

0.00 0.00
0.00 0.05 0.10 0.15 0.20 0.00 0.05 0.10 0.15 0.20

Fig. 7. Radial intensity profiles in the diffraction plane at Scherzer defocus for different spherical aberration constants, 200 kV. Mask with 100 mm diameter. Abscissa:
scattering angle in mrad. Zero denotes the vortex centers, as in Fig. 3. Full lines for Cs ¼ 0, dashed lines Cs ¼ 0.5 mm, dash-dotted lines Cs ¼ 1.2 mm. The panels a–d are for
quantum numbers m ¼0, 1, 2, 3. The abscissa is relative to the vortex centres in mrad. The profiles are scaled to the maximum of the m ¼0 beam.

1.0
0.25
0.8
0.20
0.6 0.15
0.4 0.10
0.2 0.05
0.0 0.00
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.00 0.05 0.10 0.15 0.20 0.25 0.30

0.15 0.10
0.08
0.10
0.06

0.05 0.04
0.02
0.00 0.00
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.00 0.05 0.10 0.15 0.20 0.25 0.30

Fig. 8. Same as in Fig. 7 for a mask with 50 mm diameter. Full lines Cs ¼0 and dashed lines Cs ¼1.2 mm.

which is a function of radial distance q only, we calculate the with the modified Bessel function of first kind and order zero, I0.
convolution So, the convolution can be written as
Z 1
2 2 2 2
I ¼ rs  r ð25Þ IðqÞ ¼ eq =2s rðq0 Þeq0 =2s I0 ð2qq0 Þq0 dq0 :
0

as Fig. 10 shows the effect of partial coherence on the m ¼1


vortex. The central dip rises rapidly with increasing angular width
Z 1
2
=2s2 of the source subtended on the aperture. This constitutes a very
IðqÞ ¼ rðq0 Þeðqq0Þ d2 q0 :
0 sensitive monitor of coherence. On the other hand, it renders the
creation of pure vortices difficult. Very small sources are needed,
The Gaussian can be separated into the critical source size being inversely proportional to the mask
diameter. For a 50 mm mask, the angular width of the source
2
=2s2 2
=2s2 q02 =2s2 2qq0=2s2
eðqq0Þ ¼ eq e e : should not exceed 0:01 mrad in order to see a drop of the central
intensity to 50%.
Only the last factor depends on the azimuth, and the integral over
the j variable is 8. Conclusions
Z 2p
2
e2qq0 cos j=2s dj ¼ 2pI0 ð2qq0 Þ, We have developed the theory of electrons carrying quantized
0 orbital angular momentum. To make connection to realistic
P. Schattschneider, J. Verbeeck / Ultramicroscopy 111 (2011) 1461–1468 1467

1.0 0.25
0.8 0.20
0.6 0.15

0.4 0.10

0.2 0.05

0.0 0.00
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2
0.14 0.10
0.12 0.08
0.10
0.08 0.06
0.06 0.04
0.04
0.02
0.02
0.00 0.00
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2

Fig. 9. Same as in Fig. 7 for a mask with 10 mm diameter. The results for the 3 Cs values coincide within the drawing accuracy.

1.0 angle (i.e. a very high coherence) is necessary so as to keep the


volcano structure intact. Their small angular width in the far field
may allow the creation of nm-sized or smaller electron vortices
0.8 but the demand for extremely high coherence of the source poses
a serious difficulty.
0.6

Acknowledgements
0.4

P.S. acknowledges the support of the Austrian Science Fund,


0.2 Project I543-N20. J.V. acknowledges the financial support from the
European Union under the Framework 6 Program, under a contract
for an Integrated Infrastructure Initiative, Reference no. 026019
0.0
ESTEEM. We thank Karsten Held, Stefan Löffler and Inga Ennen for
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 valuable comments.
angle [microrad]

Fig. 10. Effect of a Gaussian source distribution on the m ¼1 vortex. Radial References
intensity profile for a mask with 10 mm diameter. All intensities scaled to 1. The
abscissa is relative to the vortex center. The parameter is s of the source
subtended on the aperture: s ¼ 0 (full line), 0.05 (dashed), 0.10 (dash-dotted), [1] M. Uchida, A. Tonomura, Generation of electron beams carrying orbital
angular momentum, Nature 464 (2010) 737–739.
0.15 (dotted) mrad.
[2] J. Verbeeck, H. Tian, P. Schattschneider, Production and application of
electron vortex beams, Nature 467 (2010) 301–304.
situations, we considered a plane wave moving along the optic [3] K.Y. Bliokh, Y.P. Bliokh, S. Savel’Ev, F. Nori, Semiclassical dynamics of electron
axis of a lens system, intercepted by a round, centered aperture.8 wave packet states with phase vortices, Physical Review Letters 99 (2007)
190404.
It turns out that the movement along the optic axis can be [4] P. Schattschneider, S. Rubino, C. Hébert, J. Rusz, J. Kunes, P. Novák, E. Carlino,
separated off; the reduced Schrödinger equation operating in M. Fabrizioli, G. Panaccione, G. Rossi, Detection of magnetic circular dichro-
the plane of the aperture can be mapped onto Bessel’s differential ism using a transmission electron microscope, Nature 441 (2006) 486–488.
[5] B. McMorran, A. Agrawal, I. Anderson, A. Herzing, H. Lezec, J. McClelland,
equation. The ensuing eigenfunctions fall into families with J. Unguris, Electron vortex beams with high quanta of orbital angular
discrete orbital angular momentum ‘ m along the optic axis momentum, Science 331 (2011) 192–195.
where m is a magnetic quantum number. Those vortices can be [6] G. Molina-Terriza, J.P. Torres, L. Torner, Twisted photons, Nature Physics 3
(2007) 305–310.
produced by matching a plane wave after passage through a [7] J. Nye, M. Berry, Dislocations in wave trains, Proceedings of the Royal Society
holographic mask with a fork dislocation to the eigenfunctions of of London, Series A 336 (1974) 165–190.
the cylindrical problem. Vortices can be focussed by magnetic [8] J. Vaughan, D. Willetts, Temporal and interference fringe analysis of TEM01
laser modes, Journal of the Optical Society of America 73 (1983) 1018–1021.
lenses into volcano-like charge distributions with very narrow [9] J. Poynting, The wave motion of a revolving shaft, and a suggestion as to the
angular divergence, resembling loop currents in the diffraction angular momentum in a beam of circularly polarized light, Proceedings of the
plane. Inclusion of spherical aberration changes the ringlike shape Royal Society of London, Series A 82 (1909) 560–567.
[10] L. Allen, M. Beijersbergen, R. Spreeuw, J. Woerdman, Orbital angular
but does not destroy the central zero intensity of vortices with
momentum of light and the transformation of Laguerre–Gaussian laser
m a0. Partial coherence of the incident wave leads to a rise of the modes, Physical Review A 45 (1992) 8185–8189.
central intensity minimum. It is shown that a very small source [11] S.J. van Enk, G. Nienhuis, Eigenfunction description of laser beams and orbital
angular momentum of light, Optics Communications 94 (1992) 147–158.
[12] W. Glaser, Grundlagen der Elektronenoptik, Springer-Verlag, Wien, 1952.
[13] P. Rez, Analytic properties of the contrast transfer function in high-resolution
8
In the experiment, this aperture carries the holographic mask. electron microscopy, Acta Crystallographica A 44 (1988) 946–953.
1468 P. Schattschneider, J. Verbeeck / Ultramicroscopy 111 (2011) 1461–1468

[14] P. Schattschneider, M. Nelhiebel, B. Jouffrey, The density matrix of [17] M. Nelhiebel, N. Luchier, P. Schorsch, P. Schattschneider, B. Jouffrey, The
inelastically scattered fast electrons, Physical Review B 59 (1999) mixed dynamic form factor for atomic core-level excitation in interfero-
10959–10969. metric electron-energy-loss experiments, Philosophical Magazine B 79
[15] C. Dwyer, S.D. Findlay, L.J. Allen, Multiple elastic scattering of core-loss (1999) 941–953.
electrons in atomic resolution imaging, Physical Review B 77 (2008) 184107. [18] M. Nelhiebel, P.-H. Louf, P. Schattschneider, P. Blaha, K. Schwarz, B. Jouffrey,
[16] P. Schattschneider, C. Hébert, B. Jouffrey, Orientation dependence of ioniza- Theory of orientation-sensitive near-edge fine-structure core-level spectro-
tion edges in EELS, Ultramicroscopy 86 (2001) 343–353. scopy, Physical Review B 59 (1999) 12807–12814.

Vous aimerez peut-être aussi