<<

arXiv:2009.13943v2 [quant-ph] 12 Jan 2021 Keywords: emotc aebe rsne xesvl,wt ealdrefere detailed with extensively, studied presented is been clas of operat have , aspects and optics historical design and beam the an practical, for Theoretical, like basis the devices. system, cal forms mo optical study electron this beam an and any of mechanics of trajectory the field optics netic beam electron classical In Introduction 1 Quan T Aberrations, approximation, propagator, quantum Paraxial Paraxial , series, electron Baker magnetic Round optics, Email E-mail 1 2 unu ehnc frudmgei lcrnlne with lenses electron magnetic round of mechanics Quantum orsodn author Corresponding eie aut TertclPhysics) (Theoretical Faculty Retired : : fqatmeet nteotc fpeetdyeeto emd beam electron th present-day of of optics purpose th the main optics on The negligible. beam effects electron quantum out. of of mechanics pointed t quantum is of the part beam, understand nonlinear nonparaxial the a in for and discussed aberrations, is in aberrations uncertainties of mechanics Quantum construct are also. equation, series differential the solving by obtained ouin fisprxa qaino oin ntecs fl of case k the well In the of motion. terms of in for equation model obtained law paraxial is its lens of model Gl solutions Glaser with the field lenses for magnetic electron tor axial magnetic the for Round models Foldy-Wouthuysen-lik considered. a using is equation nique Dirac the from rived [email protected] [email protected] clrter fqatmeeto emotc,a h single the at optics, beam electron quantum of theory Scalar otBxN.20,Psa oe 1,Sllh utnt fOm of Sultanate Salalah, 211, Code: Postal 2509, No. Box Post olg fAt n ple cecs(AS,Doa Univers Dhofar (CAAS), Sciences Applied and Arts of College t rs tet eta ntttso ehooy(I)Ca (CIT) Technology of Institutes Central Street, Cross 4th ia qain od-otusntasomto,Qatmelec Quantum transformation, Foldy-Wouthuysen equation, Dirac aenAmdKhan Ahmed Sameen lsradpwrlwmdl of models law power and Glaser B ( z h elkonfnaetlsltoso h aailequat paraxial the of solutions fundamental known well the ) a eateto ahmtc n Sciences, and Mathematics of Department b , h nttt fMteaia Sciences, Mathematical of Institute The https://orcid.org/0000-0003-2968-2044 hrmn,Ceni601,India 600113, Chennai Tharamani, , https://orcid.org/0000-0003-1264-2302 a, B Abstract 1 ( aawm Jagannathan Ramaswamy , z r tde.Prxa unu propaga- quantum Paraxial studied. are ) duigtePeano-Baker the using ed rey oeo quantum of Role briefly. eeutoso motion of equations he rnfraintech- transformation e ia ehnc felectron of mechanics sical ne ihtepower the with enses u uncertainties. tum B sradpwrlaw power and aser igi h electromag- the in ving uhteinfluence the ough onfundamental nown pril ee,de- level, -particle ( o fsc emopti- beam such of ion cst h literature, the to nces z vcsmgtbe might evices ase a,Peano- map, ransfer ) b, satcei to is article is 2 mpus, ity, sn classical using an rnbeam tron ions, in the first two volumes of the encyclopedic text book of Hawkes and Kasper [1,2] and the third volume [3] presents the wave mechanics of electron beam optics, or essentially, the quantum mechanics of electron optical imaging (for classical electron beam optics see also e.g., [4–9]). In studying the quantum mechanics of imaging in electron microscopy mostly the nonrelativistic Schr¨odinger equation is used, following the pioneering work of Glaser [10–12] (see [1–3] for details). In relativistic situations, the Schr¨odinger equa- tion with relativistically corrected mass, Klein-Gordon equation, or an approximate scalar wave equation derived from the are used (for details, see [3,13–17]). Driven by the curiosity to find how classical mechanics is so successful in the design and operation of optical devices like electron microscopes and particle accelerators when the microscopic charged particles passing through the systems should be obeying quantum mechanics, a systematic study of quantum mechanics of electron beam optics, at the single-particle level, derived from the Dirac equation using a Foldy-Woutuysen-like transformation technique was initiated in [18] and the first quantum mechanical deriva- tion of the classical Busch formula for the focal length of an axially symmetric magnetic 1 electron lens, or round , was obtained. Since the electron is a - 2 parti- cle, Dirac equation is the proper equation on which a general theory of quantum electron beam optics has to be based. Of course, approximations applicable to cases like non- relativistic situations and treatments ignoring spin can be derived from such a general theory. With this in mind, following [18], a formalism of quantum charged particle beam optics applicable to devices from low energy electron microscopes to high energy particle accelerators is being developed [19–31]. A consolidated account of the formalism of quantum mechanics of charged particle beam optics, at the single-particle level, has been presented in [32] where we have studied the general theory of round magnetic lenses based on relativistic quantum mechanics us- ing both the Klein-Gordan equation (ignoring spin of the particle) and the Dirac equation 1 (taking into account the spin- 2 nature of particles like ) without assuming any specific model for the magnetic field of the system. We have shown, in general, how the relativistic quantum theory of any charged particle beam optical system can be approx- imated leading to the nonrelativistic quantum theory of the system. Following [32], we have studied in [33] the quantum mechanics of bending of a nonrelativistic monoenergetic charged particle beam by a dipole magnet in the paraxial approximation. In this article we shall study round magnetic lenses with Glaser and power law models for the axial magnetic field on the basis of the scalar quantum theory derived from the Dirac equation using the Foldy-Wouthuysen-like transformation technique. Let us consider a round magnetic lens with the z-axis as its straight optic axis. The system is formed by the axially symmetric magnetic field obtained from the vector po- tential (see e.g., [1]) 1 1 A~ (~r⊥, z)= yΠ(~r⊥, z) , xΠ(~r⊥, z) , 0 , (1) −2 2 

2 with ∞ 2 n 1 r⊥ (2n) Π(~r⊥, z) = B (z) n!(n + 1)! − 4 ! nX=0 1 1 = B(z) B′′(z)r2 + B(4)(z)r4 ..., (2) − 8 ⊥ 192 ⊥ − 2 2 2 where ~r⊥ =(x, y), r⊥ = x + y , and dB(z) d2B(z) B(0)(z) = B(z), B′(z)= , B′′(z)= , dz dz2 d3B(z) d2nB(z) B′′′(z) = , ..., B(2n)(z)= . (3) dz3 dz2n The corresponding magnetic field B~ = ~ A~ has the components ∇ × ~ 1 ′ 1 ′′′ 2 B⊥ (~r⊥, z) = B (z) B (z)r⊥ + ... ~r⊥, −2  − 8  1 1 B (~r , z) = B(z) B′′(z)r2 + B(4)(z)r4 .... (4) z ⊥ − 4 ⊥ 64 ⊥ −

Thus, B(z)= Bz(0, 0, z) is the axial magnetic field of such systems. For the classic Glaser model lens B(z) is given by the Lorentzian curve B B(z)= 0 , (5) 1+(z/a)2 where B0 = B(0) is the value of the field at the maximum of the bell-shaped distribution and a is the half-width of the distribution. Besides the Glaser model lens, we shall study the power law model lenses (see [1, 2, 34–51]) for which

B(z) zn. (6) ∝ Optical properties of these lenses have been analysed, based on classical electron optics, very extensively using diverse analytical and numerical tools. The purpose of this arti- cle is mainly to understand the quantum mechanics of such lenses though the quantum corrections to the classical results might be negligible.

2 Quantum electron beam optical Hamiltonian of a round magnetic lens

Let us consider a monoenergetic, quasiparaxial i.e., almost paraxial, beam of electrons moving along the positive z-axis of a round lens comprising of a static magnetic field B~ (~r) given in (4). In an ideal paraxial beam any electron will have ~p⊥ 0. For such a paraxial beam we can take | | ≈ Π(~r , z) B(z) (7) ⊥ ≈ 3 in (1). In a quasiparaxial beam, slightly deviating from the paraxial condition, we could have ~p p . For such a quasiparaxial beam we can take | ⊥|≪ z 1 Π(~r , z) B(z) B′′(z)r2 . (8) ⊥ ≈ − 8 ⊥

In the paraxial case only the lowest order terms in ~r⊥ would contribute to the effective field felt by the beam electrons moving close to the optic axis. Let us denote the rest mass of the electron by m and its charge, e, by q. The Dirac equation governing the − quantum dynamics of the electron, ignoring its anomalous magnetic moment, is

Ψ1(~r, t) ∂Ψ (~r, t) Ψ (~r, t) ih¯ = Hˆ Ψ (~r, t) , Ψ (~r, t)=  2  , (9) ∂t Ψ3(~r, t)      Ψ4(~r, t)    where Ψ (~r, t) is the 4-component Dirac spinor wave function associated with the electron and Hˆ is the Dirac Hamiltonian given by

2 mc 0 cπˆz c (ˆπx iπˆy) 2 − ˆ  0 mc c (ˆπx + iπˆy) cπˆz  H = 2 − cπˆz c (ˆπx iπˆy) mc 0    − − 2   c (ˆπx + iπˆy) cπˆz 0 mc   − −  = mc2β + c~α ~π,ˆ (10) · with

~πˆ = ~pˆ qA~ = ih¯ ~ qA,~ (11) − − ∇ − 1 0 0 ~σ β = , ~α = , 0 1 ~σ 0 − ! ! 0 1 0 i 1 0 σx = , σy = − , σz = , (12) 1 0 i 0 0 1 ! ! − ! in which 1 0 0 0 1 = , 0 = . (13) 0 1 ! 0 0 ! The three Pauli matrices ~σ and the four Dirac matrices (~α, β) satisfy the relations,

σ σ = σ σ , for j = k, j,k = x,y,z, σ2 = 1, j = x, y, z, j k − k j 6 j α α = α α , for j = k, j,k = x,y,z, α2 = I, j = x, y, z, j k − k j 6 j βα = α β, j = x,y,z, β2 = I, (14) j − j where 1 0 I = . (15) 0 1 !

4 When an electron is associated with the 4-component wave function Ψ (~r, t) the probability for it to be found at the position ~r at time t is given by

4 † ∗ Ψ (~r, t) Ψ (~r, t)= Ψj (~r, t)Ψj (~r, t) . (16) jX=1 Hence Ψ (~r, t) is normalized as

d3r Ψ (~r, t)† Ψ (~r, t)=1, (17) Z so that the probability of finding the electron somewhere in space is one at any time 3 ∞ ∞ ∞ t, where d r stands for −∞ −∞ −∞ dxdydz. Since the Dirac Hamiltonian in (10) is HermitianR the time evolutionR ofR Ψ (R~r, t) according to (9) is unitary and the normalization (17) is preserved in time. We are interested in studying the evolution of the beam along the optic axis i.e., z- axis. Since the beam is monoenergetic and the field is time-independent, with ~r =(x, y, z) written as (~r⊥, z), we can take

ψ1(~r⊥, z) ψ2(~r⊥, z) Ψ(~r⊥,z,t) = exp ( iEt/h¯)ψ(~r⊥, z) = exp ( iEt/h¯)   , (18) − − ψ3(~r⊥, z)    ψ (~r , z)   4 ⊥    where E = m2c4 + c2 ~p 2 is the conserved total energy of the electron with ~p = p | 0| | 0| 0 as the designq momentum. Note that E is positive for us. Let us write ~p0 = (~p0⊥,p0z). Since the beam is quasiparaxial and is moving along the positive z-axis, p > 0, p p 0z 0z ≈ 0 and ~p0⊥ p0. With the choice of Ψ as in (18) the time-dependent Dirac equation (9) becomes| | ≪ Hψˆ = Eψ, (19) the time-independent Dirac equation. With the beam of electrons entering the system through one side and leaving it through the opposite side, we are dealing with the scatter- ing states of the system, not the bound stationary states. We have to study the z-evolution of the wave function of the beam propagating through the system in order to understand the optical properties of the system. To this end we proceed as follows. Multiplying both sides of (19) from left by αz/c and rearranging the terms we get the z-evolution equation for ψ given by

∂ψ(~r⊥, z) ih¯ = ˆψ(~r , z), ˆ = p βχα qA I + α ~α ~πˆ , (20) ∂z H ⊥ H − 0 z − z z ⊥ · ⊥ with 2 1 2 2 4 ξ1 0 E + mc p0 = √E m c , χ = −1 , ξ = , (21) c − 0 ξ 1 ! s E mc2 − −

5 where the square roots are taken to be positive. Now, let

′ 1 −1 1 ψ = Mψ, M = (I + χαz) , M = (I χαz) . (22) √2 √2 − Then, we get

∂ψ′ ih¯ = ˆ′ψ′, ˆ′ = M ˆM −1 = p β + ˆ + ˆ, ∂z H H H − 0 E O ~ ˆ 1 0 ˆ 0 ξ~α⊥ πˆ⊥ = qAz , = · . (23) E − 0 1 ! O ξ−1~α ~πˆ 0 ! − − ⊥ · ⊥ The Dirac Hamiltonian Hˆ in (10) is a sum of two parts, the diagonal part mc2β and the off-diagonal part c~α ~πˆ which contains 2 2 matrix blocks only along the off-diagonal and anticommutes with· β. If the system has× an electric field also the scalar potential will contribute another diagonal term to the Hamiltonian. In general, in the Dirac Hamiltonian the diagonal part which commutes with β and does not couple the upper and lower pairs of the four-component Dirac wave function is called even and the off-diagonal part which anticommutes with β and couples the upper and lower pairs of the Dirac wave function is called odd. This structure of Hˆ helps expand it as a series of terms, with 1/mc2 as the expansion parameter, that correspond to the nonrelativistic approximation and relativistic corrections of various orders. Actually, the dimensionless expansion parameter is cp/mcˆ 2, as in the expansion of the free particle energy

cp 2 E = m2c4 + c2p2 = mc2 1+ s mc2 q   2 4 2 4 2 1 cp 1 cp 2 p p = mc 1+ 2 2 + ... mc + 3 2 , (24) " 2 mc  − 8  mc  # ≈ 2m − 8m c but, for book keeping purposes one usually regards 1/mc2 as the expansion parameter. The desired expansion of Hˆ as a series of approximations, up to any desired order, is obtained via the Foldy-Wouthuysen transformation technique which consists of a series of transformations till the desired order of accuracy is reached (see e.g., [52]). A similar ′ structure is seen in ˆ (23) in which the diagonal part, p0β+ ˆ, commutes with β, and the off-diagonal part, ˆH, anticommutes with β. This helps− us expandE ˆ′ (or ˆ = M −1 ˆ′M) O H H H as a series of approximations, with 1/p0 as the expansion parameter, starting with the paraxial approximation and followed by nonparaxial approximations. Here again, the dimensionless expansion parameter isp ˆ⊥/p0, like in the expression

2 2 2 p⊥ pz = p0 p⊥ = p0v1 − u − p0 ! q u t2 4 2 4 1 p⊥ 1 p⊥ p⊥ p⊥ = p0 1 ... p0 3 , (25)  − 2 p0 ! − 8 p0 ! −  ≈ − 2p0 − 8p0   6 for a free particle, and one uses 1/p0 as the expansion parameter for book keeping pur- poses. To get the desired result we shall be using a Foldy-Wouthuysen-like (FW-like) transformation technique. The first FW-like transformation is

(1) ′ i ψ = exp iSˆ1 ψ , Sˆ1 = β ˆ. (26) 2p0 O   This leads to the z-evolution equation for ψ(1) given by

∂ψ(1) ih¯ = ˆ(1)ψ(1), (27) ∂z H where ˆ(1) = exp β ˆ/2p ˆ′ exp β ˆ/2p H − O 0 H O 0   ∂   ih¯ exp β ˆ/2p exp β ˆ/2p − − O 0 ∂z O 0      p β + ˆ(1) + ˆ(1), ≈ − 0 E O ˆ ˆ(1) ˆ 1 ˆ2 1 ˆ ˆ ˆ ∂ 1 ˆ4 = β 2 , , + ih¯ O 3 β , E E − 2p0 O − 8p "O O E ∂z !# − 8p O 0 h i 0 1 ∂ ˆ 1 ˆ(1) = β ˆ, ˆ + ih¯ O ˆ3, (28) O −2p O E ∂z ! − 3p2 O 0 h i 0 in which ˆ(1) is even, ˆ(1) is odd, and A,ˆ Bˆ = AˆBˆ BˆAˆ, the commutator of Aˆ and Bˆ. E O − In deriving ˆ(1) we have used the relationsh i H 1 exp ( Aˆ)B exp (Aˆ) = Bˆ A,ˆ Bˆ + A,ˆ A,ˆ Bˆ − − 2! 1h i h h ii A,ˆ A,ˆ A,ˆ Bˆ + ..., (29) −3! h h h iii ∂ ∂Aˆ(z) 1 ∂Aˆ(z) exp ( Aˆ(z)) exp (Aˆ(z)) = Aˆ(z), − ∂z ∂z − 2! " ∂z #   1 ∂Aˆ(z) + Aˆ(z), Aˆ(z), + ..., (30) 3! " " ∂z ##

3 and only terms of order up to 1/p0 have been retained (see [21, 32, 53] for more details). Now, if we make a second FW-like transformation following the same recipe as for the first one, i ψ(2) = exp iSˆ ψ(1), Sˆ = β ˆ(1), (31) 2 2 2p O   0 we get ∂ψ(2) ih¯ = ˆ(2)ψ(2), ˆ(2) p β + ˆ(2) + ˆ(2), ∂z H H ≈ − 0 E O 7 ˆ(1) ˆ(2) ˆ(1) 1 ˆ(1) 2 1 ˆ(1) ˆ(1) ˆ(1) ∂ = β 2 , , + ih¯ O E E − 2p0 O − 8p "O O E ∂z !#   0 h i 1 4 β ˆ(1) , −8p3 O 0   1 ∂ ˆ(1) 1 3 ˆ(2) = β ˆ(1), ˆ(1) + ih¯ O ˆ(1) , (32) O −2p O E ∂z ! − 3p2 O 0 h i 0   where the expressions for ˆ(1) and ˆ(1) are to be inserted from (28). We can continue this series of FW-like transformationsE O till we achieve the desired accuracy. To this end, ′ (1) (1) observe that ˆ in ˆ is of order zero in 1/p0, ˆ in ˆ is of first order in 1/p0, and (2) (2) O H O H ˆ in ˆ is of second order in 1/p0. This shows that with each transformation the oddO operatorH in the Hamiltonian becomes weaker and we can stop the transformation process when the odd operator becomes weak to any desired order. Stopping with the third FW-like transformation we get

∂ψ(3) ih¯ = ˆ(3)ψ(3), ˆ(3) p β + ˆ(3), (33) ∂z H H ≈ − 0 E where we have dropped the odd term ˆ(3) as negligible for our purpose. Further, note that for a quasiparaxial beam movingO in the positive z direction ψ, or ψ(3), will be like exp (ip z/h¯) such that ih∂ψ¯ (3)/∂z p ψ. This would require the lower pair of ∼ 0 ∼ − 0 components of ψ(3) to be very small compared to the upper pair of components. In view of this, ˆ(3) can be approximated further by dropping βs from it, or replacing βs in it by 2 2 identityH matrices (see [21, 32] for more details). × Let us now retrace the transformations we have made so that we go back to the original Dirac ψ in (20):

ψ = M −1 exp β ˆ/2p exp β ˆ(1)/2p exp β ˆ(2)/2p ψ(3). (34) O 0 O 0 O 0       This inverse transformation will not take us to the original Hamiltonian ˆ in (20) since H it has been expanded in a series and truncated. Note that the Hamiltonian ˆ is not Hermitian. We are considering the beam to be quasiparaxial i.e., slightly deviatingH from the paraxial condition, and hence we can choose A~ to be given by (1) and (8). Then, 3 carrying out the transformation (34), and keeping only terms of order up to 1/p0, and dropping the non-Hermitian terms, we get

∂ψ (~r⊥, z) ih¯ = ˆ ψ (~r , z) , (35) ∂z Ho ⊥ where

ˆ = ˆ + ˆ′ + ˆ(¯h) + ˆ(¯h), (36) Ho Ho,p Ho Ho Ho,σ ˆ 1 2 1 2 2 ˆ o,p = p0 + pˆ⊥ + p0α (z)r⊥ α(z)Lz, (37) H − 2p0 2 −

8 2 ˆ′ 1 4 α(z) 2 ˆ α (z) ~ ~ 2 o = 3 pˆ⊥ 2 pˆ⊥Lz ~r⊥ pˆ⊥ + pˆ⊥ ~r⊥ H 8p0 − 2p0 − 8p0 · · 2   3α (z) 2 2 2 2 1 ′′ 3 ˆ 2 + r⊥pˆ⊥ +ˆp⊥r⊥ + α (z) 4α (z) Lzr⊥ 8p0 8 − p     + 0 α4(z) α(z)α′′(z) r4 , (38) 8 − ⊥ 2   2 ˆ(¯h) h¯ ′ 2 ′′ 2 h¯ ′′ 2 ′ ′′′ 4 o α (z) 2α(z)α (z) r⊥ + α (z) α (z)α (z) r⊥, (39) H ≈ 8p0 − 32p0 − 2    ˆ(¯h) 3¯h ′ ′′ 2 o,σ α (z)α (z)r⊥ (yΣx xΣy)+ ..., (40) H ≈ 64p0 − with qB(z) ~σ 0 α(z)= , Σ=~ . (41) 2p0 0 ~σ !

Note that ˆo depends explicitly on z through α(z). It should be noted that the design momentumHp in ˆ can be nonrelativistic or relativistic. In the relativistic case, with 0 Ho γ = 1/ 1 v2/c2 and E = γmc2, p = √E2 m2c4/c = γmv . In the nonrelativistic − 0 0 − 0 case, withq γ [1+(v2/2c2)] and E mc2 +(mv2/2), p mv . Thus, there is no nonrel- ≈ 0 ≈ 0 0 ≈ 0 ativistic approximation for ˆ except for choosing p appropriately. The approximations Ho 0 for ˆo depend only on whether the beam is paraxial or nonparaxial. Equation (35) is theH quantum electron beam optical evolution equation with ˆ as the quantum electron Ho beam optical Hamiltonian. Now, we are interested in the z-evolution of ψ (~r⊥, z) according to (35). Note that the equations (16) and (18) imply that

4 † ∗ ψ (~r⊥, z) ψ (~r⊥, z)= ψj (~r⊥, z) ψj (~r⊥, z) , (42) jX=1 is the probability of finding the electron at the position (x, y) in the transverse plane at 2 † the point z on the optic axis at any time t and d r⊥ ψ (~r⊥, z) ψ (~r⊥, z) is the probability of finding the electron somewhere in the planeR perpendicular to the axis at the point 2 ∞ ∞ z. Here we are writing d r⊥ for −∞ −∞ dxdy. The probability of finding the electron somewhere in space (16)R at any timeR Rt is always one and is conserved since it cannot disappear from space. But, the probability of finding the electron somewhere in the plane perpendicular to the axis at a point z need not have the same value i.e., be conserved, along the z-axis. This explains why ˆ in (20) is non-Hermitian. However, if we consider the probability for a beam electron toH be lost during the transit through the system to be negligible we can assume that the effect of the non-Hermitian terms in the Hamiltonian to be negligible and drop them. This is what we have done in arriving at the quantum electron beam optical Hamiltonian ˆ in (36) which is Hermitian. With Hermitian ˆ Ho Ho the z-evolution of ψ (~r⊥, z) is unitary so that we can normalize the wave function ψ (~r⊥, z) as 2 † d r⊥ ψ (~r⊥, z) ψ (~r⊥, z)=1, at any z. (43) Z 9 This means that we are considering the probability of finding the electron somewhere in the transverse (x, y) plane at a point z on the axis is same for any z and we have normalized this value to be one. ˆ ˆ ˆ′ In o in (36), o,p in (37) is responsible for the paraxial behavior, and o in (38) H H ˆ(¯h) H gives rise to third-order aberrations. The part o in (39) leads to explicitlyh ¯-dependent H ˆ(¯h) modifications of the paraxial behavior and aberrations. The part o,σ in (40) is the matrix part originating from the spin of the electron, or the 4-componentH nature of the electron wave function, besides being explicitlyh ¯-dependent. Since, the terms in both ˆ(¯h) and ˆ(¯h) are proportional toh ¯ and powers ofh ¯ their contributions are tiny and Ho Ho,σ can be dropped from ˆo. If we drop theseh ¯-dependent terms the remaining part of ˆ , namely ˆ + ˆ′H, is the ‘classical’ scalar term which acts on each component of Ho Ho,p Ho ψ (~r⊥, z) independently.  This implies that we can take, effectively, the electron wave function to be a scalar i.e., a single-component wave function like the nonrelativistic Schr¨odinger wave function, ignoring the electron spin or treating it as a spectator. Thus, we shall take ∂ψ (~r , z) ih¯ ⊥ = ˆ ψ (~r , z) , ˆ = ˆ + ˆ′ , (44) ∂z Ho,c ⊥ Ho,c Ho,p Ho as the quantum electron beam optical evolution equation with ψ (~r⊥, z) as the scalar wave function representing the beam electron and ˆ as the quantum electron beam optical Ho,c Hamiltonian. One may wonder whether it was necessary at all to start with the Dirac equation to obtain the scalar wave function description and whether it would not have been enough to start with the scalar relativistic Klein-Gordon equation. It is true that if we are to drop all theh ¯-dependent scalar and matrix terms from the quantum electron beam optical Hamiltonian there would be no difference whether we start with the Dirac equation or the Klein-Gordon equation. However, if we are to include theh ¯-dependent scalar terms then ˆ(¯h) in (39) is found to be different from the corresponding term obtained when one Ho starts with the Klein-Gordon equation (compare (39) with Eq.(5.213) in [32]). Thus, the signature of electron spin is seen even in theh ¯-dependent scalar part besides, obviously, in the matrix, or spin, part of the quantum electron beam optical Hamiltonian (see [21, 32] for more details). Foldy-Wouthuysen-like transformations are useful also in series expansions of the Klein-Gordon equation and the Helmholtz equation (see e.g., [21,23,32,52,54]; see also [55] and references therein). The formalism of quantum charged particle beam optics [32] has had a decisive influence on the development of new formalisms of scalar light beam optics based on the Helmholtz equation and vector light beam optics, incorporating polariza- tion, based on the Maxwell equations [56–62]. The use of quantum methodologies in light beam optics has provided new insights on Hamilton’s optico-mechanical analogy and also extends it into the -dependent regime [63].

10 3 Equations of motion for the quantum averages of observables

Let us look at the evolution of the observables of the beam electron defined in the plane perpendicular to the optical axis as it transits through the system. We can take over the usual quantum mechanical formalism for time evolution by replacing time t by z. In general, for an observable, say O (~r⊥, ~p⊥, z), represented by the corresponding Hermitian ˆ ~ operator O ~r⊥, pˆ⊥, z ,   Oˆ (z)= d2r ψ∗ (~r , z) Oˆ ~r ,~pˆ , z ψ (~r , z)= ψ(z) Oˆ ψ(z) (45) h i ⊥ ⊥ ⊥ ⊥ ⊥ h | | i Z   gives its expectation value, or average, in the transverse plane at z. Since Oˆ is Hermitian Oˆ (z) is real. We can take Oˆ (z) as corresponding to the value of the classical variable hO iin the transverse plane ath zi, following the Ehrenfest theorem (see e.g., [64, 65], and also [21,32]). In particular, we are interested in the z-evolution of ~r (z) and ~pˆ (z)/p , h ⊥i h ⊥i 0 representing the classical ray coordinates ~r⊥(z) and d~r⊥(z)/dz, respectively. To this end, we proceed as follows, borrowing the well known techniques of the study of time-evolution of quantum systems. We can write, as the formal solution of (44),

ψ (~r⊥, z)= Uˆ (z, zi) ψ (~r⊥, zi) , (46) where zi and z are the initial and final points of observation in the optic axis and Uˆ (z, zi) is the z-evolution operator. Substituting this expression for ψ (~r⊥, z) in (44) it is seen that ∂Uˆ (z, z ) ih¯ i = ˆ (z)Uˆ (z, z ) , Uˆ (z , z )= I,ˆ (47) ∂z Ho,c i i i where Iˆ is the identity operator. Taking the Hermitian adjoint on both sides of (47) we have ∂Uˆ † (z, z ) ih¯ i = Uˆ † (z, z ) ˆ (z), Uˆ † (z , z )= I,ˆ (48) ∂z − i Ho,c i i where we have used the fact that ˆ (z) is Hermitian. Then, multiplying both sides of (47) Ho,c from left by Uˆ † and both sides of (48) from right by Uˆ and adding the two equations we get † † † ih∂¯ (Uˆ Uˆ)/∂z = 0, showing that Uˆ Uˆ is independent of z. Since Uˆ (zi, zi)Uˆ(zi, zi)= I, it † † † follows that Uˆ Uˆ = I i.e., Uˆ(z, zi) is unitary. From the equation Uˆ(Uˆ Uˆ)=(UˆUˆ )Uˆ = Uˆ we can conclude that UˆUˆ † = I. From (47), we have

z z ∂Uˆ (z, zi) i dz = dz1 ˆo,c (z1) Uˆ (z1, zi) , (49) Zzi ∂z −h¯ Zzi H or z i z Uˆ (z, zi) = Uˆ (z, zi) I = dz1 ˆo,c (z1) Uˆ (z1, zi) . (50) zi − −h¯ zi H Z

11 This leads to the formal solution i z Uˆ (z, zi)= I dz1 ˆo,c (z1) Uˆ (z1, zi) . (51) − h¯ Zzi H Iterating this formal solution we get i z Uˆ (z, zi) = I dz1 ˆo,c (z1) − h¯ Zzi H 2 i z z2 + dz2 dz1 ˆo,c (z2) ˆo,c (z1) −h¯  Zzi Zzi H H 3 i z z3 z2 + dz3 dz2 dz1 ˆo,c (z3) ˆo,c (z2) ˆo,c (z1) −h¯  Zzi Zzi Zzi H H H + .... (52)

Note that Uˆ (z, zi) becomes exp ( i(z zi) ˆo,c/h¯) if ˆo,c is independent of z. Let us write, in general, − − H H i z Uˆ (z, zi)= P exp dz ˆo,c(z) , (53)  −h¯ Zzi H  where P[ ], say, the z-ordered exponential, represents symbolically the infinite series in ··· the right hand side of (52). An equivalent expression for Uˆ (z, zi) is given by the Magnus formula [66] (see also [21, 32, 67]): i Uˆ (z, z ) = exp Tˆ (z, z ) , i −h¯ i z   Tˆ (z, zi) = dz1 ˆo,c (z1) Zzi H 1 i z z2 + dz2 dz1 ˆo,c (z2) , ˆo,c (z1) 2 −h¯  Zzi Zzi H H 2 h i 1 i z z3 z2 + dz3 dz2 dz1 ˆo,c (z3) , ˆo,c (z2) , ˆo,c (z1) 6 −h¯ zi zi zi H H H   Z Z Z nhh i i + ˆ (z ) , ˆ (z ) , ˆ (z ) Ho,c 1 Ho,c 2 Ho,c 3 + .... hh i io (54) Note that Tˆ(z, z ) is Hermitian when ˆ is Hermitian such that Uˆ(z, z ) is unitary. i Ho,c i Now, we can write the average value Oˆ (z) for any Oˆ as h i ∗ Oˆ (z) = d2r Uˆ (z, z ) ψ (~r , z ) OˆUˆ (z, z ) ψ (~r , z ) h i ⊥ i ⊥ i i ⊥ i Z   2 ∗ † = d r⊥ ψ (~r⊥, zi) Uˆ (z, zi) OˆUˆ (z, zi) ψ (~r⊥, zi) Z = ψ (z ) Uˆ † (z, z ) OˆUˆ (z, z ) ψ (z ) . (55) h i | i i | i i For any observable the equation of motion along the z-axis is given by

† d Oˆ (z) ∂Uˆ (z, zi) h i = ψ (zi) OˆUˆ (z, zi) ψ (zi) dz * ∂z +

12 ˆ † ∂U (z, zi) + ψ (zi) Uˆ (z, zi) Oˆ ψ (zi) * ∂z +

ˆ † ∂O + ψ (zi) Uˆ (z, zi) Uˆ (z, zi) ψ (zi) * ∂z +

i † = ψ (z ) Uˆ (z, z ) ˆ , Oˆ Uˆ ( z, z ) ψ (z ) h¯ i i Ho,c i i D h i E ˆ † ∂O + ψ (zi) Uˆ (z, zi) Uˆ (z, zi) ψ (zi) * ∂z +

i ∂Oˆ = ˆo,c, O ˆ (z)+ (z), (56) h¯ H * ∂z + Dh iE where we have used (47), (48) and (55). The equation (44) is the quantum electron beam optical equation of motion for the wave function ψ (~r⊥, z) in the Schr¨odinger picture in which the wave function changes with z and an observable of the electron does not change with z in the absence of any explicit z-dependence. In the Heisenberg picture the wave function remains fixed, ψ (~r⊥, zi), and an observable of the electron changes with z as

† Oˆ(z)= Uˆ (z, zi) OˆUˆ (z, zi) , (57) so that the average value of the observable is given by

Oˆ (z)= ψ (z ) Oˆ(z) ψ (z ) . (58) h i i i D E

For the Heisenberg observable Oˆ(z) we get the equation of motion from (56) as

dOˆ(z) i ∂Oˆ 1 ∂Oˆ = ˆo,c(z), Oˆ(z) + (z)= Oˆ(z), ˆo,c(z) + (z). (59) dz h¯ H ∂z ! ih¯ H ∂z ! h i h i This is seen to correspond to the Hamilton equation of motion for the classical observables dO ∂O = O, + , (60) dz { Ho,c} ∂z where , is the classical Poisson bracket, is the classical electron beam optical { } Ho,c Hamiltonian which is obtained by replacing in ˆ the quantum operators ~r and ~pˆ by Ho,c ⊥ ⊥ the classical variables ~r⊥ and ~p⊥, respectively, and the quantum operator Oˆ is replaced by the classical observable O. Dirac’s ‘quantum classical’ correspondence rule ←→ 1 A,ˆ Bˆ A, B , (61) ih¯ ←→ { } h i takes (59) to (60). As suggested by this, we can obtain the ‘classical part’ of the quantum electron beam optical Hamiltonian ˆ i.e., ˆ = ˆ + ˆ′ , without the terms dependent Ho Ho,c Ho,p Ho onh ¯ and spin matrices, starting with the classical electron beam optical Hamiltonian and

13 replacing the classical variables in it by the corresponding Hermitian quantum operators and proper symmetrization to make the resulting Hamiltonian operator Hermitian. When the observable is not explicitly dependent on z the equation of motion (56) becomes d Oˆ (z) i h i = ˆ , Oˆ (z). (62) dz h¯ Ho,c Dh iE We are particularly interested in the equations of motion for x, y,p ˆx/p0 andp ˆy/p0 which are independent of z. Let us write x (z) = x , pˆ (z) = pˆ , etc., for the sake of h i h i h xi h xi simplicity of notation. The desired equations of motion are as follows:

d x 1 1 2 α(z) 2 ˆ h i = pˆx + α(z) y + 3 pˆ⊥pˆx + 2 pˆ⊥y 2 pˆxLz dz p0 h i h i 2p0 2p0 − 2 D E D E D E α (z) ~ ~ ~ ~ x~r⊥ pˆ⊥ ~r⊥ pˆ⊥x + xpˆ⊥ ~r⊥ + pˆ⊥ ~r⊥x − 4p0 · · · · D ED E D E D E 3α2(z) 1 + r2 p + p r2 α′′(z) 4α3(z) yr2 , (63) 4p ⊥ x x ⊥ − 8 − ⊥ 0 D E D E  D E d y 1 1 2 α(z) 2 ˆ h i = pˆy α(z) x + 3 pˆ⊥pˆy 2 pˆ⊥x +2 pˆyLz dz p0 h i − h i 2p0 − 2p0 2 D E D E D E α (z) ~ ~ ~ ~ y~r⊥ pˆ⊥ + ~r⊥ pˆ⊥y + ypˆ⊥ ~r⊥ + pˆ⊥ ~r⊥y − 4p0 · · · · 2 D E D E D E D E 3α (z) 2 2 1 ′′ 3 2 + r⊥py + pyr⊥ + α (z) 4α (z) xr⊥ , (64) 4p0 8 − D E D E  D E 1 d pˆ α(z) α(z) h xi = α2(z) x + pˆ + pˆ2 pˆ p dz − h i p h yi 2p3 ⊥ y 0 0 0 D E α2(z) + pˆ ~r ~pˆ + ~r ~pˆ pˆ + pˆ ~pˆ ~r + ~pˆ ~r pˆ 4p2 x ⊥ · ⊥ ⊥ · ⊥ x x ⊥ · ⊥ ⊥ · ⊥ x 0 D E D E D E D E 3α2(z) xpˆ2 + pˆ2 x − 4p2 ⊥ ⊥ 0 D E D E 1 α′′(z) 4α3(z) 2 Lˆ x + pˆ r2 −8p − z y ⊥ 0   D E D E 1 α4(z) α(z)α′′(z) r2 x , (65) −2 − ⊥ 1 d pˆ  α(z) Dα(z)E h yi = α2(z) y pˆ pˆ2 pˆ p dz − h i − p h xi − 2p3 ⊥ x 0 0 0 D E α2(z) + pˆ ~r ~pˆ + ~r ~pˆ pˆ + pˆ ~pˆ ~r + ~pˆ ~r pˆ 4p2 y ⊥ · ⊥ ⊥ · ⊥ y y ⊥ · ⊥ ⊥ · ⊥ y 0 D E D E D E D E 3α2(z) ypˆ2 + pˆ2 y − 4p2 ⊥ ⊥ 0 D E D E 1 ′′ 3 ˆ 2 α (z) 4α (z) 2 Lzy pˆxr⊥ −8p0 − −   D E D E

14 1 α4(z) α(z)α′′(z) r2 y . (66) −2 − ⊥  D E Note that all the terms on the right hand sides of the above equations are real, as should 2 ˆ be. A term like ( pˆ⊥y 2 pˆxLz ) may not look explicitly real, but can be seen to be real h i − h i2 ˆ by checking that the operator (ˆp⊥y 2ˆpxLz) is Hermitian. The following should be noted. Even− after we have dropped the explicitlyh ¯-dependent terms from the Hamiltonian ˆ the effect of quantum mechanics on the equations of Ho motion can be seen in the terms not linear in x , y , pˆx , and pˆy . For example, let us consider the term x3 (part of the term r2 xh =i hx3i +h yi2x ) andh i express it in terms of h i h ⊥ i h i h i x which would correspond to the result of a position measurement. Let δx = x x . hThen,i δx = 0. We can write −h i h i x3 = ( x + δx)3 = x 3 +3 x 2δx +3 x (δx)2 +(δx)3 h i h i h i h i D E D E D E = x 3 +3 x (δx)2 + (δx)3 . (67) h i h i D E D E This shows that the quantum electron beam optical equations of motion differ from the corresponding classical electron beam optical equations of motion since a term like x3 h i cannot be replaced by x 3. The difference is seen to depend on terms like (∆x)2 = (δx)2 , the square of the uncertaintyh i in x, and higher order central moments like (δx)3 h. Thus,i h i in principle, the quantum uncertainties will influence the equations of motion for the position and momentum of the beam electron and hence influence the performance of the system. For example, when an aperture is introduced in the path of the beam to limit the transverse momentum spread one will be introducing uncertainties in the position coordinates, ∆x and ∆y, and hence the corresponding momentum uncertainties, ∆p x ≥ h/¯ 2∆x and ∆py h/¯ 2∆y, in accordance with the Heisenberg uncertainty principle. This is uncontrollable.≥ However, in practice, these quantum effects might be too tiny to be of any significance in the present-day electron optical systems. This explains the remarkable success of classical electron optics.

4 Paraxial approximation

In the equations of motion (63-66) the terms on the right hand side linear in x , y , pˆ , and pˆ , are the result of ˆ , the paraxial part of ˆ , which is quadratich i inh~ri h xi h yi Ho,p Ho,c ⊥ and ~pˆ . The terms not linear in x , y , pˆ , and pˆ , are the result of ˆ′ . Now, let ⊥ h i h i h xi h yi Ho us keep only the terms linear in x , y , pˆx , and pˆy , on the right hand side of the equations of motion. In other words,h i leth i ush takei h i

∂ψ (~r , z) ih¯ ⊥ = ˆ ψ (~r , z) , (68) ∂z Ho,p ⊥ as the quantum paraxial electron beam optical evolution equation, with the quantum paraxial electron beam optical Hamiltonian ˆ given by (37). This paraxial, or linear, Ho,p

15 approximation leads to the paraxial equations of motion,

x (z) 0 α(z) 1 0 x (z) h i h i d y (z) y (z)    α(z)0 0 1    1h i = − 2 1h i , (69) pˆx (z) α (z)0 0 α(z) pˆx (z) dz  p0   p0   1 h i   − 2   1 h i   pˆy (z)   0 α (z) α(z) 0   pˆy (z)   p0 h i   − −   p0 h i        following from (63-66). Let us write this equation as

d ~r⊥ (z) ρ(z) 1 ~r⊥ (z) 1h ~i = 2 1h ~i , dz pˆ⊥ (z) ! α (z)1 ρ(z) ! pˆ⊥ (z) ! p0 − p0 D E D 0E α(z) with ρ(z)= . (70) α(z) 0 − ! Let us now introduce the Larmor XY z-coordinate frame with its X and Y axes rotating along the z-axis, coinciding with the z-axis of the laboratory frame, at the z-dependent z rate dθ(z)/dz = α(z) such that we can write, with θ (z, zi)= zi dzα(z), R x cos θ (z, z ) sin θ (z, z ) X X = i i = R (z, z ) . (71) y sin θ (z, z ) cos θ (z, z ) Y i Y ! − i i ! ! ! Then,

pˆx ih∂/∂x¯ ih∂/∂X¯ PˆX = − = R (z, zi) − = R (z, zi) . (72) pˆy ih∂/∂y¯ ih∂/∂Y¯ Pˆ ! − ! − ! Y !

Let us recall that zi and z are the initial and final points of observation in the optic axis. Note that α(z)= qB(z)/2p0 = qB(z)/2γmv0 = ωL(z)/v0 where ωL(z) is the instantaneous Larmor frequency at z. Now, we have

R~ z ~r⊥ (z) R (z, z ) 0 ⊥ ( ) h i i 1 ~ = R  1D ~ E  . (73) pˆ⊥ (z) ! 0 (z, zi) ! Pˆ⊥ (z) p0 p0     D E   This implies that, with dR (z, zi) /dz = ρ(z)R (z, zi),

R~ (z) d ~r⊥ (z) ρ(z)R (z, z ) 0 ⊥ h i i 1 ~ = R  1D ~ E  dz pˆ⊥ (z) ! 0 ρ(z) (z, zi) ! Pˆ⊥ (z) p0 p0     D E   R~ ⊥ (z) R (z, zi) 0 d + R  1D ~ E  . 0 (z, zi) ! dz Pˆ⊥ (z) p0       (74)

16 From (70) and (73) we have

d ~r⊥ (z) ρ(z) 1 R (z, zi) 0 1h ~i = 2 R dz pˆ⊥ (z) ! α (z)1 ρ(z) ! 0 (z, zi) ! p0 − D E R~ ⊥ (z)  1D ~ E  . (75) × Pˆ⊥ (z) p0       Now, equating the right hand sides of (74) and (75) we get ~ d R⊥ (z) R (z, z )−1 0 0 1 = i  1D ~ E  R −1 2 dz Pˆ⊥ (z) 0 (z, zi) ! α (z)1 0 ! p0 −       R~ ⊥ (z) R (z, zi) 0 R  1D ~ E  × 0 (z, zi) ! Pˆ⊥ (z) p0     ~   0 1 R⊥ (z) = 2  1D ~ E  . (76) α (z) 0 ! Pˆ⊥ (z) − p0       Thus, in the rotating XY z-coordinate system,

d R~ (z) ⊥ 1 ~ˆ D E = P ⊥ (z) (77) dz p0   and ~ˆ 1 d P ⊥ (z)   = α2(z) R~ (z), (78) p dz − ⊥ 0 D E leading to 2 d R~ ⊥ (z) + α2(z) R~ (z)=0, (79) D dz2E ⊥ D E the well known classical paraxial equations of motion for the position coordinates of the electron in the rotating coordinate system. Note that in the nonrelativistic case we can 2 take p0 = 2meU where U is the electric potential which has accelerated the electron. Then we have α2(z)= eB2(z)/8mU. ~ The transfer map for the averages of R~ ⊥ and Pˆ⊥ between the transverse planes D E   at zi and z can be obtained by integrating (76). We can write the result as

R~ ⊥ (z) R~ ⊥ (zi) h i g (z, zi) h (z, zi) h i 1 ~ = ′ ′ 1 ~  Pˆ⊥ (z)  g (z, zi) h (z, zi)  Pˆ⊥ (z )  p0 ! p0 i         ~r (z ) M ⊥ i = (z, zi) 1h i , (80) ~rˆ⊥ (z ) p0 i ! D E 17 observing that the Larmor XY z frame and the laboratory xyz frame coincide at z = zi. Note that the elements of the second row of M (z, zi) are the z-derivatives of the elements of the first row because of (77). We can recognize g (z, zi) and h (z, zi) as the two linearly independent solutions of the paraxial equation (79) corresponding to the initial conditions x (zi) = 1, pˆx (zi) /p0 = d x (z)/dz z=zi = 0 and x (zi) = 0, pˆ (z ) /p = d x (z)/dzh i = 1, respectively,h i andh thei same| conditions for yh also.i Then, h xi i 0 h i |z=zi the two solutions have to satisfy the initial conditions ′ ′ g (zi, zi)=1, g (zi, zi)=0, h (zi, zi)=0, h (zi, zi)=1. (81) From (73) and (80) we get

~r⊥ (z) g (z, z ) R (z, z ) h (z, z ) R (z, z ) ~r⊥ (zi) h i = i i i i h i . (82) 1 ~pˆ (z) g′ (z, z ) R (z, z ) h′ (z, z ) R (z, z ) 1 ~pˆ (z ) p0 ⊥ ! i i i i ! p0 ⊥ i ! D E D E Thus, in terms of the two fundamental solutions g (z, zi) and h (z, zi) we can write R 1 R ~ ~r⊥ (z)= g (z, zi) (z, zi) ~r⊥ (zi)+ h (z, zi) (z, zi) pˆ⊥ (zi) , (83) h i h i p0 D E to represent the ‘classical’ trajectory of the quantum average of position of the electron entering the system with ~r (z ) and ~pˆ (z ) as the average position and momentum h ⊥i i ⊥ i in the transverse plane of entry at zi. D E There are two ways of finding the fundamental solutions. One is to solve the paraxial equation (79), a second order linear ordinary differential equation, analytically. We will get two linearly independent solutions for the differential equation which can be combined linearly to get the two fundamental solutions with the required initial conditions. An alternative method is as follows. From (44), (46), (52), and (76), we have z M (z, zi) = P exp dzµ(z)  Zzi  z z z2 = I + dz1 µ (z1)+ dz2 dz1 µ (z2) µ (z1) Zzi Zzi Zzi z z3 z2 + dz3 dz2 dz1 µ (z3) µ (z2) µ (z1) Zzi Zzi Zzi + ..., 0 1 with µ(z)= . (84) α2(z) 0 − ! In the theory of linear ordinary differential equations this method of solving a system of linear differential equations is known as the Peano-Baker method and the above series expression for the transfer matrix M (z, zi) is known as the Peano-Baker series (see e.g., [68, 69]). Explicitly calculating the elements of the first row of M (z, zi) we get

z z2 2 g (z, zi) = 1 dz2 dz1 α (z1) − Zzi Zzi z z4 z3 z2 2 2 + dz4 dz3 α (z3) dz2 dz1 α (z1) Zzi Zzi Zzi Zzi ..., (85) − 18 and

z z2 2 h (z, zi) = (z zi) dz2 dz1 α (z1)(z1 zi) − − Zzi Zzi − z z4 z3 z2 2 2 + dz4 dz3 α (z3) dz2 dz1 α (z1)(z1 zi) Zzi Zzi Zzi Zzi − .... (86) − Note that the solutions defined by (85) and (86) satisfy automatically the required initial conditions (81).

5 Paraxial quantum propagator

The z-evolution of the wave functionψ(~r⊥, z) from the vertical plane at zi on the axis to the plane at z is given by (46). This equation can be written in integral form as

2 ψ (~r⊥, z)= d ri⊥ K (~r⊥, z; ~r⊥i, zi) ψ (~r⊥i, zi) , (87) Z 2 ∞ ∞ where d r⊥i stands for −∞ −∞ dxidyi, and R R R K (~r⊥, z; ~r⊥i, zi)= ~r⊥ Uˆ (z, zi) ~r⊥i , (88) D E is the quantum electron beam optical propagator given by the matrix element of the z- evolution operator Uˆ (z, zi) in the position representation. When the beam is paraxial the z-evolution of ψ(~r⊥, z) is given by (68). Then, the corresponding paraxial quantum electron beam optical propagator becomes

Kp (~r⊥, z; ~r⊥i, zi) = ~r⊥ Uˆp (z, zi) ~r⊥i D E z i = ~r⊥ P exp dz ˆo,p(z) ~r⊥i . (89)   −h¯ Zzi H  

The paraxial quantum electron beam optical Hamiltonian ˆo,p being quadratic in ~ H ~r⊥, pˆ⊥ , Kp (~r⊥, z; ~r⊥i, zi) can be calculated exactly (see [21, 23, 32, 70] for details). The result is p exp i p (z z ) 0 ¯h 0 − i Kp (~r⊥, z; ~r⊥i, zi) =   i2πhh¯ (z, zi) ip0 ′ 2 ~ 2 exp h (z, zi)r⊥ 2R⊥ ~r⊥i + g(z, zi)r⊥i , × " 2¯hh(z, zi) − · #   if h(z, zi) =0, (90) ′ 6 1 i g (z, zi) 2 Kp (~r⊥, z; ~r⊥i, zi) = exp p0 (z zi)+ r⊥ g(z, zi) "h¯ − 2g(z, zi) !# ~ 2 R⊥ δ ~r⊥i , if h(z, zi)=0, (91) × g(z, zi) −    19 where R~ ⊥ =(X,Y ) are the coordinates in the rotating coordinate system given by (71). Thus, for a paraxial beam z-evolution of the wave function is given by

′ p0 ip0 h (z, zi) 2 ψ (~r⊥, z) = exp (z zi)+ r⊥ i2πhh¯ (z, zi) " h¯ − 2h(z, zi) !#

2 ip0 2 ~ d r⊥i exp g(z, zi)r⊥i 2R⊥ ~r⊥i ψ (~r⊥i, zi) , × "2¯hh(z, z ) − · # Z i   if h(z, z ) =0 (92) i 6 and ′ ~ 1 i g (z, zi) 2 R⊥ ψ (~r⊥, z) = exp p0 (z zi)+ r⊥ ψ , zi , g(z, zi) "h¯ − 2g(z, zi) !# g(z, zi)    if h(z, zi)=0. (93) As is well known, equations (92) and (93) represent the general law of propagation of a paraxial wave function in the case of a round magnetic lens [3, 10–12] and form the basis for the development of Fourier transform methods in the electron optical imaging techniques (see [3] for details). Note that, as follows from (80), if h (z, zi) = 0 for a particular z then R (z) is the image point of R (z ) and the situation corresponds to h i h i i the stigmatic, or point-to-point, imaging with Larmor rotation. We shall not go further into the well known theory of imaging by electron lenses (see [3, 21, 32] and references therein for more details).

6 Glaser model round magnetic lens

For the round magnetic lens with Glaser’s bell-shaped model for B(z) given by (5) we have α0 qB0 α(z)= 2 , α0 = . (94) 1+(z/a) 2p0 The corresponding paraxial equation is d2 R (z) α2 h i + 0 R (z)=0, (95) dz2 (1+(z/a)2)2 h i in which R denotes X or Y in the rotating Larmor frame of reference. Following Glaser (see e.g., [2]), let us take au(ϕ) z = a cot ϕ, R (ϕ)= . (96) h i sin ϕ When z varies from via 0 to , ϕ varies from π via π/2 to 0. Then, the equation (95) becomes −∞ ∞ d2u(ϕ) + ω2u(ϕ)=0, with ω = 1+ α2a2. (97) dϕ2 0 q 20 Note that ω is a constant and hence the general solution of (97) is given by

u(ϕ)= C sin[ω(ϕ γ)], (98) − where C and γ are arbitrary constants of integration which have to be fixed to satisfy the initial conditions. Thus, the general solution of the paraxial equation (95) becomes

Ca sin[ω(ϕ γ)] R (ϕ)= − , (99) h i sin ϕ in terms of ϕ. Let the electron move from an initial point in the plane at z = zi to the right to the plane at z > zi. As is well known (see e.g., [71]), the two fundamental solutions of the paraxial equation (95), in terms of ϕ and ϕi, are given by 1 g (ϕ,ϕ ) = (ω sin ϕ cos [ω (ϕ ϕ )] + cos ϕ sin [ω (ϕ ϕ )]) i ω sin ϕ i − i i − i a sin [ω (ϕ ϕi)] h (ϕ,ϕi) = − . (100) − ω sin ϕi sin ϕ

−1 −1 Substituting ϕ = cot (z/a) and ϕi = cot (zi/a) we get g(z, zi) and h(z, zi) satisfying ′ ′ the required initial conditions: g(zi, zi) = 1, g (zi, zi) = 0, h(zi, zi) = 0, h (zi, zi) = 1. Then, the ‘classical’ trajectory of ~r (z), in the laboratory frame of reference is given by h ⊥i (83). Properties of the Glaser model magnetic lens have been well studied since the early days of electron microscopy. We just wanted to show how the classical paraxial theory of such a lens follows from quantum mechanics by identifying the classical ray coordinates with the quantum averages of position and momentum and using the paraxial approxi- mation at the quantum level. The quantum paraxial electron beam optical propagator for the Glaser model lens is obtained by substituting the two fundamental solutions (100) in (92) assuming that h(z, z ) = 0 for the plane of observation at z. The resulting exact i 6 propagator, under the paraxial approximation, could be useful in the study of propaga- tion of electron vortex beams in a round magnetic lens (see e.g., [72–74], and references therein, for recent work on related topics).

7 Round magnetic lenses with B(z) zn ∝ Let us now consider the power law model magnetic lenses for which

n B(z)= B0knz , (101) where n is a nonnegative or negative integer. The case n = 0 corresponds to a constant magnetic field in the z-direction which cannot be used to focus an electron beam. In this case we know the exact solution in the classical situation: helical motion of the electron with cyclotron frequency around the z-direction. In the quantum situation the eigenstates

21 correspond to the Landau levels and one can construct coherent state wave packets for which x and y follow the classical helical paths [75]. The field corresponding to n = 1 also cannoth i beh usedi to focus an electron beam [41]. So, we shall consider n to be an integer− > 0 or < 1. For any− integer n, > 0 or < 1, we have − n qB0 α(z)= α0knz , with α0 = . (102) 2p0 The corresponding paraxial equation of motion is d2 R (z) h i + α2k2 z2n R (z)=0, (103) dz2 0 n h i where, as earlier, R denotes X or Y in the rotating Larmor frame of reference. With α0 −1 n having the dimension of (length) and knz being dimensionless for both positive and n+1 negative values of the integer n, note that α0knz is dimensionless. Then, defining R (z)= √zu(z), (104) h i and introducing the dimensionless variable α k zn+1 ζ = 0 n , (105) n +1 the paraxial equation (103) becomes d2u(ζ) du(ζ) 1 ζ2 + ζ + ζ2 u(ζ)=0. (106) dζ2 dζ − 4(n + 1)2 ! Recall that n is not 0 or 1 for us. Equation (106) is the Bessel differential equation of order 1/2(n + 1) and the two− linearly independent solutions for it are the Bessel functions of the first kind J1/2(n+1)(ζ) and J−1/2(n+1)(ζ). Bessel function of the first kind for any fractional order ν is defined by 1 ζ ν ∞ ( 1)j ζ 2j Jν(ζ)= − , (107) Γ(ν + 1) 2! (ν + 1)jj! 2! jX=0 where (a)j is the Pochhammer symbol, or the rising factorial, 1, for j =0, (a) = , (108) j a(a + 1)(a + 2) ... (a + j 1), for j 1, ( − ≥ and Γ is the gamma function (for more details on Bessel functions see e.g., [76–79]). In our case ν = 1/2(n+1). Thus, the fundamental solutions of the paraxial equation (103) ± for any integer n> 0 are given by n+1 n+1 (g) α0knz (g) α0knz gn (z, zi) = Cn,1√zJ 1 + Cn,2√zJ− 1 , 2(n+1) n +1 ! 2(n+1) n +1 ! n+1 n+1 (h) α0knz (h) α0knz hn (z, zi) = Cn,1√zJ 1 + Cn,2 √zJ− 1 , (109) 2(n+1) n +1 ! 2(n+1) n +1 !

22 (g) (g) (h) (h) where the constants Cn,1, Cn,2, Cn,1 , and Cn,2 are to be chosen such that the initial −n conditions (81) are satisfied. Note that when B(z)= B0k−nz we have ζ = α0k−n/((n n−1 2 2 − − 1)z ) and in (106) we can write 1/(2( n+1)) as 1/(2(n 1)) . Note also that Jν( ζ)= ν − − − ( 1) Jν(ζ). As a result, the corresponding fundamental solutions of the paraxial equation (103)− for any integer n< 1 can be written as − (g) α0k−n (g) α0k−n g−n (z, zi) = C−n,1√zJ 1 + C−n,2√zJ− 1 , 2(n−1) (n 1)zn−1 ! 2(n−1) (n 1)zn−1 ! − − (h) α0k−n (h) α0k−n h−n (z, zi) = C−n,1√zJ 1 + C−n,2√zJ− 1 , 2(n−1) (n 1)zn−1 ! 2(n−1) (n 1)zn−1 ! − − (110)

(g) (g) (h) (h) where the constants C−n,1, C−n,2, C−n,1, and C−n,2 are to be chosen such that the initial conditions (81) are satisfied. We shall look at some examples in the following. For any integer n > 0, we shall take zi = 0 so that the electron moves from the field-free region to the field region on the right (z > 0). Then the initial conditions to be satisfied are:

′ ′ gn(0, 0)=1, gn(0, 0)=0, hn(0, 0)=0, hn(0, 0)=1. (111) From (107) it is observed that

n+1 α0knz √zJ 1 =0, at z =0, (112) 2(n+1) n +1 ! and 1 n+1 − 2(n+1) α0knz 1 α0kn √zJ− 1 = , at z =0. (113) 2(n+1) n +1 ! 2n+1 2(n + 1)! Γ 2n+2 This suggests that we can take  

1 2(n+1) (g) (g) 2n +1 α0kn Cn,1 =0,Cn,2 =Γ , (114) 2n +2 2(n + 1)! so that gn(0, 0) = 1. In other words,

1 2(n+1) n+1 2n +1 α0kn α0knz gn(z, 0)=Γ √zJ− 1 . (115) 2(n + 1)! 2(n + 1)! " 2(n+1) n +1 !#

It is seen from (107) that the first term of gn(z, 0) is a constant and from the second term onwards, the j-th term is z2(j−1)(n+1). From this it follows that g′ (0, 0) = 0 as required. ∝ n Similar arguments lead to the result:

1 − 2(n+1) n+1 2n +3 α0kn α0knz hn(z, 0) = Γ √zJ 1 . 2(n + 1)! 2(n + 1)! " 2(n+1) n +1 !# (116)

23 ′ Using (107) it can be verified that hn(0, 0)=0 and hn(0, 0) = 1. For any integer n< 1 the field vanishes as z . We shall consider the electron − −→ ±∞ to move from the field-free region at towards the field region on the right i.e., zi = −∞ ′ . Then the initial conditions are: limz→−∞ g−n (z, ) = 1, limz→−∞ g−n (z, )= 0,−∞ lim h (z, ) = 0, lim h′ (z, ) = 1.−∞ From (107) it is observed−∞ that z→−∞ −n −∞ z→−∞ −n −∞ 1 2(n−1) α0k−n 1 α0k−n lim √zJ 1 (117) z→−∞ 2(n−1) (n 1)zn−1 ! −→ Γ 2n−1 2(n 1)! − 2(n−1) − and   1 − 2(n−1) α0k−n 1 α0k−n lim √zJ− 1 z. (118) z→−∞ 2(n−1) (n 1)zn−1 ! −→ Γ 2n−3 2(n 1)! − 2(n−1) − Then, it follows that we should take   1 − 2(n−1) (g) 2n 1 α0k−n (g) C−n,1 =Γ − , C−n,2 =0, (119) 2(n 1)! 2(n 1)! − − so that lim g (z, ) = 1. In other words, z→−∞ −n −∞ 1 − 2(n−1) 2n 1 α0k−n α0k−n g−n(z, )=Γ − √zJ 1 . (120) −∞ 2(n 1)! 2(n 1)! " 2(n−1) (n 1)zn−1 !# − − − It is seen from (107) that the first term of g−n(z, ) is a constant and from the second term onwards, the j-th term is z−2(j−1)(−∞n−1). From this it follows that ′ ∝ limz→−∞ g (z, ) = 0 as required. In view of the asymptotic limits in (117) and (118) we can take ∞ 1 2(n−1) 2n 3 α0k−n h−n(z, ) = Γ − −∞  2(n 1)! 2(n 1)!  − −  α0k−n √zJ− 1 × " 2(n−1) (n 1)zn−1 !# − 1 − 2(n−1) 2n 1 α0k−n lim ziΓ − − zi→−∞ 2(n 1)! 2(n 1)! − − α0k−n √zJ 1 , (121) × " 2(n−1) (n 1)zn−1 !#) − ′ such that limz→−∞ h−n (z, ) = 0 and limz→−∞ h−n (z, ) = 1. Let us now see how the−∞ Peano-Baker method works. From−∞ the general expressions in (85) and (86) we have

z z2 2 2n gn (z, zi) = 1 (α0kn) dz2 dz1 z1 − Zzi Zzi z z4 z3 z2 4 2n 2n +(α0kn) dz4 dz3 z3 dz2 dz1 z1 Zzi Zzi Zzi Zzi ... (122) − 24 and

z z2 2 2n hn (z, zi) = (z zi) (α0kn) dz2 dz1 z1 (z1 zi) − − Zzi Zzi − z z4 z3 z2 4 2n 2n +(α0kn) dz4 dz3 z3 dz2 dz1 z1 (z1 zi) Zzi Zzi Zzi Zzi − ..., (123) − where the electron is considered to be moving from zi to z. It should be noted that the initial conditions are satisfied automatically. For any integer n > 0, with zi = 0, it is straightforward to see from (122) and (123) that (α k zn+1)2 (α k zn+1)4 g (z, 0) = 1 0 n + 0 n ... n − (2n + 1)(2n + 2) (2n + 1)(2n + 2)(4n + 3)(4n + 4) − ∞ ( 1)j α k zn+1 2j = − 0 n 2n+1 j=0 j! 2(n + 1) ! X 2(n+1) j   1 2(n+1) n+1 2n +1 α0kn α0knz = Γ √zJ− 1 (124) 2(n + 1)! 2(n + 1)! " 2(n+1) n +1 !# and

n+1 2 n+1 4 (α0knz ) (α0knz ) hn(z, 0) = z 1 + ... " − (2n + 2)(2n + 3) (2n + 2)(2n + 3)(4n + 4)(4n + 5) − # ∞ ( 1)j α k zn+1 2j = z − 0 n 2n+3 j=0 j! 2(n + 1) ! X 2(n+1) j   1 − 2(n+1) n+1 2n +3 α0kn α0knz = Γ √zJ 1 . (125) 2(n + 1)! 2(n + 1)! " 2(n+1) n +1 !# For any integer n< 1, with z = , we have − i −∞ 2 1 α0k−n g−n(z, ) = 1 −∞ − (2n 1)(2n 2) zn−1 ! − − 1 α k 4 + 0 −n ... (2n 1)(2n 2)(4n 3)(4n 4) zn−1 ! − − − − − ∞ ( 1)j α k 2j = − 0 −n 2n−1 n−1 j=0 j! 2(n 1)z ! X 2(n−1) j −   − 1 2n 1 α k 2(n−1) = Γ − 0 −n 2(n 1)! 2(n 1)! − − α0k−n √zJ 1 (126) × " 2(n−1) (n 1)zn−1 !# − 25 and

2 1 α0k−n h−n(z, ) = z 1 −∞  − (2n 2)(2n 3) zn−1 ! − −  1 α k 4 + 0 −n ... (2n 2)(2n 3)(4n 4)(4n 5) zn−1 ! −  − − − − 2  1 α0k−n lim zi 1 − zi→−∞   − (2n 1)(2n 2) zn−1 !  − −  4  1 α k− + 0 n ... (2n 1)(2n 2)(4n 3)(4n 4) zn−1 ! −  − − − −   ∞ ( 1)j α k 2j  = z − 0 −n  2n−3 n−1  j=0 j! 2(n 1)z !  X 2(n−1) j −     ∞ j  2j  ( 1) α0k−n lim zi  −  − zi→−∞ 2n−1 2(n 1)zn−1 !  j=0 2(n−1) j!  X j −     1  2n 3 α k 2(n−1)  = Γ − 0 −n  2(n 1)! 2(n 1)!  − −  α0k−n √zJ− 1 × " 2(n−1) (n 1)zn−1 !# − 1 − 2(n−1) 2n 1 α0k−n lim ziΓ − − zi→−∞ 2(n 1)! 2(n 1)! − − α0k−n √zJ 1 . (127) × " 2(n−1) (n 1)zn−1 !#) − We see that the solutions (124-127) are the same as those obtained analytically in (115), (116), (120), and (121). Thus, the Peano-Baker method is a constructive process, incor- porating the initial conditions, leading to the same solutions as the analytical solutions. When it is not possible to get the analytical solutions in a particular case the Peano-Baker method provides a computational scheme to get the approximate solutions (see e.g., [80]). The above analytical solutions are well known and, based on them, practical aspects of the performances of axially symmetric magnetic lenses with the Glaser and power law models for B(z) have been analysed extensively in the literature (see [1, 2, 34–51]). Here, our objective has been mainly to demonstrate, with the examples of some models of round magnetic lenses, how quantum mechanics leads to the classical trajectories for the quantum average of position when the small quantum corrections are neglected.

26 8 Quantum mechanics of aberrations

When the incoming electron beam is not ideally paraxial we have to retain in the quantum electron beam optical Hamiltonian ˆo,c terms higher than quadratic in x , y , pˆx , and pˆ which lead to terms not linearH in x , y , pˆ , and pˆ in the equationsh i h i ofh motioni h yi h i h i h xi h yi (63-66). Thus, for a quasiparaxial beam we shall take

ˆ = ˆ + ˆ′ , (128) Ho,c Ho,p Ho ˆ ˆ′ as the Hamiltonian, where o,p and o are, respectively, the paraxial Hamiltonian and the lowest order non-paraxialH HamiltonianH given in (37-38). From (55) we know that when the beam is paraxial the relation between Oˆ (zi) and Oˆ (z), for any observable O, is given by h i h i Oˆ (z)= Uˆ † (z, z ) OˆUˆ (z, z ) (z ) , (129) h i p i p i i where D E i z Uˆp (z, zi)= P exp dz ˆo,p(z) . (130)  −h¯ Zzi H  From (82) we have the paraxial transfer map:

ˆ † ˆ x p(z) Up (z, zi) xUp (z, zi) (zi) h i h ˆ † ˆ i  y p(z)   Up (z, zi) yUp (z, zi) (zi)  1h i = 1h † i pˆx p(z) Uˆ (z, zi)ˆpxUˆp (z, zi) (zi)  p0   p0 p   1 h i   h i  pˆy p(z)  1 ˆ † ˆ   p0   Up (z, zi)ˆpyUp (z, zi) (zi)   h i   p0 h i      g (z, zi) R (z, zi) h (z, zi) R (z, zi) = ′ R ′ R g (z, zi) (z, zi) h (z, zi) (z, zi) ! x (z ) h i i  y (zi)  1h i . (131) pˆx (zi) ×  p0   1 h i   pˆy (zi)   p0 h i    To get the transfer map for the quasiparaxial beam we have to calculate

ˆ † ˆ x (z) U (z, zi) xU (z, zi) (zi) h i h ˆ † ˆ i  y (z)   U (z, zi) yU (z, zi) (zi)  1h i = 1h † i , (132) pˆx (z) Uˆ (z, zi)ˆpxUˆ (z, zi) (zi)  p0   p0   1 h i   1 h i  pˆy (z)  ˆ † ˆ   p0   U (z, zi)ˆpyU (z, zi) (zi)   h i   p0 h i      where i z Uˆ (z, zi) = P exp dz ˆo,c(z)  −h¯ Zzi H  z P i ˆ ˆ′ = exp dz o,p(z)+ o(z) . (133) −h¯ zi H H   Z   27 To get the transfer map (132), one has to use the formalism of the time-dependent per- turbation theory of quantum mechanics based on the interaction picture, replacing time t by z, treating ˆ′ (z) as a perturbation. The results corresponding to imaging by a round Ho magnetic lens are as follows (for details see, [21, 32]). Let us now choose zi = zob, the position of the object plane, and z = zim, the position of the image plane where a magnified, inverted, and rotated, image of the object is formed. Under ideal conditions we would have point-to-point imaging such that in (131) g (z , z ) = M, h (z , z ) = 0, g′ (z , z ) = 1/f, and h′ (z , z ) = 1/M where im ob − im ob im ob − im ob − M is the magnification and f is the focal length of the lens. Under quasiparaxial conditions the transfer map becomes

x (z ) x (z ) (∆x)(z ) h i im h ip im im  y (zim)   y p (zim)   (∆y)(zim)  1h i = 1h i + 1 pˆx (zim) pˆx p (zim) (∆px)(zim)  p0   p0   p0   1 h i   1 h i   1   pˆy (zim)   pˆy p (zim)   (∆py)(zim)   p0 h i   p0 h i   p0     MR (z , z )  0  = im ob − 1 R (z , z ) 1 R (z , z ) − f im ob − M im ob ! x (z ) (δx)(z ) h i ob ob  y (zob)   (δy)(zob)  1h i + 1 , pˆx (zob) (δpx)(zob) ×  p0   p0   1 h i   1   pˆy (zob)   (δpy)(zob)   p0 h i   p0      (134) where (∆x)(zim), (∆y)(zim), (∆px)(zim), and (∆py)(zim) are the third-order aberrations, lowest order deviations from the paraxial results involving the quantum averages of third- ~ order polynomials in ~r⊥, pˆ⊥ as seen in the following. Explicit expressions for (δx)(zob), (δy)(zob), (δpx)(zob), and (δpy)(zob) are given by C (δx)(z ) = pˆ pˆ2 (z ) ob p3 x ⊥ ob 0 D E K + pˆ , ~pˆ ~r + ~r ~pˆ + x, pˆ2 (z ) 2p2 x ⊥ · ⊥ ⊥ · ⊥ ⊥ ob 0 Dn  o n oE k 1 + pˆ , Lˆ y, pˆ2 (z ) p2 x z − 2 ⊥ ob 0 n o n o A ~ ~ + x, pˆ⊥ ~r⊥ + ~r⊥ pˆ⊥ (zob) 2p0 · · a Dn  oE + x, Lˆ y, ~pˆ ~r + ~r ~pˆ (z ) 2p z − ⊥ · ⊥ ⊥ · ⊥ ob 0 Dn o n  oE F 2 2 2 + pˆx,r⊥ (zob)+ D xr⊥ (zob) d yr⊥ (zob) , (135) 2p0 − Dn oE D E D E C (δy)(z ) = pˆ pˆ2 (z ) ob p3 y ⊥ ob 0 D E 28 K + pˆ , ~pˆ ~r + ~r ~pˆ + y, pˆ2 (z ) 2p2 y ⊥ · ⊥ ⊥ · ⊥ ⊥ ob 0 Dn  o n oE k 1 + pˆ , Lˆ + x, pˆ2 (z ) p2 y z 2 ⊥ ob 0 n o n o A ~ ~ + y, pˆ⊥ ~r⊥ + ~r⊥ pˆ⊥ (zob) 2p0 · · a Dn  oE + y, Lˆ x, ~pˆ ~r + ~r ~pˆ (z ) 2p z − ⊥ · ⊥ ⊥ · ⊥ ob 0 Dn o n  oE F + pˆ ,r2 (z )+ D yr2 (z )+ d xr2 (z ) , (136) 2p y ⊥ ob ⊥ ob ⊥ ob 0 Dn oE D E D E 1 K 2 k 2 (δpx)(zob) = 3 pˆxpˆ⊥ (zob) 3 pˆypˆ⊥ (zob) p0 −p − p 0 D E 0 D E A ~ ~ 2 pˆx, pˆ⊥ ~r⊥ + ~r⊥ pˆ⊥ (zob) −2p0 · · a Dn  oE pˆ , Lˆ + pˆ , ~pˆ ~r + ~r ~pˆ (z ) −2p2 x z y ⊥ · ⊥ ⊥ · ⊥ ob 0 Dn o n  oE F x, pˆ2 (z ) −2p2 ⊥ ob 0 Dn oE D 2 ~ ~ pˆx,r⊥ + x, pˆ⊥ ~r⊥ + ~r⊥ pˆ⊥ (zob) −2p0 · · Dn o n  oE d 1 x, Lˆ + pˆ ,r2 (z ) E xr2 (z ) , (137) −p z 2 y ⊥ ob − ⊥ ob 0 n o n o D E 1 K 2 k 2 (δpy)(zob) = 3 pˆypˆ⊥ (zob)+ 3 pˆxpˆ⊥ (zob) p0 −p p 0 D E 0 D E A ~ ~ 2 pˆy, pˆ⊥ ~r⊥ + ~r⊥ pˆ⊥ (zob) −2p0 · · a Dn  oE pˆ , Lˆ + pˆ , ~pˆ ~r + ~r ~pˆ (z ) −2p2 y z x ⊥ · ⊥ ⊥ · ⊥ ob 0 Dn o n  oE F y, pˆ2 (z ) −2p2 ⊥ ob 0 Dn oE D 2 ~ ~ pˆy,r⊥ + y, pˆ⊥ ~r⊥ + ~r⊥ pˆ⊥ (zob) −2p0 · · Dn o n  oE d 1 y, Lˆ pˆ ,r2 (z ) E yr2 (z ) , (138) −p z − 2 x ⊥ ob − ⊥ ob 0 n o n o D E where, A,ˆ Bˆ = AˆBˆ + BˆAˆ, the anticommutator of Aˆ and Bˆ. With g = g (z, zob), { } ′ ′ ′ ′ h = h (z, zob), g = g (z, zob), and h = h (z, zob), the aberration coefficients are given by

1 zim C = dz α4 αα′′ h4 +2α2h2h′2 + h′4 , 2 zob − Z n  o 1 zim K = dz α4 αα′′ gh3 + α2(gh)′hh′ + g′h′3 , 2 zob − Z n  o

29 zim 1 1 1 k = dz α′′ α3 h2 αh′2 , Zzob 8 − 2  − 2  1 zim A = dz α4 αα′′ g2h2 +2α2gg′hh′ + g′2h′2 α2 , 2 zob − − Z n  o zim 1 a = dz α′′ α3 gh αg′h′ , Zzob 4 −  −  1 zim F = dz α4 αα′′ g2h2 + α2 g2h′2 + g′2h2 + g′2h′2 +2α2 , 2 zob − Z n    o 1 zim D = dz α4 αα′′ g3h + α2gg′(gh)′ + g′3h′ , 2 zob − Z n  o zim 1 1 1 d = dz α′′ α3 g2 αg′2 , Zzob 8 − 2  − 2  1 zim E = dz α4 αα′′ g4 +2αg2g′2 + g′4 . (139) 2 zob − Z n  o In (135-136) C, K, k, A, a, F , D, and d are, respectively, the coefficients of terms causing spherical aberration, isotropic coma, anisotropic coma, isotropic astigmatism, anisotropic astigmatism, field curvature, isotropic distortion, and anisotropic distortion in the image (see e.g., [1] for detailed descriptions of these geometrical aberrations). The terms with the coefficient E in (137-138) cause aberrations of the ray gradients which do not affect the image by a single lens, but will have to be taken into account in multilens systems. This aberration has no name in electron optics and has been named pocus in light optics [81, 82]. The expressions for all the aberration coefficients in (139) are the same as in classical electron optics. 2 As noted earlier, a term like pˆxpˆ⊥ (zob) in δx (zob), for example, is ( pˆ ( pˆ 2 + pˆ 2))(z ) plus additionalh termsi dependingh oni quantum uncertainties h xi h xi h yi ob which are uncontrollable in principle. Thus, quantum uncertainties contribute to aber- rations. There are also other tiny quantum corrections to aberrations due to theh ¯- dedependent terms dropped from the quantum beam optical Hamiltonian. Theh ¯- 2 ˆ(¯h) ˆ dependent term r⊥ in o in (39) is a paraxial term which, if added to o,p, will modify α(z) affecting∝ theH Larmor rotation, focal length, and the aberration coefficients,H ˆ(¯h) 4 in a tiny way. The otherh ¯-dependent term in o , r⊥, is a perturbation term which, if ˆ′ H ∝ added to o, will modify the pocus and will have a tiny influence in a multilens system. Let usH end this section with a note on the expression for the spherical aberration coefficient C in (139). There are many expressions available for the various aberration co- efficients in the literature. As explained in [1], though there may be several expressions for an aberration coefficient, very different in appearance, they are otherwise equivalent and can be obtained from one another by partial integration and replacing the second deriva- tives of g and h using the paraxial equation. Thus, for C there are various equivalent expressions, the most important and immensely influential being Scherzer’s expression which showed that it is always positive for a round lens, free of space charge and with a static electromagnetic field, and hence the spherical aberration is unavoidable in such lenses. We shall study the equivalence of the expression for C in (139) with Scherzer’s

30 expression and an expression used in [42] (see also [1]). The expressions in (139) for all the aberration coefficients, obtained using quantum mechanics [21,23,32] are the same as those obtained in [83] using the Lie algebraic approach to classical electron optics which reproduces all the classical results exactly. This is not surprising since the quantum elec- tron beam optics developed in [21,23,32] becomes the Lie algebraic approach to classical electron optics when the quantum classical correspondence rule (61) is used. Lie ←→ algebraic methods have been developed extensively for classical charged particle beam optics, particularly accelerator beam optics (see e.g., [83–86]). Let us now start with the expression for C given in (139), rewritten as

1 zim C = dz α4h4 +2α2h2h′2 αα′′h4 + h′4 . (140) 2 zob − Z n o

Integrating the last term by parts, using the boundary conditions h (zob, zob) = 0 and h (z , z ) = 0 and the relation h′′ (z, z ) = α(z)2h (z, z ) following from the paraxial im ob ob − ob equation of motion (79) having h (z, zob) as one of its solutions, we get

zim z z z ′4 ′ ′3 ′3 z ′2 ′′ 2 2 ′2 dz h = dz h h = hh zob dz 3hh h =3 dzα h h . (141) Zzob Zzob | − Zzob Zzob Integrating the third term by parts leads to

zim zim dz αα′′h4 = dzα′′ αh4 − Zzob − Zzob  zim ′ 4 zim ′ ′ 4 3 ′ = α αh zob + dzα α h +4αh h − | Zzob zim    = dz α′2h4 +4αα′h3h′ . (142) zob Z n o Integrating the term αα′h3h′ by parts, and using the boundary conditons and the paraxial equation, we find

zim zim zim ′ ′ 3 ′ ′ 3 ′ 3 ′ zim 3 ′ dz αα h h = dzα αh h = α αh h zob dzα αh h Zzob Zzob | − Zzob zim       = dz αα′h3h′ 3α2h2h′2 + α4h4 , (143) zob − − Z n o which can be written in the equivalent form

zim 1 zim dz αα′h3h′ = dz 3α2h2h′2 + α4h4 . (144) zob 2 zob − Z Z n o Replacing the third and the last terms in the integral (140) using the relations (141), (142), and (144), we get

1 zim C = dz 5α2h2h′2 + α′2h4 +4αα′h3h′ + α4h4 . (145) 2 zob Z n o

31 Now, if we write the third term in the above equation as (2 + 2)αα′h3h′ and replace one part 2αα′h3h′ using (144), keeping the other part 2αα′h3h′ as is, we obtain the celebrated result of Scherzer 1 zim C = dz 2α4h4 + h2 (hα′ + h′α)2 + α2h2h′2 . (146) 2 zob Z n o Now, we shall relate the expression for C in (139) to another expression 1 zim C = dz h4 4b4 bb′′ +5b′2 , (147) 48 zob − Z n o in which e b(z)= B(z)=2α(z), (148) r2mU used in [42] (see also [1]). To this end we proceed as follows. As earlier, we use integration by parts, boundary conditions on h, and the paraxial equation. First, we find an expression for the integral of αα′h3h′, different from (143), as given by zim zim ′ ′ 3 ′ ′ 3 zim ′ 3 dz αα h h = αα h h zob dz h αα h Zzob | − Zzob  zim    = dz α′2 + αα′′ h4 +3αα′h3h′ , (149) − zob Z n  o which can be written in the equivalent form zim 1 zim dz αα′h3h′ = dz α′2 + αα′′ h4. (150) zob −4 zob Z Z   Similarly, for the integral of α2h2h′2 we find zim zim ′ 2 2 ′2 2 2 ′ zim 2 2 ′ dzα h h = α h h h zob dz h α h h Zzob | − Zzob  zim    = dz 2αα′h3h′ 2α2h2h′2 + α4h4 , (151) zob − − Z n o which can be written in the equivalent form zim 1 z dzα2h2h′2 = dz 2αα′h3h′ + α4h4 . (152) zob 3 zob − Z Z n o Let us replace the last term in (140) using (141) leading to the expression 1 zim C = dz α4h4 αα′′h4 +5α2h2h′2 . (153) 2 zob − Z n o Next, we replace the third term in (153) using (152) to obtain 1 zim 8 10 C = dz α4h4 αα′′h4 αα′h3h′ . (154) 2 Zzob  3 − − 3  Then, we replace the last term in (154) using (150) to get (147), 1 zim C = dz h4 16α4 αα′′ +5α′2 , (155) 12 zob − Z n o which gives the expression for C used in [42] when α = b/2 as seen in (148).

32 9 Concluion

To summarize, we have considered the scalar theory of quantum mechanics of electron beam optics, at the single-particle level, derived from the Dirac equation using a Foldy- Wouthuysen-like transformation technique. Guided by the Ehrenfest theorem, quantum averages of position and momentum of a beam electron in a plane transverse to the optic axis of an electron optical system have been identified with the classical ray coordinates. Round magnetic electron lenses with Glaser and power law models for the axial magnetic field have been studied in particular. We have found that in the paraxial approximation the quantum average of position obeys the classical paraxial equation of motion. The fun- damental solutions of the paraxial equations for the lenses considered, obtained by solving the differential equations, are well known. In the case of the power law model lenses the fundamental solutions have also been constructed using the Peano-Baker series. We have discussed the quantum propagator for paraxial propagation of the beam wave function along the optic axis of the system. Quantum mechanics of aberrations due to deviation from paraxial condition has been discussed briefly. Role of quantum uncertainties in the nonlinear part of the equations of motion for a nonparaxial beam, and in aberrations, has been pointed out. As remarked by Hawkes [87], quantum corrections to the classical theory of electron optics are ‘fortunately usually negligible’, and a long article by Majert and Kohl [88] on the simulation of atomically resolved elemental maps with a multislice algorithm for relativistic electrons shows that there are practical situations in which it is essential to use the Dirac theory. Thus, with the development of nano-level technology in future electron beam optical devices quantum effects may have to be considered seriously.

References

[1] P.W. Hawkes, E. Kasper, Principles of Electron Optics - Vol.1: Basic Geometrical Optics, 2nd Ed., Elsevier, 2017.

[2] P.W. Hawkes, E. Kasper, Principles of Electron Optics - Vol.2: Applied Geometrical Optics, 2nd Ed., Elsevier, 2017.

[3] P.W. Hawkes, E. Kasper, Principles of Electron Optics - Vol.3: Wave Optics, Aca- demic Press, 1994.

[4] A.B. El-Kareh, J.C.J. El-Kareh, Electron Beams, Lenses, and Optics - Vols.1 & 2, Academic Press, 1970.

[5] H. Wollnik, Optics of Charged Particles, Academic Press, 1987.

[6] M. Szilagyi, Electron and Ion Optics, Plenum Press, 1988.

[7] H. Liebl, Applied Charged Particle Optics, Springer, 2008.

[8] J. Orloff, Handbook of Charged Particle Optics, 2nd Ed., Taylor & Francis, 2009.

33 [9] H. Rose, Geometrical Charged Particle Optics, 2nd Ed., Springer, 2012. [10] W. Glaser, Grundlagen der Elektronenoptik, Springer, 1952. [11] W. Glaser, P. Schiske, Elektronenoptische abbildung auf grund der wellenmechanik, Annalen der Physik 12 (1953) 240-280, https://doi.org/10.1002/andp.19534470408. [12] W. Glaser, Elektronen und Ionenoptik, in: S. Fl¨ugge (Ed.), Handbuch der Physik, Vol.33, pp. 123-395, Springer, 1956. [13] H.A. Ferwerda, B.J. Hoenders, C.H. Slump, Fully relativistic treatment of electron optical image formation based on the Dirac equation, Opt. Acta 33 (1986) 145-157, http://dx.doi.org/10.1080/713821923. [14] H.A. Ferwerda, B.J. Hoenders, C.H. Slump, The fully relativistic foundation of linear transfer theory in electron optics based on the Dirac equation, Opt. Acta 33 (1986) 159-183, http://dx.doi.org/10.1080/713821925. [15] T.R. Groves, Charged particle optics theory: An introduction, Taylor & Francis, 2015. [16] A. Lubk, Paraxial quantum mechanics, in: P.W. Hawkes (Ed.), Advances in Imaging and Electron Physics, Vol.206, pp. 1-323, Academic Press, 2018. [17] G. Pozzi, Particles and waves in electron optics and microscopy, in: P.W. Hawkes (Ed.) Advances in Imaging and Electron Physics, Vol.194, pp. 1-334, Academic Press, 2016. [18] R. Jagannathan, R. Simon, E.C.G. Sudarshan, N. Mukunda, Quantum theory of magnetic electron lenses based on the Dirac equation, Phys. Lett. A 134 (1989) 457- 464, http://dx.doi.org/10.1016/0375-9601(89)90685-3. [19] R. Jagannathan, Quantum theory of electron lenses based on the Dirac equation, Phys. Rev. A 42 (1990) 6674-6689, http://dx.doi.org/10.1103/PhysRevA.42.6674. [20] S.A. Khan, R. Jagannathan, Quantum mechanics of charged particle beam transport through magnetic lenses, Phys. Rev. E 51 (1995) 2510-2515, https://doi.org/10.1103/PhysRevE.51.2510. [21] R. Jagannathan, S.A. Khan, Quantum theory of the optics of charged particles, in: P.W. Hawkes (Ed.), Advances in Imaging and Electron Physics, Vol. 97, pp. 257-358, Academic Press, 1996, http://dx.doi.org/10.1016/S1076-5670(08)70096-X. [22] M. Conte, R. Jagannathan, S.A. Khan, M. Pusterla, Beam optics of the Dirac particle with anomalous magnetic moment, Particle Accelerators 56 (1996) 99-126, http://cds.cern.ch/record/307931/files/p99.pdf.

34 [23] S.A. Khan, Quantum Theory of Charged-Particle Beam Optics, PhD Thesis, Univer- sity of Madras, Chennai, India, 1997, Complete thesis available from Dspace of IMSc Library, The Institute of Mathematical Sciences, Chennai, India, where the doctoral research was done, http://www.imsc.res.in/xmlui/handle/123456789/75.

1 [24] R. Jagannathan, The Dirac equation approach to spin- 2 particle beam optics, in: P. Chen (Ed.), Proceedings of the 15th Advanced ICFA Beam Dynamics Workshop on Quantum Aspects of Beam Physics, Monterey, California, USA, 1998, pp. 670-681, World Scientific, 1999.

1 [25] S.A. Khan, Quantum theory of magnetic quadrupole lenses for spin- 2 particles, in: P. Chen (Ed.), Proceedings of the 15th Advanced ICFA Beam Dynamics Workshop on Quantum Aspects of Beam Physics, Monterey, California, USA, 1998, pp. 682-694, World Scientific, 1999.

[26] S.A. Khan, Quantum aspects of accelerator optics, in: A. Luccio, W. MacKay (Eds.), Proceedings of the 1999 Conference (PA99), New York, 1999, pp. 2817-2819, http://dx.doi.org/10.1109/PAC.1999.792948.

[27] R. Jagannathan, Quantum mechanics of Dirac particle beam optics: Single-particle theory, in: P. Chen (Ed.) Proceedings of the 18th Advanced ICFA Beam Dynamics Workshop on Quantum Aspects of Beam Physics, Capri, Italy, 2000, pp. 568-577, World Scientific, 2002, https://doi.org/10.1142/9789812777447_0047.

[28] S.A. Khan, Quantum formalism of beam optics, in: P. Chen (Ed.), Proceed- ings of the 18th Advanced ICFA Beam Dynamics Workshop on Quantum As- pects of Beam Physics, Capri, Italy, 2000, pp. 517-526, World Scientific, 2002, http://dx.doi.org/10.1142/9789812777447_0042.

[29] R. Jagannathan, Quantum mechanics of Dirac particle beam transport through op- tical elements with straight and curved optical axes, in: P. Chen, K. Reil (Eds.) Proceedings of the 28th Advanced ICFA Beam Dynamics and Advanced & Novel Accelerators Workshop, Hiroshima, Japan, 2003, pp. 13-21, World scientific, 2004, https://doi.org/10.1142/9789812702333_0002.

[30] S.A. Khan, Quantum aspects of charged particle beam optics, in: A. Al-Kamli, N. Can, G.O. Souadi, M. Fadhali, A. Mahdy, M. Mahgoub (Eds.), Proceedings of the 5th Saudi International Meeting on Frontiers of Physics - 2016 (SIMFP 2016), Gizan, Saudi Arabia, AIP Conference Proceedings, 1742 (2016) pp. 030008-1–030008- 4, http://dx.doi.org/10.1063/1.4953129.

[31] S.A. Khan, E.C.G. Sudarshan and the quantum mechanics of charged-particle beam optics, Current Science 115 (2018) 1813-1814, http://www.currentscience.ac.in/Volumes/115/09/1813.pdf.

35 [32] R. Jagannathan and S.A. Khan, Quantum Mechanics of Charged Particle Beam Optics: Understanding Devices from Electron Microscopes to Particle Ac- celerators, Taylor & Francis, 2019, https://doi.org/10.1201/9781315232515, http://isbn.nu/9781138035928/.

[33] S.A. Khan and R. Jagannathan, Quantum mechanics of bending of a nonrel- ativistic charged particle beam by a dipole magnet, Optik 206 (2020) 163626, https://doi.org/10.1016/j.ijleo.2019.163626.

[34] P.W. Hawkes, Magnetic lens theory, in: P.W. Hawkes (Ed.), Magnetic electron lenses, pp. 1-51, Springer, 1982, http://dx.doi.org/10.1007/978-3-642-81516-4.

[35] U.F. Gianola, Investigation of magnetic lenses having the axial field H(0, z) = γ/zn, Proc. Phys. Soc. (London) B 65 (1952) 597-603, http://iopscience.iop.org/0370-1301/65/8/309.

[36] H. H¨ansel, Uber¨ eine magnetische feldverteilung mit exact i¨osbarer paraxialgleichung, Optik 21 (1964) 273-280.

[37] A. Alshwaikh, T. Mulvey, in: D.L. Misell (Ed.), Developments in Electron Microscopy and Analysis, 1977, Institute of Physics, Bristol, Conference Series No.36 (1977) pp. 25-28.

[38] S.M. Al-Hilly, T. Mulvey, in: M.J. Goringe (Ed.), Electron Microscopy and Analysis, 1981, Institute of Physics, Bristol, Conference Series No.61 (1982) pp. 103-106.

[39] T. Mulvey, Unconventional lens design, in: P.W. Hawkes (Ed.), Magnetic Electron Lenses, pp.359-417, Springer, 1982, http://dx.doi.org/10.1007/978-3-642-81516-4.

[40] M. Lenc, Immersion objective lenses in electron optics, PhD Dissertation, Delft Uni- versity of Technology, The Netherlands, 1992.

[41] A. Crewe, On the peculiarities of monopole and multipole focusing, Optik 112 (2001) 181-183, https://doi.org/10.1078/0030-4026-00035.

[42] P.W. Hawkes, On the optical properties of magnetic lenses with fields of the form B(z) z−n, n = 2, 3, 4, Optik 113 (2002) 273-275, https://doi.org/10.1078/0030-4026-00154∝ .

[43] Z. Liu, Differential algebraic analysis of optical properties of monopole, dipole, and quadrupole lenses, Optik 114 (2003) 518-520, https://doi.org/10.1078/0030-4026-00302.

[44] A. Crewe, A comment on the properties of multipole lenses, Optik 114 (2003) 449- 450, https://doi.org/10.1078/0030-4026-00294.

36 [45] A.S.A. Alamir, Spiral distortion of magnetic lenses with fields of the form B(z) z−n, n = 2, 3, 4, Optik 114 (2003) 525-528, https://doi.org/10.1078/0030-4026-00319∝ .

[46] A. Crewe, Electron focusing in magnetic fields of the form zn, Optik 115 (2004) 31-35, https://doi.org/10.1078/0030-4026-00317.

[47] A.S.A. Alamir, On the chromatic aberration of magnetic lenses with a field distribu- tion in the form of an inverse power low (B(z) z−n), Optik 115 (2004) 227-231, https://doi.org/10.1078/0030-4026-00358. ∝

[48] A.S.A. Alamir, On the optical properties of monopole, multipole magnetic lenses, Optik 116 (2005) 429-432, https://doi.org/10.1016/j.ijleo.2005.02.016.

[49] A.S.A. Alamir, Magnetic lenses performance with fields of the form B(z) zn, Optik ∝ 120 (2009) 610-613, https://doi.org/10.1016/j.ijleo.2008.02.006.

[50] A.S.A. Alamir, Radial and spiral distortion of magnetic lenses with fields of the form B(z) zn, Optik 120 (2009) 984-986, https://doi.org/10.1016/j.ijleo.2008.04.004∝ .

[51] A.S.A. Alamir, On chromatic aberration of magnetic lenses with a field distri- bution in the form of power low model B(z) zn, Optik 122 (2011) 273–275, https://doi.org/10.1016/j.ijleo.2009.12.016∝ .

[52] J.D. Bjorken and S.D. Drell, Relativistic Quantum Mechanics, McGraw-Hill, 1994.

[53] R.M. Wilcox, Exponential operators and parameter differentiation in quantum physics, J. Math. Phys. 8 (1967) 962-982, https://doi.org/10.1063/1.1705306.

[54] W. Greiner, Relativistic Quantum Mechanics: Wave Equations, 3rd Ed., Springer, 2000.

[55] L. Fishman, One-way wave equation modeling in two-way wave propagation prob- lems, in: B. Nilsson, L. Fishman (Eds.), Mathematical Modelling of Wave Phenomena 2002, Mathematical Modelling in Physics, Engineering, and Cognitive Sciences, Vol. 7, V¨axj¨oUniversity Press, V¨axj¨o, Sweden, (2004), pp. 91-111.

[56] S.A. Khan, R. Jagannathan, R. Simon, Foldy-Wouthuysen transforma- tion and a quasiparaxial approximation scheme for the scalar wave the- ory of light beams, E-Print: arXiv:physics/0209082 [physics.optics] (2002) http://arXiv.org/abs/physics/0209082.

[57] S.A. Khan, The Foldy-Wouthuysen transformation technique in optics, Optik 117 (2006) 481-488, http://dx.doi.org/10.1016/j.ijleo.2005.11.010.

37 [58] S.A. Khan, The Foldy-Wouthuysen transformation technique in optics, in: P.W. Hawkes (Ed.), Advances in Imaging and Electron Physics, Vol.152, pp. 49-78, Aca- demic Press, 2008, http://dx.doi.org/10.1016/S1076-5670(08)00602-2. [59] S.A. Khan, Quantum methodologies in Helmholtz optics, Optik 127 (2016) 9798- 9809. http://dx.doi.org/10.1016/j.ijleo.2016.07.071. [60] S.A. Khan, Linearization of wave equations, Optik 131 (2017) 350-363, http://dx.doi.org/10.1016/j.ijleo.2016.11.073. [61] S.A. Khan, Passage from scalar to vector optics and the Mukunda-Simon- Sudarshan theory for paraxial systems, J. Mod. Opt. 63 (2016) 1652-1660, http://dx.doi.org/10.1080/09500340.2016.1164257. [62] S.A. Khan, Quantum methods in light beam op- tics, Optics & Photonics News 27 (2016) 47, http://www.osa-opn.org/home/articles/volume_27/december_2016/features/optics_in_2016/. [63] S.A. Khan, Hamilton’s optical-mechanical analogy in the wavelength-dependent regime, Optik 130 (2017) 714-722. http://dx.doi.org/10.1016/j.ijleo.2016.07.071. [64] W. Greiner, Quantum Mechanics, 4th Ed., Springer, 2001. [65] D.J. Griffiths, D.F. Schroeter, Introduction to Quantum Mechanics, 3rd Ed., Cam- bridge University Press, 2018. [66] W. Magnus, On the exponential solution of differential equations for a linear operator, Commun. Pure Appl. Math. 7 (1954) 649-673. https://doi.org/10.1002/cpa.3160070404 [67] S. Blanes, F. Casas, J.A. Oteo, J. Ros, The Magnus expan- sion and some of its applications, Phys. Rep. 470 (2009) 151-238, https://doi.org/10.1016/j.physrep.2008.11.001. [68] L.A. Pipes, L.R. Harvill, Applied Mathematics for Engineers and Physicists, 3rd Ed., Dover, 2014. [69] M. Baake, M., U. Schl¨agel, The Peano-Baker series, Proc. Steklov Inst. Math. 275 (2011) 167-171, https://doi.org/10.1134/S0081543811080098, https://arxiv.org/abs/1011.1775. [70] K.B. Wolf, On time-dependent quadratic quantum Hamiltonians, SIAM J. Appl. Math. 40 (1981) 419-431, http://www.jstor.org/stable/2101339. [71] F. Lenz, Properties of electron lenses, in: P.W. Hawkes (Ed.), Magnetic electron lenses, pp. 119-161, Springer, 1982, http://dx.doi.org/10.1007/978-3-642-81516-4.

38 [72] S. L¨offler, A. Hamon, D. Aubry, P. Schattschneider, A quantum propagator for electrons in a round magnetic lens, in: M. H¨ytch, P.W. Hawkes (Eds.), Ad- vances in Imaging and Electron Physics, Vol.215, pp. 89-105, Elsevier, 2020, https://doi.org/10.1016/bs.aiep.2020.06.003. [73] L. Zou, P. Zhang, A.J. Silenko, Paraxial wave function and Guoy phase for a rel- ativistic electron in a uniform magnetic field, E-Print: arXiv:2003.04717[quant-ph] (2020), https://arxiv.org/abs/2003.04717. [74] K.Y. Bliokh, I.P. Ivanov, G. Guzzinati, L. Clark, R. Van Boxem, A. B´ech´e, R. Juchtmans, M.A. Alonso, P. Schattschneider, F. Nori, J. Verbeeck, The- ory and applications of free-electron vortex states, Phys. Rep. 690 (2017) 1-70, http://dx.doi.org/10.1016/j.physrep.2017.05.006. [75] I.A. Malkin, V.I. Man’ko, Coherent states of a charged parti- cle in a magnetic field, Soviet Phys. JETP, 28 (1969) 527-532, http://jetp.ac.ru/cgi-bin/dn/e_028_03_0527.pdf. [76] I.S. Gradshteyn, I.M. Ryzhik, Table of Integrals, Series, and Products, Elsevier, 2007. https://doi.org/10.1016/C2009-0-22516-5. [77] M. Abramowitz, I.A. Stegun, Handbook of Mathematical Functions with Formu- las, Graphs, and Mathematical Tables, Dover, 2014, http://dlmf.nist.gov/ and http://people.math.sfu.ca/~cbm/aands/. [78] G.B. Arfken, H.J. Weber, F.E. Harris, Mathematical Methods for Physicists: A Comprehensive Guide, 7th Ed., Academic Press, 2012. [79] V. Lakshminarayanan, L.S. Varadharajan, Special Functions for Optical Science and Engineering, SPIE: The International Society for Optics and Photonics, 2015, https://spie.org/Publications/Book/2207307?SSO=1. [80] N. Vaseghi, M.S. Abrishamian, Exact frequency domain method for the analysis of scattering from multilayer bi-anisotropic cylindrical structures, Appl. Opt. 59 (2020) 3447-3457, https://doi.org/10.1364/AO.388258. [81] A.J. Dragt, E. Forest, K.B. Wolf, Foundations of a Lie algebraic theory of ge- ometrical optics, in: J.S. Mondragon, K.B. Wolf (Eds.), Lie Methods in Op- tics, Springer Lecture Notes in Physics, Vol.250, pp. 105-158, Springer, 1986, http://dx.doi.org/10.1007/3-540-16471-5_4. [82] S.A. Khan, Aberrations in Helmholtz optics, Optik 153 (2018) 164-181, https://doi.org/10.1016/j.ijleo.2017.10.006 [83] A.J. Dragt, E. Forest, Lie algebraic theory of charged particle optics and electron microscopes, in: P.W. Hawkes (Ed.), Advances in Imag- ing and Electron Physics, Vol. 67, pp. 65-120, Academic Press, 1986, http://dx.doi.org/10.1016/S0065-2539(08)60329-7.

39 [84] A.J. Dragt, F. Neri, G. Rangarajan, D.R. Douglas, L.M. Healy, R.D. Ryne, Lie alge- braic treatment of linear and nonlinear beam dynamics, Ann. Rev. Nucl. Part. Sci. 38 (1988) 455-496, https://doi.org/10.1146/annurev.ns.38.120188.002323.

[85] A.J. Dragt, Lie Methods for Nonlinear Dynamics with Appli- cations to Accelerator Physics, University of Maryland, 2018, https://physics.umd.edu/dsat/dsatliemeth.

[86] T. Radli˘cka, Lie algebraic methods in charged particle optics, in: P.W. Hawkes (Ed.), Advances in Imaging and Electron Physics, Vol.151, pp. 241-362, Academic Press, 2008, http://dx.doi.org/10.1016/S1076-5670(07)00404-1.

[87] P.W. Hawkes, Dirac, c and a Supper date, Ultramicroscopy 213 (2020) 112981, https://doi.org/10.1016/j.ultramic.2020.112981.

[88] S. Majert, H. Kohl, Simulation of atomically resolved elemental maps with a smul- tislice algorithm for relativistic electrons, in: P.W. Hawkes, M. H¨ytch (Eds.), Advances in Imaging and Electron Physics, Vol.211, pp. 1-120, Elsevier, 2019, https://doi.org/10.1016/bs.aiep.2019.04.001

40