Exact Analytic Solutions of Maxwell's Equations Describing Propagating Nonparaxial Electromagnetic Beams

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/262105913

Exact analytic solutions of Maxwell's equations describing propagating


nonparaxial electromagnetic beams

Article  in  Applied Optics · April 2014


DOI: 10.1364/AO.53.004524

CITATIONS READS

6 501

2 authors:

Roger Garay Michel Zamboni-Rached


Universidad Nacional de Itapúa University of Campinas
2 PUBLICATIONS   11 CITATIONS    146 PUBLICATIONS   1,782 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Biology and electromagnetics. Does a virus have a metamaterial invisibility cloak ? View project

Computational holographic techniques applied in experimental generation of nondiffracting beams and structured light View project

All content following this page was uploaded by Michel Zamboni-Rached on 07 May 2014.

The user has requested enhancement of the downloaded file.


Exact analytic solutions of Maxwell’s equations
describing propagating nonparaxial electromagnetic
beams.

Roger L. Garay-Avendaño and Michel Zamboni-Rached

DECOM-FEEC, University of Campinas, Campinas, SP, Brazil.

Abstract
In this paper we propose a method capable of describing in exact and analytic
form the propagation of nonparaxial scalar and electromagnetic beams. The main
features of the method presented here are its mathematical simplicity and the fast
convergence in the cases of highly nonparaxial electromagnetic beams, enabling us
to obtain high-precision results without the necessity of lengthy numerical simula-
tions or other more complex analytical calculations. The method can be used in
electromagnetism (optics, microwaves) as well as in acoustics.

Keywords: Nonparaxial scalar and electromagnetic beams; Bessel beams; Wave equation;
Maxwell equations; polarization. Linear, azimuthal and radial polarization

1 Introduction

In recent years particular attention has been paid to the nonparaxial electromagnetic
beams which possess highly focused field distributions, with spot sizes of a few wave-
lengths.
These beams can be treated, in general, neither with paraxial theories, nor with scalar
approaches. In these cases the vectorial nature of the field must be taken into account,
with a mathematical description beyond the paraxial approximation.
Up to now, different vectorial formulations to obtain nonparaxial electromagnetic
beams have been explored in the literature [3, 4]. Some methods start with paraxial
solutions, and then include small nonparaxial corrections [5]. The main limitation with

1
those approaches is the slow convergence presented by them when applied to the cases
of highly nonparaxial beams. Other proposed methods are mathematically too complex
[6, 7] and can only be used to get particular solutions, such as Gaussian beams [8, 9].
The aim of the present work is to adapt to the case of beams an exact scalar method
developed to pulses[1], and afterword, based on that scalar method, to develop two new
vectorial approaches capable of describing nonparaxial electromagnetic beams as exact
solutions of the Maxwell equations.
Our methods, compared to the others, have the advantage of being faster (fast conver-
gence) when dealing with highly nonparaxial beams. Another advantage is its mathemat-
ical simplicity: the only necessary integrations are those to find simple Fourier coefficients
of the longitudinal wavenumber spectrum, S(kz ) (with −ω/c ≤ kz ≤ ω/c), of the desired
beam, all the other calculations are based on simple derivatives.
These new methods can be applied to describe beams of interest to optical tweez-
ers[10,11], atom guiding, high-resolution microscopy [12], etc..
Section 2 starts by exposing some concepts of paraxial and nonparaxial beams and
then proposes a scalar method for describing nonparaxial scalar beams as exact solutions
of the wave equation. In Section 3 we apply our new vectorial methods for obtaining
exact nonparaxial electromagnetic beams with linear, radial and azimuthal polarization.
Section 4 is devoted to the conclusions.

2 Scalar Analysis

The first subsection forwards a brief overview of the concepts involving paraxial and
nonparaxial scalar beams. The rest of this section is devoted to the description of the
method for describing scalar nonparaxial beams.

2
2.1 Paraxial and nonparaxial scalar beams

In free space, any propagating ∗ beam with azimuthal symmetry, ψ(ρ, z, t), solution of
the wave equation, ∇2 ψ − ∂ct
2
ψ = 0, can be written as a superposition of zero-order Bessel
beams[19] over the longitudinal wavenumber kz :

Z ω/c  q 
ψ(ρ, z, t) = a exp[−iωt] J0 ρ ω 2 /c2 − kz2 exp[ikz z]S(kz )dkz (1)
−ω/c

where ω is the angular frequency, S(kz ) is a weight function (spectrum) and a is a con-
stant with unitary magnitude and with dimensions depending on the physical quantity
represented by ψ.
In eq.(1), the range of integration over kz is limited to −ω/c ≤ kz ≤ ω/c to avoid
evanescent waves. However, this range considers propagating (kz > 0) and counter-
propagating (kz < 0) Bessel beams. To avoid the counter-propagating components in
eq.(1), the spectrum S(kz ) can be chosen in order to minimize or even cancel their con-
tribution to the resulting beam [13-18].
Note also that, depending on the spectrum S(kz ), the resulting wave can be paraxial
or nonparaxial. For paraxial beams, the spectra are strongly concentrated around kz =
ω/c, with a bandwidth ∆kz << ω/c. On the other hand, nonparaxial beams possess
wide spectrum bandwidth (∆kz of the same order of ω/c), but can also present narrow
bandwidths if concentrated far from kz = ω/c.
The main problem to obtain exact analytic solutions describing totally propagating
nonparaxial beams is to solve the integral (1).
In the following, we present a mathematical method capable to solve (1) for any
spectral function S(kz ), allowing us to obtain exact analytic solutions to the wave equation
describing nonparaxial scalar beams.

All beams considered in this work are totally propagating, with no evanescent component.

3
2.2 Describing nonparaxial propagating scalar beams

Let us start by considering in (1) a constant spectrum S(kz ) = (1/2)(c/ω). Then we


get:
Z ω/c  q 
1c
ψ(ρ, z, t) = a exp[−iωt] J0 ρ ω 2 /c2 − kz2 exp[ikz z] dkz (2)
−ω/c 2ω
Solving the integral in (2), we obtain[20]:
s 
ω2 ω2
ψ(ρ, z, t) = a exp[−iωt] sinc  2 ρ2 + 2 z 2  (3)
c c

where, sinc[x] = sin[x]/[x].


If we consider in (1) another type of spectrum such as S(kz ) =
(c/2ω) exp [−i2πnkz /K], with K = 2ω/c, we get:
 s 
ac ω/c ω2 2πn
Z    
ψ(ρ, z, t) = exp[−iωt] J0 ρ − kz2  exp i z+ kz dkz (4)
2ω −ω/c c2 K

Using the result (3), we can solve the integral in (4) and obtain:

s 
2
ω2 ω

ψ(ρ, z, t) = a exp[−iωt] sinc  2 ρ2 + z + πn  (5)
c c
By noting that the integral solution (1) possesses finite integration limits, we can
expand the spectrum S(kz ) as a Fourier series:

2πn
X  
S(kz ) = Rn exp i kz (6)
n=−∞ K

within −ω/c ≤ kz ≤ ω/c, being the Fourier coefficients given by:

1 Z ω/c 2π
 
Rn = S(kz ) exp −i nkz dkz (7)
K −ω/c K

where K = 2ω/c.
So, by using (6) in the integral solution (1), we obtain the exact solution for any kind
of spectrum S(kz ):

4
s 
∞ 2
ω2 ω
X 
ψ(ρ, z, t) = a K exp[−iωt] Rn sinc  2 ρ2 + z + πn . (8)
n=−∞ c c
Equation (8), with Rn given by (7), resumes our scalar approach, i.e., it describes in
exact and analytic way any axially symmetric and totally propagating beam in free space
without boundaries. Therefore, it also describes nonparaxial propagating beams.
In this way, our problem now is reduced to the evaluation of the Fourier coefficients Rn
by using (7). Note that we can obtain these coefficients by any reasonable approximation
and that even so Eq.(8) still is an exact solution of the wave equation. Obviously, we
consider a finite number of terms in (7) and (8): in this case 2N + 1, with −N ≤ n ≤ N .
The requisite is that the chosen number N should allow a good representation of the
spectrum S(kz ) through the Fourier series.
Besides its simplicity, the present method presents another important advantage: in
general, the more nonparaxial is the desired beam (great values of ∆kz ), the faster the
convergence of the solution (8) which describes it. It is simple to understand this feature.
Usually nonparaxial beams are characterized by wide bandwidth spectra S(kz ), which in
general require a small number of terms in their Fourier series representation, and this fact
has as a consequence a small number of terms in solution (8) for a satisfactory description
of the correspondent beams (fast convergence). This property makes our method effective
when dealing with highly nonparaxial beams.

Example: As an example, we study a Gaussian spectrum, given by:

√b
(
π
exp[−b2 (kz − k̄z )2 ] , 0 ≤ kz ≤ ω/c
S(kz ) = (9)
0 , otherwise
where k̄z is the central position of the spectrum and b is a constant.
In (9), the bandwidth is given by ∆kz = 2/b. Here, k̄z and ∆kz will determine the
type of spectrum (paraxial or nonparaxial).

5
First, we obtain the Fourier coefficients Rn of the Gaussian spectrum substituting
(9) into (7). If the gaussian given by (9) possesses negligible values outside the interval
[0, ω/c], the Fourier coefficients can be easily calculated, resulting in

−π 2 n2
" #

 
Rn ≈ exp −i nk¯z exp , (10)
K K 2 b2
which are substituted into (8) to furnish an exact analytic solution of the wave equation
related to the spectrum (9).
For this example, we use a highly nonparaxial spectrum localized around k̄z = 0.4 ω/c
with a bandwidth ∆kz = 0.5 k¯z , and therefore b = 2/∆kz = 4/k¯z
Figure 1(a) shows a plot of the desired spectrum (9) by a red continuous line. It also
depicts the approximated spectra obtained through the Fourier series (6) for N = 2, 4, 6
and 10.

Figure 1: Nonparaxial scalar beam. (a) Nonparaxial Gaussian spectrum, eq.(9), repre-
sented by the continuous line and its representation through the Fourier series, eq.(6),
with N = 2, 4, 6 and 10; (b) The 3D intensity pattern of the resulting nonparaxial beam
given by eq.(8) with N = 10 and its orthogonal projection.

According to this graph, we can see that with a small number of Fourier coefficients

6
(i.e. 2N + 1 = 21) we can obtain a spectrum very similar to the original one in (9)
and therefore the same (small) number of terms in the summation of the solution (8) is
enough to represent very well the resulting highly nonparaxial beam, confirming the fast
convergence of our method in situations like this.
The 3D intensity pattern of the resulting nonparaxial beam, obtained from the so-
lution (8), with its orthogonal projection are shown in Fig.(1b). We can note the very
small transverse spot size ∆ρ ≈ λ. Another point that should be noted is the gaussian
shape with high localization in the longitudinal direction, a feature that is not found in
ordinary beams.

3 Vectorial (electromagnetic) case

In this section, based on our scalar approach, we develop two vectorial methods for
describing nonparaxial electromagnetic beams as exact solutions of the Maxwell equations.

3.1 First vectorial method: Obtaining nonparaxial totally prop-


agating electromagnetic beams with radial and azimuthal
polarizations

In this first vectorial approach we make use of the vector, A, and scalar, V , potentials.
Considering the free space (no currents and no charges) and the Lorenz gauge, we
have for the time harmonic case that ∇2 A + (ω 2 /c2 )A = 0, with V = −i(c2 /ω)∇ · A.
Now we can use our solution (8), with eq.(7), for any cartesian component of A.
In particular, we can use the component Az of vector potential written in cylindrical
coordinates (Aρ , Aφ , Az ).
Once we have the vector potential, the magnetic and electric fields can be obtained
through the well known relations:

7
B = ∇×A (11)

c2
E = i ∇(∇ · A) + iωA (12)
ω
This is the first vectorial method for obtaining exact nonparaxial beam solutions to the
Maxwell equations. We note, due to the fact that we choose the solution for any cartesian
component of A as being given by our scalar exact solution (8), that the calculation of
the fields is direct, involving only derivatives. This is an important feature of the present
method because it allows one to obtain exact solutions to the Maxwell equations without
the necessity of solving complex integrals.

3.1.1 Nonparaxial beams with radial polarization

~ = Eρ ρ̂ + Ez ẑ), we choose:
To obtain beams with radial polarization, (E

A(ρ, z, t) = ẑAz (ρ, z, t) = ẑA(ρ, z) exp(−iωt) (13)

where Az (ρ, z, t) is solution of the wave equation with azimuthal symmetry. The scalar
potential is given by (Lorenz gauge):

c2 ∂
" #
V (ρ, z, t) = −i Az (ρ, z, t) (14)
ω ∂z
By using equations (11,12) we get, in cylindrical coordinates,

Bρ = 0 (15)

Bφ = − Az (ρ, z, t) (16)
∂ρ
Bz = 0 (17)

and

8
c2 ∂ 2
Eρ = i Az (ρ, z, t) (18)
ω ∂ρ∂z
Eφ = 0 (19)
c2 ∂ 2
Ez = i Az (ρ, z, t) + iωAz (20)
ω ∂2z

Now, taking advantage of the solution found in the scalar case, we choose Az (ρ, z, t)
equal to the analytic solution obtained in (8). So, substituting (8) into (15-17) and (18-
20), we obtain the electric and magnetic field components of the electromagnetic beam
with radial polarization (i.e. Eφ = 0):


ω2
!
X sin h
Eρ = a K i 2 exp[−iωt] Rn (cnπ + ωz)ρ 3 5
c n=−∞ h
!
cos h sin h
−3 4 − 3 (21)
h h
Eφ = 0 (22)

"
X cos h sin h
Ez = a K iω exp[−iωt] Rn − + 3 −
n=−∞ h2 h
2
!#
ω sin h cos h sin h
ρ2 2 3 5 −3 4 − 3 (23)
c h h h
Bρ = 0 (24)
2 ∞
!
ω X sin h cos h
Bφ = −a K 2
exp[−iωt] Rn ρ − 2 (25)
c n=−∞ h3 h
Bz = 0 (26)

where Rn are the Fourier coefficients (related with the general spectrum S(kz )) given by
(7), a = 1 Tm and h is defined by:

s
2
ω2 2 ω

h= 2
ρ + z + πn (27)
c c

9
Example
To use the obtained solutions for radially polarized beams, we choose a rectangular
spectrum defined by:

1 1

= , for kzmin ≤ kz ≤ kzmax




kzmax − kzmin ∆kz
S(kz ) = (28)




0 , otherwise
where 0 ≤ kzmin ≤ kzmax ≤ ω/c. Here, the central spectrum position (k̄ = (kzmin +
kzmax )/2) and the spectral bandwidth (∆kz = kzmax − kzmin ) determine if the spectrum is
paraxial or nonparaxial. For this example we choose a nonparaxial case with k̄z = 0.5 ω/c
and ∆kz /k̄z = 1.20.
To obtain the Fourier coefficients Rn for the spectrum in (28), we use it in (7) to get

i 2π 2π
    
Rn = exp −i nkzmax − exp −i nkzmin (29)
2πn∆kz K K

Figure 2(a) shows the desired spectrum (28) in continuous line. It also depicts the
approximated spectra obtained through the Fourier series (6) for N = 80.
Substituting (29) into (21−26), we obtain the exact analytic solutions for the electric
and magnetic fields. These solutions describe the radially polarized nonparaxial electro-
magnetic beam in cylindrical coordinates. Figures 2(b)-(c) plot the 3D intensity patterns
of the longitudinal Ez and transverse Eρ components of the electric field, respectively.
The vectorial diagram of Eρ is shown in Fig. 2(d).
In Figures 2(b) and 2(c), we can see that the intensity of the resulting longitudinal
component is almost one order of magnitude higher than the transverse component of the
electromagnetic beam, i.e. |Eρ |2 < |Ez |2 . We can also see that the intensity spot size of
the Eρ component in Fig. 2(d) is of the same order of magnitude of the wavelength.
These properties are typical characteristics of highly nonparaxial electromagnetic

10
Figure 2: Radially polarized nonparaxial beam: (a) The Rectangular nonparaxial spec-
trum given by eq.(28) and its representation through eq.(6) with N = 80. (b)-(c) The 3D
intensity patterns for the transverse, Eρ , and longitudinal, Ez , components of the electric
field with their orthogonal projections on the plane z = 0. (d) The Vectorial diagram of
Eρ on the plane z = 0.

beams, making it clear why a simple scalar analysis is not sufficient to fully describe these
waves. The main reason is due to the common simplification that scalar analysis makes,
by disregarding the longitudinal component of the electric field. This point remarks the
importance of a full vectorial analysis where all the components of electromagnetic field

11
are taken into account in solving the Maxwell’s equations.

3.1.2 Nonparaxial beams with azimuthal polarization

We can obtain nonparaxial electromagnetic beams with azimuthal polarization (E =


Eφ φ̂) directly from the radial polarization just by using duality:



 E0 → c B

(30)
1
B0 → − E




c
In this way, from (21−26), we get

Eρ = 0 (31)
2 ∞
!
ω sin h cos h
X
Eφ = a K exp[−iωt] Rn ρ − 2 (32)
c n=−∞ h3 h
Ez = 0 (33)
2 ∞
!
ω X sin h
Bρ = −a K i exp[−iωt] Rn (cnπ + ωz)ρ 3
c3 n=−∞ h5
!
cos h sin h
−3 − 3 (34)
h4 h
Bφ = 0 (35)

ω2
"
ω cos h sin h
Rn − 2 + 3 − ρ2 2
X
Bz = −a K i exp[−iωt]
c n=−∞ h h c
#
sin h cos h sin h
(3 5 − 3 4 − 3 ) (36)
h h h

where Rn are the Fourier coefficients given in (7) and h was defined in (27).
As an example, we use again the rectangular spectrum defined in (28). Then the
Fourier coefficients, eq.(29), for this particular nonparaxial spectrum are to be used in
(31-36).

12
Figure 3(a) shows the 3D intensity pattern of the electric field E = Eφ φ̂ and its
orthogonal projection. Figure 3(b) shows the field intensity on the plane z = 0 and Figure
3(c) shows a zoom of the spot given in Fig.3(b), with a vectorial diagram superimposed.
One should observe that the transverse spot size has the same magnitude of the
wavelength. As we mentioned above, this is a typical characteristic of nonparaxial elec-
tromagnetic beams.
Finally, we note that in this first vectorial method, the expressions obtained for both
examples, for beams with radial and azimuthal polarizations, represent exact analytic
solutions of the Maxwell’s equations. This is an strong advantage with respect to other
methods that use approximated solutions or lengthy numerical simulations to obtain the
electromagnetic field components.

3.2 Second vectorial method: Nonparaxial propagating electro-


magnetic beams with linear polarization

In this section we develop a second vectorial method, which allows us to obtain


nonparaxial propagating electromagnetic beams with linear polarization.
Let us suppose that we wish to obtain a linearly polarized electromagnetic beam
~ = Ey ŷ + Ez ẑ with, as always, the propagation along the z direction.
E
~ ·E
Remembering that we are considering free space, the Gauss law (∇ ~ = 0) requires

that

∂ Z
Ez = − Ey dz (37)
∂y
Now, if we suppose that ψ is a solution of the wave equation, then its derivatives
with respect to the cartesian coordinates are solutions as well. Therefore, if we define
R
ψ = Ey dz, then we obtain:

13
Figure 3: (a) The 3D intensity pattern of the electric field of an Azimuthally polarized
nonparaxial beam obtained from eq.(32) and its orthogonal projection; (b) The beam
intensity pattern on the plane z = 0. A zoom with a vectorial diagram of the electric field
on z = 0 is shown in (c).


Ey = ψ (38)
∂z

Ez = − ψ (39)
∂y
Note that in (38) and (39), we obtain the components of the electric field (Ey and
Ez ) from the solution ψ of the scalar wave equation.

14
Now, we can take advantage of our scalar solution which describes nonparaxial scalar
beams, i.e., we define ψ as given by eq.(8) and then substitute it into (38) and (39). With
this, we get the electric field solution:

Ex = 0 (40)
 ∞ !
ω X ω cos[h] sin [h]

Ey = a K exp[−iωt] Rn (nπ + z) −
c n=−∞ c h2 h3
(41)
 ∞ !
ω X ω sin[h] cos[h]

Ez = a K exp[−iωt] Rn ρ sin φ − , (42)
c n=−∞ c h3 h2

where Rn are the Fourier coefficients given by eq.(7), h was defined in eq.(27) and a = 1 V.
Note that the expressions (40 − 42) are exact analytic solutions representing the
transverse and longitudinal components of the electric field. The magnetic field can be
found through the Faraday law, B = −i/ω ∇ × E.
Example:
For the last example, we use the Gaussian nonparaxial spectrum defined in (9). There-
fore, we use the Fourier coefficients given by (10) into (40−42) to obtain the exact ex-
pressions for the electric field components (Ey and Ez ).
The 3D intensity pattern of the longitudinal electric field component, Ez , and its
orthogonal projection on the plane z = 0, are showed in Fig.4(a).
Figure 4(b) depicts the intensity pattern for the transverse electric field component,
Ey , on the plane z = 0 and its orthogonal projection. A zoom of Fig. 4(b) with a
vectorial diagram is plotted in graph 4(c). We note in this last figure that the spot size
is approximately equal to the wavelength.
We can also see in the Figures 4(a) and 4(b) that the maximum intensity of the
resulting longitudinal component of the electric field is grater than that of the transverse
component. This clearly invalidates scalar analysis and enforces the necessity of a full

15
Figure 4: (a) The 3D intensity pattern of the longitudinal electric field component, Ez , on
the plane z = 0 and its orthogonal projection; (b) The same for the transverse component
Ey ; (c) A vectorial diagram of Ey on the plane z = 0.

vectorial treatment, where all components of the electromagnetic field are taken into
account in solving the Maxwell’s equations.

16
4 Conclusions

The first part of this work is devoted to a method capable to describe, in exact and
analytical way, any totally propagating scalar beam with azimuthal symmetry, including,
as particular cases, nonparaxial beams.
Subsequently, based on this scalar approach, we have developed two vectorial meth-
ods for describing nonparaxial electromagnetic beams as exact solutions of the Maxwell
equations.
Compared to others, our methods have the advantage of being faster (fast conver-
gence) when dealing with highly nonparaxial beams. Another advantage is the math-
ematical simplicity of our approaches: the only necessary integrations are those to
find simple Fourier coefficients of the longitudinal wavenumber spectrum, S(kz ) (with
−ω/c ≤ kz ≤ ω/c), of the desired beam, all the other calculations are based on simple
derivatives.
As examples we have obtained exact analytical solutions of the Maxwell equations
describing nonparaxial electromagnetic beams with radial, azimuthal and linear polariza-
tions.
These new methods can be applied to describe beams of interest to optical tweezers,
atom guiding, high-resolution microscopy, etc..

5 Acknowledgements

The authors are very grateful to Jose Luis Prego for continuous discussions and kind
collaboration.
This work was supported by CNPq and FAPESP (Brazil).
E-mail addresses: roger.garay.a@gmail.com and mzamboni@dmo.fee.unicamp.br

17
References

[1] Michel Zamboni-Rached and E. Recami, “Sub-luminal wave bullets: exact localized
subluminal solutions to the wave equations,” Phys. Rev. A 77, 033824 (2008);

[2] A. Carnicer, I. Juvells, D. Maluenda, R. Martı́nez-Herrero, and P. M. Mejı́as: “The


longitudinal component of paraxial fields”, European Journal of Physics, vol.33, no.
5, p. 1235 (2012).

[3] A. April: “Nonparaxial TM and TE beams in free space”, Opt. Lett., vol.33, no.14,
pp. 1563-1565 (2008).

[4] R. Borghi, A. Ciattoni and M. Santarsiero: “Exact axial electromagnetic field for
vectorial gaussian and attened gaussian boundary distributions”, J. Opt. Soc. Am.
A, vol.19, no.6, pp. 1207-1211 (2002).

[5] M. Lax, W. H. Louisell, and W. B. McKnight: “From maxwell to paraxial wave


optics”, Phys. Rev. A, vol.11, pp. 1365-1370 (1975).

[6] P. C. Chaumet: “Fully vectorial highly nonparaxial beam close to the waist”, J. Opt.
Soc. Am. A, vol.23, no.12, pp. 3197-3202 (2006).

[7] A. Ciattoni, B. Crosignani, and P. D. Porto: “Vectorial analytical description of


propagation of a highly nonparaxial beam”, Optics Communications, vol.202, pp.
17-20 (2002).

[8] A. April: “Bessel-gauss beams as rigorous solutions of the helmholtz equation”, J.


Opt. Soc. Am. A, vol.28, no.10, pp. 2100-2107 (2011).

18
[9] S. Nemoto: “Nonparaxial gaussian beams”, Appl. Opt., vol.29, no.13, pp. 1940-1946
(1990).

[10] J. E. Curtis, B. A. Koss, and D. G. Grier: “Dynamic holographic optical tweezers”,


Optics Communications, vol.207, pp. 169-175 (2002).

[11] J. Arlt, V. Garces-Chavez, W. Sibbett and K. Dholakia: “Optical micromanipulation


using a bessel light beam”, Optics Communications, vol.197, pp. 239-245 (2001).

[12] T. Ito and S. Okazaki: “Pushing the limits of lithography”, Nature, vol.406, no.6799,
pp. 1027-1031 (2000).

[13] Besieris, I. M., A. M. Shaarawi, and R. W. Ziolkowski, “A bidirectional traveling


plane wave representation of exact solutions of the scalar wave equation,” J. Math.
Phys., Vol. 30, 1254-1269 (1989).

[14] Ioannis Besieris and Amr Shaarawi, “Paraxial localized waves in free space,” Opt.
Express 12, 3848-3864 (2004).

[15] M. Zamboni-Rached, E. Recami, H. Figueroa “New localized Superluminal solutions


to the wave equations-with finite total energies and arbitrary frequencies”, European
Physics Journal D, Vol. 21, 217-228 (2002).

[16] M. Zamboni-Rached, “Unidirectional decomposition method for obtaining exact lo-


calized wave solutions totally free of backward components,” Physical Review A, Vol.
79, 013816 (2009).

[17] M.Zamboni-Rached: “Localized Waves: Structure and Applications”, M.Sc. thesis


(Phys. Dept., State University of Campinas, 1999);

19
[18] M.Zamboni-Rached, “Localized waves in diffractive/dispersive media”, PhD
Thesis, Aug.2004, (State University of Campinas, DMO/FEEC [download at
http://libdigi.unicamp.br/document/?code=vtls000337794].

[19] Localized Waves: Theory and Applications, ed. by H.E.Hernández-Figueroa,


M.Zamboni-Rached, and E.Recami (J.Wiley; New York, 2008).

[20] I.S.Gradshteyn and I.M.Ryzhik: Integrals, Series and Products, 5th edition
(Ac.Press; New York, 1965).

[21] Classical Electrodynamics, ed. J. D. Jackson ( 3rd ed. Wiley, pp. 239-242, 1998).

20

View publication stats

You might also like