Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Propagation of generalized vector Helmholtz-Gauss beams through paraxial optical systems

Open Access Open Access

Abstract

We introduce the generalized vector Helmholtz-Gauss (gVHzG) beams that constitute a general family of localized beam solutions of the Maxwell equations in the paraxial domain. The propagation of the electromagnetic components through axisymmetric ABCD optical systems is expressed elegantly in a coordinate-free and closed-form expression that is fully characterized by the transformation of two independent complex beam parameters. The transverse mathematical structure of the gVHzG beams is form-invariant under paraxial transformations. Any paraxial beam with the same waist size and transverse spatial frequency can be expressed as a superposition of gVHzG beams with the appropriate weight factors. This formalism can be straightforwardly applied to propagate vector Bessel-Gauss, Mathieu-Gauss, and Parabolic-Gauss beams, among others.

©2006 Optical Society of America

1. Introduction

The problem of finding vector beam solutions ofMaxwell equations has been studied by several researchers [1, 2, 3, 4, 5, 6, 7, 8]. In this direction, the existence of the vector Helmholtz-Gauss (vHzG) beams, which constitute a general family of localized electromagnetic beams, was demonstrated theoretically in a recent paper [8] for propagation in free space. Special cases of the vHzG beams are the transverse electric (TE) and transverse magnetic (TM) Gaussian vector beams [2], nondiffracting vector Bessel beams [3], polarized Bessel-Gauss [4], Mathieu-Gauss, and Parabolic-Gauss beams, modes in cylindrical waveguides and cavities [9, 10], and scalar Helmholtz-Gauss beams [11, 12].

In this paper, we introduce a useful generalized form of the vHzG beams that we will refer to as generalized vHzG (gVHzG) beams. The paraxial propagation of the gVHzG beams is studied, not only in free space, but also through more general types of paraxial optical systems characterized by complex ABCD matrices, including lenses, Gaussian apertures, cascaded paraxial systems, and systems having quadratic amplitude as well as phase variations about the axis. By following a coordinate-free approach, rather than proposing solutions in a particular coordinate system, it was possible to derive an elegant and closed-form expression for the electromagnetic field, the vector angular spectrum, and the Poynting vector at the output plane of the ABCD system. It is found that the gVHzG beams are a class of vector fields which exhibit the property of form-invariance under paraxial optical transformations. The formulation described in this paper can be useful in applications where the polarization of the fields is of major concern [5, 6, 13, 14, 15].

2. Propagation of the generalized vector HzG beams

Consider an electromagnetic paraxial beam with time dependence exp(-iωt) travelling in the z direction (unit vector ) through an ABCD axisymmetric optical system with input and output planes located at z=z 1 and z=z 2, respectively. The system is characterized by an ABCD matrix with, in general, complex elements A, B, C, and D that satisfy the unimodular relation AD-BC=1.

Let us express both the position vector R and the wave vector K as

R=(r,z),r=(x,y)=(r,θ);K=(k,kz),k=(kx,ky)=(k,ϕ),

where r and k are the positions at the transverse planes of the configuration and frequency spaces, respectively. Additionally, we will denote a general vector field as F=f+fz , where f=(fx, fy ) represents the transverse part of the field. The transverse nabla operators in the configuration and frequency spaces are denoted by ∇=∂/∂x+y ∂/y and ∇=k ̂x∂/∂kx + y∂/∂ky , respectively.

Throughout the paper, the suffixes “1” and “2” denote the evaluation of the respective physical quantity (or operator) at the input (z=z 1) and output (z=z 2) transverse planes of the ABCD system.

2.1. Definition of the generalized vector HzG beams

We commence the analysis by defining the transverse electric e(r) and magnetic h(r) vectors of a first-class (TM) gVHzG beam at the input plane of the ABCD system as [8]

e1(r1)=exp(iKr122q1)1W(r1;κ1),h1(r1)=(εκ)12ẑ×e1(r1),

where K=|K|=ω(µε)1/2 is the wave number.

The electric vector field in Eqs. (2) results from the product of two functions, each depending on one parameter. First, the Gaussian modulation is characterized by a complex beam parameter q 1=q1+iq1, where the superscripts 𝓡 and 𝓘 denote the real and imaginary parts of a complex quantity, respectively. For simplicity in dealing with the ABCD system, the parameter q 1 is used throughout the paper. The physical meaning of q 1 is contained in the known relation 1/q 1=1/ 1+i2/Kw12, where ℜ1 is the radius of curvature of the phase fronts, and w 1 is the 1/e amplitude spot size of the Gaussian modulation. In assuming a complex q1 we are allowing for the possibility that the Gaussian apodization has an initial converging (q1<0) or diverging (q1>0) spherical wavefront. Additionally, the condition q𝓘 1<0 must be fulfilled in order to satisfy the physical requirement that the field amplitude vanishes as r becomes arbitrarily large.

The gradient ∇1 W(r 1;κ 1) in Eq. (2) provides the vector nature of the transverse electric field. The scalar function W(r 1;k 1) is a solution of the two-dimensional Helmholtz equation [12+κ12]W=0, and physically describes the transverse shape of an ideal scalar nondiffracting beam [16, 22, 23]. Since a vector beam e 1(r 1) is determined from a scalar function W(r 1;κ 1), throughout the paper we will refer to the latter as the seed function. The function W can be formally expanded in terms of plane waves as

W(r1;κ1)=ππg(ϕ)exp[iκ1(x1cosϕ+y1sinϕ)]dϕ,

where κ 1 and g(ϕ) are the transverse wave number and the angular spectrum of the ideal scalar nondiffracting beam, respectively. Since g(ϕ) is arbitrary, an infinite number of profiles can be obtained, see section 2.6 for important special cases.

The fields in Eq. (2) are purely transverse and correspond to the zeroth-order electric and magnetic fields of the perturbative series expansion of the Maxwell equations provided by Lax et al. [1]. The Lax expansion also showed that in next-order correction small longitudinal field components must be present, and their values are obtained from the transverse components through

ez,1=iK1·e1=exp(iKr122q1)(iκ12KW+1W·r1q1),
hz,1=iK1·h1=εμexp(iKr122q1)(ẑ×1W)·r1q1.

TE polarized gVHzG beams: For the sake of space, throughout the paper we shall deal with the explicit expressions for the TM beams. However second-class (TE) beams can be readily obtained from Eqs (2) by applying the duality property, i.e. replacing E with (µ/ε)1/2 H and (β/ε)1/2 H with -E, namely

e1(TE)(r1)=exp(iKr122q1)[ẑ×1W],h1(TE)(r1)=εμexp(iKr122q1)1W.

The corresponding longitudinal components are obtained by applying the divergence operator as shown in Eqs. (4).

2.2. Classification of the generalized vector HzG beams

In the theory of nondiffracting beams, the parameter κ 1 in Eq. (3) is customarily assumed to be real and positive [16, 22, 23]. For the sake of generality, we will let κ 1=κ1+iκ1 be arbitrarily complex allowing the possibility of having three kinds of beams:

Ordinary VHzG (oVHzG) beams correspond to purely real κ 1=κ1 for which the seed function W(r 1;κ 1) is a two-dimensional purely oscillatory (or standing wave) function. The physical meaning of κ1 is clear, as it governs the oscillatory behavior of the function W in the transverse direction. Far from the z axis, the spatial period of the field oscillations tends monotonically to 2π/κ1. The special case when κ1>0 and q 1=-iKw12/2 is purely imaginary leads to the ordinary vector solutions discussed in Ref. [8], where w1 is the standard 1/e amplitude spot size of the Gaussian apodization.

Modified VHzG (mVHzG) beams correspond to purely imaginary κ 1=iκ1. In this case W(r 1,κ1) is an evanescent seed function which satisfies the modified Helmholtz equation [12-(κ1)2]W=0. Two concrete examples of seed functions of the modified kind are provided by the cosh-Gaussian beams [17] and the modified Bessel beams [18, 19].

Generalized VHzG (gVHzG) beams correspond to the general case when k 1 is complex. As we will see, the gVHzG can be interpreted as intermediate solutions between oVHzG and mVHzG beams.

In order that fields in Eq. (2) satisfy the paraxial approximation, it is needed that K≫1/w 1, i.e. the Gaussian spot width is many wavelengths wide, and additionally that K≫|κ 1|, i.e. the spatial transverse beam oscillations must be many wavelengths wide. Two limiting cases of the gVHzG beams are of particular interest: (a) pure vector nondiffracting beams are obtained when |q1|→∞, and (b) the special value κ 1=0 leads to the generalized vector Laplace-Gauss beams, for which the seed functionW is now a solution of the 2D Laplace equation [8].

2.3. Propagation of the electromagnetic field through the ABCD system

Paraxial propagation of the electric vector field e 1(r 1) through the ABCD system can be performed by solving the Huygens diffraction integral [20, 21]

e2(r2)=Kexp(iKL0)i2πBe1(r1)exp[iK2B(Ar122r1·r2+Dr22)]d2r1,

where e 2(r 2) is the output transverse electric field, and L 0 is the optical path length from the input to the output plane of the ABCD system measured along the optical axis. The vector integral in Eq. (6) can be treated as a pair of independent scalar integrals by decomposing the transverse vector e 1(r 1) into orthogonal linearly polarized parts. After substituting each Cartesian component of Eq. (2) and the expansion given by Eq. (3) into Eq. (6), the integration can be performed applying the changes of variables xj=uj cosϕ-vj sinϕ, and yj=uj sinϕ+vj cosϕ, for j=1,2. Upon returning to the original variables and regrouping the Cartesian components into a vector function, we obtain (see the detailed derivation in the appendix A)

e2(r2)=κ1κ2exp(iκ1κ2B2K)G(r2,q2)2W(r2;κ2),
h2(r2)=(εμ)12ẑ×e2(r2),

where

G(r2,q2)=exp(iKL0)A+Bq1exp(iKr222q2),

is the output field of a scalar Gaussian beam with input parameter q 1 travelling axially through the ABCD system, and the transformation laws for the parameters q 1 and κ 1 from the plane z 1 to the plane z 2 are

q2=Aq1+BCq1+D,κ2=κ1q1Aq1+B.

Equations (7) are the main result of this paper, they permit an arbitrary gVHzG beam to be propagated in closed-form through an axisymmetric ABCD optical system with real or complex matrix elements. The presence of the fundamental Gaussian beam G(r 2,q 2) in Eq. (7a) provides the confinement mechanism which ensures the transverse intensity distribution vanishes at large values of r, and that the beam is square integrable. As expected, Eqs. (7) reduces to Eq. (2) when the ABCD matrix becomes the identity matrix.

Like the input field [Eqs. (4)], the longitudinal components of the electric ez ,2 and the magnetic hz ,2 fields at the output plane can be readily obtained by applying the operator [(i/K)∇2·] over the corresponding transverse components e 2(r 2) and h 2(r 2), respectively.

2.4. Poynting vector of the generalized vector HzG beams

The time-averaged Poynting vector is given by 〈S〉=Re(E×H*)/2. It can be decomposed as 〈S〉=〈sz ẑ+〈s〉, where 〈sz 〉 is the longitudinal part which determines the flow of energy in the direction of propagation z, and 〈s〉 is the transverse part which determines the flow of energy perpendicular to this direction. The Poynting vector can be calculated at the input (j=1) and output (j=2) planes using the corresponding expressions for the electric and magnetic fields, we have

Sj=12(εμ)12fj2{jW2ẑ+Re[(iκj2KW+jW·rjqj)jW*]},

where f 1=exp(iKr12/2q1 ), and f 2=(κ12 )exp(-1κ2B/2K)G(r 2,q 2).

From Eq. (10) we note that the energy flux density along the longitudinal direction is proportional to the squared magnitude of the transverse electric field vector. SinceW and qj are, in general, complex, the beam exhibits a transverse flow of energy whose radial part is a manifestation of diffraction. For a paraxial beam it is expected that the longitudinal part of the energy flux be much more significant than the transverse part, in fact a simple analysis of orders of magnitude in Eq. (10) reveals that the longitudinal flow is at least Kw 1 times the transverse one. Finally, for a lossless medium, the light power carried by the beam in longitudinal direction ∬〈S j·ẑd2 r j remains constant for any z plane.

2.5. Propagation of the vector angular spectrum

The electric field e(r) of the gVHzG beam at either the input or output planes of the ABCD system admits the plane wave expansion of the form e(r)=(1/2π)∬∞-∞ẽ(k)exp(i k·r)d2 k, where ẽ(k) is the vector angular spectrum whose functional form is obtained by Fourier inversion

e˜(k)=12πe(r)exp(ik·r)d2r.

After inserting Eqs. (2) and (7a) into Eq.(11) and performing the integrals we obtain the vector angular spectra at the input and output planes of the ABCD system, namely

e˜1(k1)=iexp(iq1κ122K)exp(iq1k122K)˜1W(k1;q1κ1K),
e˜2(k2)=κ1Kκ2q2exp(iq2κ22+κ1κ2B2K)G˜(k2,q2)˜2W(k2;Bκ22Kκ1),

where ˜ =k ̂x∂/∂kx + y∂/∂ky is the transverse nabla operator in the K space, and

G˜(k2,q2)=(iq2K)exp(iKL0)A+Bq1exp(iq2k222K)

is the Fourier transform of the Gaussian function G(r 2,q 2) in Eq. (8).

2.6. Remarks on the coordinates systems and polarization basis

The electromagnetic fields in Eqs. (2), (7) and (12) are completely general in the sense that they do not depend on a particular coordinate system. The vector beam solutions are constructed starting from scalar solutions of the 2D Helmholtz equation which can be formally expanded in terms of plane waves according to Eq. (3). Although this integral expansion constitute a general integral solution, it is important to note that the 2D Helmholtz equation can also be solved in several orthogonal coordinate systems using the separation of variables method [11, 24, 25]. This fact leads to have complete and orthogonal families of eigenfunctions of the 2D Helmholtz equation.

Of particular interest are the families of eigenfunctions of the 2D Helmholtz equation expressed in Cartesian, polar, elliptic, and parabolic coordinates [11]. For Cartesian coordinates (x,y), generalized vector Gaussian beams can be constructed from superpositions of fundamental plane waves of the form W=exp(iκ⃗1·r 1)=exp[ 1(x cosκ+ysinκ)], see, for example, the cosine-Gauss beams studied in Ref. [11]. The case of the polar coordinates (r,θ) corresponds to eigenfunctions W=Jem(κr)exp(imθ) for which gVHzG beams reduce to the mth-order generalized vector Bessel-Gauss beams [4]. For elliptic coordinates (ξ,η), generalized vector even Mathieu-Gauss beams of mth-order and ellipticity parameter ε can be constructed from the eigenfunctions W=Jem(ξ,ε)cem(η,ε), where Je(·) and ce(·) are the radial and angular even Mathieu functions of mth-order and parameter ε, respectively [22]. For parabolic coordinates (u,v), generalized vector even parabolic-Gauss beams can be constructed from the eigenfunctions W=Pe(u2κ;p)Pe(v2κ;p), where Pe(·) is a parabolic cylinder function of parameter p and even parity [23].

On the other hand, the gradient operator in Eqs. (2) and (7a) can also be expressed in several coordinate systems [24, 25], with the consequence that several polarization basis may be used to decompose the field vector e 2(r 2) at any point r2 into two orthogonal polarized transverse parts. For instance, in polar coordinates, the transverse vector fields can be split into radial and azimuthal polarized components, or in elliptic coordinates, into elliptic and hyperbolic polarized components. Explicit expressions for these polarization basis associated to particular coordinate systems can be found elsewhere [24, 25].

It is important to emphasize that a general vector beam solution is found by the superposition of TM and TE vector modes, i.e.

E=αE(TM)+βE(TE),

where α and β are arbitrary constants. By combining the different polarization basis with the different seed functions W(r 2;κ 2) a large variety of beam profiles with specific polarization states could be constructed through superposition. As example, consider the basis of circular polarizations uû±=(±i ŷ)/√2. It is easy to verify that the gradient of W in Eqs. (2) and (7a) can be expressed in the basis of vectors û± as

W=12(WxiWy)û++12(Wx+iWy)û.

Now, gVHzG beams with pure left (+) and right (-) circularly polarized beams can be constructed from Eq. (15) using the superposition e ±=e (TM)±ie(TE) at either the input (j=1) or output (j=2) planes of the ABCD system, we have explicitly

ej±(rj)=2fj(WxjiWyj)û±,

where f 1=exp(iKr12/2q 1), and f2=(κ 1/κ 2)exp(- 1 κ 2 B/2K)G(r 2,q 2). A similar approach can be applied for the vector angular spectra in Eq. (12) to derive the corresponding circular polarization states.

Finally, from Eqs. (2) and (5), it is clear that the polarization of the transverse electric field of the TM and TE beams is defined entirely by the operations ∇W and ×∇W, respectively. Both vector fields Ψ(1)=∇W and Ψ(2)=×∇W constitute two independent vector solutions of the 2D vector Helmholtz equation ∇2Ψ+κ 2Ψ=0 [24, 25]. Now, if we set the function W=Wm to be the m-th eigensolution belonging to a countable set of complete and orthogonal solutions of the scalar Helmholtz equation, then, because of the linearity and the one-to-one mapping of the gradient operator, the properties of linear independence, orthogonality, and completeness exhibited by the family of scalar solutions Wm are transferred to the corresponding families of vector fields Ψ(1) and Ψ(2). Additionally, the transverse fields of a TM and a TE beams are orthogonal, even when their seed functions Wm are equal. In this sense, the gVHzG beams exhibit similar polarization properties than the ideal vector nondiffracting beams [3, 7, 16] and waveguides with constant cross-section [9, 10, 24].

3. Physical discussion of the propagation properties

In Section 2 we have demonstrated that localized vector beam solutions of the Maxwell equations can be propagated through an ABCD optical system in a closed and coordinate-free form. Particular attention was focused on the propagation of the vector beams and their vector angular spectra. Several considerations for the coordinate system and polarization basis were also discussed. To get involved in the propagation details of the gVHzG beams, let us study in this section two very illustrative examples.

3.1. Free space propagation

We consider first the free space propagation along a distance L=z 2-z 1. The input and output fields are given by Eqs. (2) and (7a) with A=1, B=L, C=0, and D=1. From Eqs. (9) the transformation laws become

q2=q1+L,κ2=κ1q1q1+L,

where we note that the product q 2 κ 2=q 1 κ 1 remains constant under free-space propagation.

 figure: Fig. 1.

Fig. 1. Physical picture of the decomposition of a gVHzG beam propagating in free space in terms of fundamental vector Gaussian beams whose mean propagation axes lie on the surface of a double cone.

Download Full Size | PDF

In a similar fashion to the scalar HzG beams [11], to gain a basic understanding of the features of the gVHzG beams propagating in free space, one may consider that a gVHzG beam is formed as a superposition of fundamental vector Gaussian beams (see Fig. 1) whose mean propagation axes lie on the surface of a double cone, whose amplitudes are modulated angularly by the function g(ϕ). This physical picture is evident after replacing Eqs. (17) into Eq. (7a) and observing that the TM polarized gVHzG beams in vacuum can be rewritten as

e2(r2)=ππg(ϕ)g2(r2;ϕ)dϕ,

where

g2(r2;ϕ)=iexp(iκ122KLζ)exp(iKL)ζexp(iKr222q1ζ)exp(iκ1·r2ζ)κ1,

with ζ≡1+L/q 1 and κ 11(cosϕ +sinϕ ŷ) Equation (19) represents the free space propagation along a distance L of a tilted Gaussian beam with input parameter q 1 whose mean wave vector has a projection κ 1 over the transverse plane [11], and whose polarization vector points in direction of the vector κ 1.

The generatrix of the double cone shown in Fig. 1 corresponds to the linear propagation of the centroid of the individual Gaussian beams, and from Eqs. (7a) it is found to be

rgen(z)=q12κ1𝒥Kq1𝒥+κ1𝓡q1𝒥+κ1𝒥q1𝓡Kq1𝒥(zz1).

In Fig. 1 we identify three important transverse planes:

• The initial plane at z=z 1.

• The waist plane (z=zwaist ) corresponds to the plane where the width of the elementary Gaussian beams is minimum, i.e. where the radial factor exp (iKr22/2q2) becomes a real Gaussian envelope. Using this condition, from Eqs. (17) we get zwaist=z1q1. At the waist plane the parameter q becomes purely imaginary qwaist=iq1, whereas the parameter κ reduces to κwaist 1(1-iq1/q1). From the general expression of the Poynting vector Eq. (10) we note that if W and κ is set to be purely real, then at the waist plane the energy flow becomes purely longitudinal.

• The vertex plane (z=zvertex ) corresponds to the plane where the main propagation axes of the constituent Gaussian beams intersect. As shown in Fig. 1, the pseudo-nondiffracting region delimits the region where significant interference of the constituent vector Gaussian beams occurs, and where the transverse beam profile exhibits a standing-wave behavior. The evaluation of the condition rgen =0 in Eq. (20) yields

zvertex=z1q12κ1𝒥κ1𝓡q1𝒥+κ1𝒥q1𝓡.

Note that at the vertex plane the parameter κ becomes purely real κvertex=κ1+κ1 q1/q1, with the consequence that at this plane the beam profile belongs to the oVHzG kind with qvertex 1 q 1/κvertex . At the vertex plane the extent of the pseudo-nondiffracting region is maximum, and its 1/e amplitude Gaussian spot size can be calculated with wvertex =[KIm(1/qvertex )/2]-1/2.

In general, the initial, the waist, and the vertex planes are located at different axial positions, as shown in Fig. 1. The ordinary VHzG beams studied in Ref. [8] constitute the special case when z 1=0, q𝓡 1=0, and κ1=0, for which the three planes coincide at z=0, and the cone generatrix reduces to the expected rgen =(κ𝓡 1/K)z.

On the other side, the mVHzG beams occur when κ1=0; if we additionally set q1=0 then rgen =r 0=q1 κ1/K becomes a constant, and therefore the mVHzG beams may be viewed as a superposition of vector Gaussian beams whose axes are parallel to the z axis and lie on the surface of a circular cylinder of radius r 0.

Finally, we remark that the transverse fields of the gVHzG beams propagating in free space satisfy the paraxial wave equation [12+i2K∂/∂z]{e 1,h 1}=0 and correspond to the purely transverse zeroth-order electric and magnetic fields of the perturbative series expansion of the Maxwell equations provided by Lax et al. [1].

3.2. Propagation through a GRIN medium

Let us consider now the propagation of the gVHzG beams through a graded refractive-index (GRIN) medium with quadratic index variation n(r)=n 0(1-r 2/2a 2). The ABCD transfer matrix from plane z 1 to plane z 2=z 1+L is given by

[ABCD]=[cos(La)asin(La)sin(La)acos(La)].

For a general input vector field of the form (2), the propagated vector field at a distance z 2 is described by Eq. (7a). Substitution of the matrix elements in Eq. (22) into Eqs. (9) yields the parameter transformations:

q2=aq1cos(La)+asin(La)q1sin(La)+acos(La),κ2=κ1q1q1cos(La)+asin(La),

where we note that, under propagation, the parameters q and κ vary periodically with a longitudinal period 2πa, therefore, the initial field distribution self-reproduces after a distance 2πa.

To show the role played by the gVHzG beams as intermediate vector solutions between oVHzG and mVHzG beams, let us assume that the input field at z 1=0 belongs to the oVHzG kind (i.e. κ1=0) with an initial real Gaussian apodization of width w1 (i.e. q 1=iq1=-iKw12/2). For a propagation distance L=LF=πa/2, the ABCD matrix Eq. (22) reduces to [0,a;-1/a,0] which is indeed identical to the matrix transformation from the first to the second focal plane of a converging thin lens of focal length a, i.e. a Fourier transformer. At the Fourier plane L=LF , from Eqs. (23) we see that both parameters q 2=ia 2/q𝓘 1 and κ 2=iκ1 q1/a become purely imaginary. It is now evident that if an oVHzG profile is Fourier transformed, a mVHzG profile will be obtained, and vice versa. The intermediate profiles belong to the gVHzG kind where, for the particular case of the GRIN medium, the transition between both types of beams is characterized by the continuous transformations given in Eqs. (23).

The special case when the parameter a of the GRIN medium is equal to the Rayleigh distance (i.e. zR=Kw12/2) of the initial Gaussian apodization is of particular interest. From Eqs. (23) we see that the Gaussian width q 2=q 1=-ia remains constant under propagation and that the wave number κ 2=κ 1 exp(-iL/a) rotates at a constant rate over the complex plane (κ2,κ2) as the beam propagates through the GRIN medium. For brevity, we will refer to the case when a=Kw12/2 as a balanced propagation, and a non-balanced otherwise.

In Figs. 2 and 3 we show the propagation of the transverse intensity distribution and the electric vector field for several circularly polarized gVHzG beams with κ 1=30 mm-1 through a GRIN medium with a=1/√2π m. The input fields are given by Eq. (16) with j=1 for K=2π/λ and λ=632.8 nm. The animations were constructed by calculating the field at 200 transverse planes evenly spaced from the input (z 1=0) to the output (z 2=4LF =2πa) planes using Eq. (16) with j=2 to generate the left or right circularly polarized fields as the case may be.

The fields shown in Fig. 2 correspond to a seed function W (r 1) given by the superposition of N plane waves of the general form

W(r1)=n=1NAnexp[iκ1r1cos(θ1ϕn)],

where An are complex amplitudes. For Fig. 2(a) we have chosen a left circularly polarized oVHzG beam in a balanced condition (q 1=-ia=-i/√2π) with N=3, An={1,1,1}, and ϕn ={90°,-30°,-150°}. Note that the width of the constituent Gaussian beams remains constant under propagation because the beams are balanced. Following the established convention, at a given z plane, the transverse components of the fields rotate anti-clockwise for left-handed circular polarization as time increases. The field at the plane z 2=LF is shown Fig. 2(b), where we note that for the selected amplitudes An =1 the beam polarization becomes purely radial. To show the non-balanced condition, in Fig. 2(c) we propagated the same gVHzG by setting now q 1=0.4-i0.8/√2π and keeping all remaining parameters unchanged. The video shows clearly that the width of the constituent Gaussian beams change under propagation and reach a minimum at the plane where q 2 in Eqs. (23) becomes purely imaginary (~1.22LF )

In Fig. 2(e) we show a right circularly polarized balanced oVHzG beam constructed with N=8 constituent Gaussian beams. By means of the amplitudes and phases of the coefficients An it is possible to adjust the polarization state of the resulting beam. In this case we set An=i such that the electric field at the plane z 2=LF now becomes purely azimuthal, as shown in Fig. 2(f). In Fig. 2(g) we set An =exp(-iπn/4) such that the electric field vectors at each point on the plane z 2=LF become parallel, as shown in Fig. 2(h).

Figures 3(a) to 3(d) show the vector propagation of ordinary vector Bessel-Gauss (VBG)

beams with even parity. The first propagation shown in Fig. 3(a) corresponds to a balanced input VBG beam of the form in Eq. (16) with seed function W (r 1)=iJ 3 (κ 1 r 1)cos (3θ 1). The presence of the factor i in the seed function produces that, at the plane z 2=LF , the VBG beam be azimuthally polarized and belongs to the modified VHzG kind. The second propagation shown in Fig. 3(c) corresponds to a non-balanced VBG beam withW (r 1)=J 3 (κ 1 r 1)cos (3θ 1) and complex input parameter q 1=0.4-i0.8/√2π.

The propagation of a fourth-order helical vector Mathieu-Gauss (VMG) beam is shown in Figs. 3(e) and 3(f). The seed function is given by the superposition of even and odd Mathieu beams [11, 22], namely W (r 1)=Je4(ξ,3)ce4(η,3)+iJo4(ξ,3)se4(η,3), where (ξ,h) are the elliptic coordinates defined as x=f cosh ξ cos η and y=f sinh ξ sin η, with f being the semifocal distance. In this case the vector beam is balanced but the initial field belongs to the gVHzG kind with κ 1=30+i15 mm-1. The input field is given by Eq. (16) where the Cartesian partial derivatives are expressed in elliptic coordinates as follows

x=1f(cosh2ξcos2η)(sinhξcosηξcoshξsinηη),
y=1f(cosh2ξcos2η)(coshξsinηξsinhξcosηη),

As the beam propagates, the parameters q and κ vary according to Eq. (23). For this value of κ 1, the typical elliptic annular intensity pattern of the ordinary helical VMG beams occurs approximately at z⋍0.28LF , while the expected circular annular pattern of the modified helical VMG beams occurs at z⋍1.28LF .

 figure: Fig. 3.

Fig. 3. Propagation of the transverse intensity distribution and electric vector field for generalized vector Bessel-cosine-Gauss, Mathieu-Gauss, and parabolic-Gauss beams. The parameter data for the propagations are included within the text. The movies show the evolution from z=0 to z=4LF . (Movie files: 3.1 MB, 2.5 MB, 3.6 MB, and 2.2 MB)

Download Full Size | PDF

 figure: Fig. 2.

Fig. 2. Propagation of the transverse intensity distribution and the electric vector field for circularly polarized gVHzG beams constructed with finite superposition of vector Gaussian beams. The parameter data for the propagations are included within the text. The movies show the evolution from z=0 to z=4LF . (Movie files: 2.4 MB, 2.3 MB, 3.3 MB, and 3.3 MB)

Download Full Size | PDF

Finally, in Fig. 3(g) we show the propagation of a travelling vector Parabolic-Gauss (VPG) beam with TM polarization. The electric field is given directly by Eq. (7a) with a seed function given by the superposition of even and odd Parabolic nondiffracting beams [11, 23], namely W(r1)=Pe(u2κ;p)Pe(v2κ;p)+iPO(u2κ;p)PO(v2κ;p) with parabolicity parameter p=2. The Cartesian derivatives are expressed in the Parabolic coordinates x=(v 2-u 2)/2, and y=uv as follows

x=1u2+v2(uuvv),
y=1u2+v2(vuuv),

The beam is again balanced, but now we start the propagation assuming a purely imaginary κ 1=i30 mm-1 corresponding to a VPG of the modified kind. As expected for this initial condition, the beam now will belong to the ordinary kind at the plane z2=LF .

To finish this section, let us remark that linearly polarized gVHzG beams can be also constructed as discussed in subsection 2.6. Since the propagation of these vector beams reduces to the propagation of scalar HzG beams, a variety of theoretical and experimental evolutions of linearly polarized gVHzG beams can be seen elsewhere, see for instance Refs. [11, 12, 26].

4. Conclusions

In this paper we have introduced a generalized form of the vector HzG beams that can be propagated in a closed and elegant form through axisymmetric paraxial optical systems characterized by ABCD transfer matrices. Once the choice of polarization of the transverse component is made, the propagation of the vector beams is completely characterized by the transformation of two independent complex parameters q and κ. The derivation of the new formulation has included the possibility of propagation in complex lenslike media having at most quadratic transverse variations of the index of refraction and the gain or loss.

Apart from a complex amplitude factor, the output field has the same mathematical structure as the input field, thus the gVHzG beams constitute a class of vector fields whose form is invariant under paraxial optical transformations. As a consequence, the transverse polarization of the gVHzG beams does not change under paraxial ABCD lossless transformations. This form-invariance property should not to be confused with the shape-invariance property of the scalar Hermite-Gauss or Laguerre-Gauss beams which preserve, excepting a scaling factor, the same transverse shape under paraxial lossless transformations. The intensity shape of the gVHzG beams will change because κ 1 and κ 2 are not proportional to each other through a real factor leading to different profiles of the function W, and moreover because the parameters q 1 and κ 1 are transformed according to different laws. Gaussian apodized fields with arbitrary polarization can be built up with a suitable superposition of constituent gVHzG beams with the same Gaussian envelope and transverse spatial frequency.

A. Appendix: Derivation of the output field e2(r2) [Eq. (7a)]

By inserting e 1(r 1) [Eq. (2)] into the Huygens diffraction integral Eq. (6) we obtain

e2(r2)=Kexp(iKL0)i2πBd2r1exp(iKr122q1)exp[iK2B(Ar122r1·r2+Dr22)]
×1[ππdϕg(ϕ)exp[iκ1(x1cosϕ+y1sinϕ)]],

where W(r 1;κ 1) has been already expressed using the expansion given by Eq. (3).

The gradient operator ∇1 can be introduced within the angular integral and its application gives raise to two Cartesian field components, namely e 2(r 2)=ex +ey ŷ. Working with the x component we obtain the expression

ex=Kexp(iKL0)i2πBd2r1exp(iKr122q1)exp[iK2B(Ar122r1·r2+Dr22)]
×iκ1ππdϕg(ϕ)cosϕexp[iκ1(x1cosϕ+y1sinϕ)],

that can be rearranged as

ex=Kκ1exp(iKL0)2πBexp(iKDr222B)ππdϕg(ϕ)cosϕ
×d2r1exp[iKr122(1q1+AB)]exp[iKBr1·r2+iκ1(x1cosϕ+y1sinϕ)],

The double integral can be evaluated by splitting it into two single integrals and the result is

πa2exp{14a2[(κ1cosϕKBx2)2+(κ1sinϕKBy2)2]},

where a 2=-i (K/2)(q11+A/B).

Replacing Eq.(30) into Eq. (29) and after expanding the quadratic terms we obtain

ex=exp(iKL0)exp(κ124a2)exp(iKDr222B)exp(K2r224a2B2)
×ππ[Kκ12a2Bcosϕdϕ]g(ϕ)exp[Kκ12a2B(x2cosϕ+y2sinϕ)].

Now, from the expansion in Eq. (3), it is easy to identify that the integral corresponds indeed to the derivative of the function W(r 2;κ 2), then we have

ex=κ1κ2exp(iκ1κ2B2K)G(r2,q2)x2W(r2;κ2)

where the parameters q 2 and κ 2 are given by Eq. (9), G(r 2,q 2) is the fundamental Gaussian beam [Eq. (8)], and the unimodular condition AD-BC=1 has been applied to eliminate D. The expression for ey is also given by Eq. (32) with the difference that the partial derivative is now taken with respect to the y coordinate. Superposing both Cartesian components and using the gradient operator definition yields

e2(r2)=κ1κ2exp(iκ1κ2B2K)G(r2,q2)2W(r2;k2)

which is indeed Eq.(7a).

Acknowledgments

This research was supported by Consejo Nacional de Ciencia y Tecnología (grant 42808) and by Tecnológico de Monterrey (grant CAT007). M. A. Bandres acknowledge support from Secretar ía de Educación Pública of México.

References and links

1. M. Lax, W. H. Louisell, and W. B. McKnight, “From Maxwell to paraxial wave optics,” Phys, Rev. A 11, 1365–1370 (1975). [CrossRef]  

2. L. W. Davis and G. Patsakos, “TM and TE electromagnetic beams in free space,” Opt. Lett. 6, 22–23 (1981). [CrossRef]   [PubMed]  

3. Z. Bouchal and M. Olivík, “Non-diffractive vector Bessel beams,” J. Mod. Opt. 42, 1555–1566 (1995). [CrossRef]  

4. D. G. Hall, “Vector-beam solutions of Maxwell’s wave equation,” Opt. Lett. 21, 9–11 (1996). [CrossRef]   [PubMed]  

5. A. A. Tovar and G. H. Clark, “Concentric-circle-grating, surface-emitting laser beam propagation in complex optical systems,” J. Opt. Soc. Am. A 14, 3333–3340 (1997). [CrossRef]  

6. A. Flores-Pérez, J. Hernández-Hernández, R. Jáuregui, and K. Volke-Sepúlveda, “Experimental generation and analysis of first-order TE and TM Bessel modes in free space,” Opt. Lett. 31, 1732–1734 (2006). [CrossRef]   [PubMed]  

7. K. Volke-Sepulveda and E. Ley-Koo, “General construction and connections of vector propagation invariant optical fields: TE and TM modes and polarization states,” J. Opt. A: Pure Appl. Opt. , 8, 867–877 (2006). [CrossRef]  

8. M. A. Bandres and J. C. Gutiérrez-Vega, “Vector Helmholtz-Gauss and vector Laplace-Gauss beams,” Opt. Lett. 30, 2155–2157 (2005). [CrossRef]   [PubMed]  

9. L. J. Chu, “Electromagnetic waves in elliptic hollow pipes of metal,” J. Appl. Phys. 9, 583–591 (1938). [CrossRef]  

10. R. D. Spence and C. P. Wells, “The propagation of electromagnetic waves in parabolic pipes,” Phys. Rev. 62, 58–62 (1942). [CrossRef]  

11. J. C. Gutiérrez-Vega and M. A. Bandres, “Helmholtz-Gauss waves,” J. Opt. Soc. Am. A 22, 289–298 (2005). [CrossRef]  

12. C. López-Mariscal, M. A. Bandres, and J. C. Gutiérrez-Vega, “Observation of the experimental propagation properties of Helmholtz-Gauss beams,” Opt. Eng. 45, 068001 (2006). [CrossRef]  

13. Q. Zhan, “Trapping metallic Rayleigh particles with radial polarization,” Opt. Express 12, 3377–3382 (2004). [CrossRef]   [PubMed]  

14. V. G. Niziev and A. V. Nesterov, “Influence of beam polarization on laser cutting efficiency,” J. Phys. D: Appl. Phys. 32, 1455–1461 (1999). [CrossRef]  

15. R. Dorn, S. Quabis, and G. Leuchs, “Sharper focus for a radially polarized light beam,” Phys. Rev. Lett. 91, 233901 (2003). [CrossRef]   [PubMed]  

16. Z. Bouchal, “Nondiffracting optical beams: physical properties, experiments, and applications” Czech. J. Phys. 53, 537–578 (2003). [CrossRef]  

17. L. W. Casperson, D. G. Hall, and A. A. Tovar, “Sinusoidal-Gaussian beams in complex optical systems,” J. Opt. Soc. Am. A 14, 3341–3348 (1997). [CrossRef]  

18. S. Ruschin, “Modified Bessel nondiffracting beams,” J. Opt. Soc. Am. A 11, 3224–3228 (1994). [CrossRef]  

19. M. Santarsiero, “Propagation of generalized Bessel-Gauss beams through ABCD optical systems,” Opt. Commun. 132, 1–7 (1996). [CrossRef]  

20. S. A. Collins, “Lens-system diffraction integral written in terms of matrix optics,” J. Opt. Soc. Am. 60, 1168–1177 (1970). [CrossRef]  

21. A. E. Siegman, Lasers (University Science, 1986).

22. J. C. Gutiérrez-Vega, M.D. Iturbe-Castillo, and S. Chávez-Cerda, “Alternative formulation for invariant optical fields: Mathieu beams,” Opt. Lett. 25, 1493–1495 (2000). [CrossRef]  

23. M. A. Bandres, J. C. Gutiérrez-Vega, and S. Chávez-Cerda, “Parabolic nondiffracting optical wave fields,” Opt. Lett. 29, 44–46 (2004). [CrossRef]   [PubMed]  

24. J. A. Stratton, Electromagnetic theory (McGraw-Hill, New York, 1941)

25. P. Morse and H. Feshbach, Methods of Theoretical Physics (McGraw-Hill, New York, 1953).

26. M. Guizar-Sicairos and J. C. Gutiérrez-Vega, “Generalized Helmholtz-Gauss beams and its transformation by paraxial optical systems,” Opt. Lett. 31, 2912–2914 (2006). [CrossRef]   [PubMed]  

Supplementary Material (8)

Media 1: AVI (2405 KB)     
Media 2: AVI (2359 KB)     
Media 3: AVI (3342 KB)     
Media 4: AVI (3308 KB)     
Media 5: AVI (3186 KB)     
Media 6: AVI (2596 KB)     
Media 7: AVI (3643 KB)     
Media 8: AVI (2287 KB)     

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (3)

Fig. 1.
Fig. 1. Physical picture of the decomposition of a gVHzG beam propagating in free space in terms of fundamental vector Gaussian beams whose mean propagation axes lie on the surface of a double cone.
Fig. 3.
Fig. 3. Propagation of the transverse intensity distribution and electric vector field for generalized vector Bessel-cosine-Gauss, Mathieu-Gauss, and parabolic-Gauss beams. The parameter data for the propagations are included within the text. The movies show the evolution from z=0 to z=4LF . (Movie files: 3.1 MB, 2.5 MB, 3.6 MB, and 2.2 MB)
Fig. 2.
Fig. 2. Propagation of the transverse intensity distribution and the electric vector field for circularly polarized gVHzG beams constructed with finite superposition of vector Gaussian beams. The parameter data for the propagations are included within the text. The movies show the evolution from z=0 to z=4LF . (Movie files: 2.4 MB, 2.3 MB, 3.3 MB, and 3.3 MB)

Equations (42)

Equations on this page are rendered with MathJax. Learn more.

R = ( r , z ) , r = ( x , y ) = ( r , θ ) ; K = ( k , k z ) , k = ( k x , k y ) = ( k , ϕ ) ,
e 1 ( r 1 ) = exp ( i K r 1 2 2 q 1 ) 1 W ( r 1 ; κ 1 ) , h 1 ( r 1 ) = ( ε κ ) 1 2 z ̂ × e 1 ( r 1 ) ,
W ( r 1 ; κ 1 ) = π π g ( ϕ ) exp [ i κ 1 ( x 1 cos ϕ + y 1 sin ϕ ) ] d ϕ ,
e z , 1 = i K 1 · e 1 = exp ( i K r 1 2 2 q 1 ) ( i κ 1 2 K W + 1 W · r 1 q 1 ) ,
h z , 1 = i K 1 · h 1 = ε μ exp ( i K r 1 2 2 q 1 ) ( z ̂ × 1 W ) · r 1 q 1 .
e 1 ( T E ) ( r 1 ) = exp ( i K r 1 2 2 q 1 ) [ z ̂ × 1 W ] , h 1 ( T E ) ( r 1 ) = ε μ exp ( i K r 1 2 2 q 1 ) 1 W .
e 2 ( r 2 ) = K exp ( i K L 0 ) i 2 π B e 1 ( r 1 ) exp [ i K 2 B ( A r 1 2 2 r 1 · r 2 + D r 2 2 ) ] d 2 r 1 ,
e 2 ( r 2 ) = κ 1 κ 2 exp ( i κ 1 κ 2 B 2 K ) G ( r 2 , q 2 ) 2 W ( r 2 ; κ 2 ) ,
h 2 ( r 2 ) = ( ε μ ) 1 2 z ̂ × e 2 ( r 2 ) ,
G ( r 2 , q 2 ) = exp ( i K L 0 ) A + B q 1 exp ( i K r 2 2 2 q 2 ) ,
q 2 = A q 1 + B C q 1 + D , κ 2 = κ 1 q 1 A q 1 + B .
S j = 1 2 ( ε μ ) 1 2 f j 2 { j W 2 z ̂ + Re [ ( i κ j 2 K W + j W · r j q j ) j W * ] } ,
e ˜ ( k ) = 1 2 π e ( r ) exp ( i k · r ) d 2 r .
e ˜ 1 ( k 1 ) = i exp ( i q 1 κ 1 2 2 K ) exp ( i q 1 k 1 2 2 K ) ˜ 1 W ( k 1 ; q 1 κ 1 K ) ,
e ˜ 2 ( k 2 ) = κ 1 K κ 2 q 2 exp ( i q 2 κ 2 2 + κ 1 κ 2 B 2 K ) G ˜ ( k 2 , q 2 ) ˜ 2 W ( k 2 ; B κ 2 2 K κ 1 ) ,
G ˜ ( k 2 , q 2 ) = ( i q 2 K ) exp ( i K L 0 ) A + B q 1 exp ( i q 2 k 2 2 2 K )
E = α E ( T M ) + β E ( T E ) ,
W = 1 2 ( W x i W y ) u ̂ + + 1 2 ( W x + i W y ) u ̂ .
e j ± ( r j ) = 2 f j ( W x j i W y j ) u ̂ ± ,
q 2 = q 1 + L , κ 2 = κ 1 q 1 q 1 + L ,
e 2 ( r 2 ) = π π g ( ϕ ) g 2 ( r 2 ; ϕ ) d ϕ ,
g 2 ( r 2 ; ϕ ) = i exp ( i κ 1 2 2 K L ζ ) exp ( i K L ) ζ exp ( i K r 2 2 2 q 1 ζ ) exp ( i κ 1 · r 2 ζ ) κ 1 ,
r gen ( z ) = q 1 2 κ 1 𝒥 K q 1 𝒥 + κ 1 𝓡 q 1 𝒥 + κ 1 𝒥 q 1 𝓡 K q 1 𝒥 ( z z 1 ) .
z vertex = z 1 q 1 2 κ 1 𝒥 κ 1 𝓡 q 1 𝒥 + κ 1 𝒥 q 1 𝓡 .
[ A B C D ] = [ cos ( L a ) a sin ( L a ) sin ( L a ) a cos ( L a ) ] .
q 2 = a q 1 cos ( L a ) + a sin ( L a ) q 1 sin ( L a ) + a cos ( L a ) , κ 2 = κ 1 q 1 q 1 cos ( L a ) + a sin ( L a ) ,
W ( r 1 ) = n = 1 N A n exp [ i κ 1 r 1 cos ( θ 1 ϕ n ) ] ,
x = 1 f ( cosh 2 ξ cos 2 η ) ( sinh ξ cos η ξ cosh ξ sin η η ) ,
y = 1 f ( cosh 2 ξ cos 2 η ) ( cosh ξ sin η ξ sinh ξ cos η η ) ,
x = 1 u 2 + v 2 ( u u v v ) ,
y = 1 u 2 + v 2 ( v u u v ) ,
e 2 ( r 2 ) = K exp ( i K L 0 ) i 2 π B d 2 r 1 exp ( i K r 1 2 2 q 1 ) exp [ i K 2 B ( A r 1 2 2 r 1 · r 2 + D r 2 2 ) ]
× 1 [ π π d ϕ g ( ϕ ) exp [ i κ 1 ( x 1 cos ϕ + y 1 sin ϕ ) ] ] ,
e x = K exp ( i K L 0 ) i 2 π B d 2 r 1 exp ( i K r 1 2 2 q 1 ) exp [ i K 2 B ( A r 1 2 2 r 1 · r 2 + D r 2 2 ) ]
× i κ 1 π π d ϕ g ( ϕ ) cos ϕ exp [ i κ 1 ( x 1 cos ϕ + y 1 sin ϕ ) ] ,
e x = K κ 1 exp ( i K L 0 ) 2 π B exp ( i K D r 2 2 2 B ) π π d ϕ g ( ϕ ) cos ϕ
× d 2 r 1 exp [ i K r 1 2 2 ( 1 q 1 + A B ) ] exp [ i K B r 1 · r 2 + i κ 1 ( x 1 cos ϕ + y 1 sin ϕ ) ] ,
π a 2 exp { 1 4 a 2 [ ( κ 1 cos ϕ K B x 2 ) 2 + ( κ 1 sin ϕ K B y 2 ) 2 ] } ,
e x = exp ( i K L 0 ) exp ( κ 1 2 4 a 2 ) exp ( i K D r 2 2 2 B ) exp ( K 2 r 2 2 4 a 2 B 2 )
× π π [ K κ 1 2 a 2 B cos ϕ d ϕ ] g ( ϕ ) exp [ K κ 1 2 a 2 B ( x 2 cos ϕ + y 2 sin ϕ ) ] .
e x = κ 1 κ 2 exp ( i κ 1 κ 2 B 2 K ) G ( r 2 , q 2 ) x 2 W ( r 2 ; κ 2 )
e 2 ( r 2 ) = κ 1 κ 2 exp ( i κ 1 κ 2 B 2 K ) G ( r 2 , q 2 ) 2 W ( r 2 ; k 2 )
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.