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

Mode symmetry in microstructured fibres revisited

Open Access Open Access

Abstract

Motivated by the recent development of hollow-core, microstructured optical fibres without a translationally-periodic cladding structure, an analysis of the symmetry-induced properties of the modes of a uniform fibre with CMv point group symmetry is presented. Modes are classified according to the irreducible representations of the longitudinal components of the electric and magnetic fields; the radial component of the electric field and azimuthal component of the magnetic field share the symmetry of the longitudinal component of the electric field, while the radial component of the magnetic field and the azimuthal component of the electric field have the same symmetry as the longitudinal magnetic field. The symmetry class of the lowest order modes in fibres with point group CMv with M from 1 to 10 is given. The work is closely related to previous analyses, but clarifies the symmetry relationship between the two orthogonal modes in a degenerate pair and how the different symmetries of the longitudinal electric and magnetic field components emerge from the governing equations.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

McIsaac, in two seminal papers in 1975, essentially solved the problem of the symmetry-induced characteristics of modes in a uniform waveguide [1, 2]. He showed that modes can be classified according to the irreducible representations of the symmetry group of the waveguide and that all modes are either non-degenerate or doubly degenerate. For each class of mode he gave expressions, based on a Fourier expansion in the azimuthal variable, for the form of the longitudinal electric and magnetic fields in a waveguide whose transverse structure has a point group symmetry of either CM or CMv, for arbitrary M. His analysis also provided the minimum sector of the waveguide required to determine the complete electromagnetic field, and the boundary conditions satisfied by the longitudinal components on the edge of the minimum sector. This work came to prominence in the context of the development of microstructured optical fibres and, in particular, photonic crystal fibres with hexagonal symmetry (i.e. point group C6v). Examples of the use and development of McIsaac’s symmetry analysis include [3–14].

The motivation for revisiting this topic comes from the development of a new generation of anti-resonant microstructured fibres (often termed negative curvature fibres) that dispense with a translationally-periodic cladding structure but which demonstrate highly efficient guidance of light in a low refractive index core, e.g. air ([15, 16], and references therein). These fibres are not restricted to hexagonal symmetry, and examples have been published where the point group of the idealised transverse structure is C7v [17–19], C8v [20–23], or C10v [22]. The aim of this work is to provide a complete analysis of the symmetries of the field components for all mode classes for fibres with a CMv point group. In doing this two aspects of modal symmetry not fully covered by McIsaac’s and subsequent work are clarified. One concerns how the connection between the symmetries of the electric and magnetic field components emerges from the underlying governing equations. The second is the precise relationship between the field components (transverse as well as longitudinal) in doubly degenerate modes. Although these points have been touched upon in a number of previous papers (details are given in the next section), it is anticipated that bringing this analysis together in a consistent and coherent form will be of use to the community.

2. Symmetry analysis

The analysis is based on a fibre that is uniform along its length (the z direction) and with a transverse structure defined by an isotropic dielectric function ϵ(r, θ). Cylindrical polar coordinates (r, θ, z) are used throughout because they fit naturally with the symmetries of the fibre. It is assumed that the transverse dielectric profile has a point group CMv. The symmetry operations that leave the structure invariant are rotations about the fibre axis by 2/M, where p = 0, 1, 2,, M − 1, and mirror lines in the plane of the fibre that are separated by angles p′π/M, where p′ = 0, 1,, M − 1. For even values of M there are two distinct types of mirror lines, while for odd M all the mirrors are equivalent. For simplicity it is assumed that the x-axis, i.e. θ = 0, is a mirror line. In the case of a hollow-core, anti-resonant fibre whose cladding consists of a ring of thin-walled capillaries [15,16], M corresponds to the number of capillaries; for even values of M the distinct types of mirror lines are those passing through or between the capillaries respectively (see Fig. 1).

 figure: Fig. 1

Fig. 1 Schematic diagram of the cross section of hollow-core, anti-resonant fibres consisting of thin-walled capillaries uniformly distributed around an outer jacket. Air regions are white and glass regions are shaded. (a) and (b) have point groups C7v and C8v respectively. Mirror lines are shown by dashed and dotted lines; in (a) all mirror lines are equivalent, while in (b) there are two distinct sets of mirror lines represented by the two line types.

Download Full Size | PDF

The analysis is divided into three parts. In the following subsection, the symmetries of a scalar wave equation are analysed and expressions for the generic form of solutions with different symmetry types (i.e. the different irreducible representations) are presented. These expressions are equivalent to those given by [1], except for a clarification of the form of the doubly degenerate solutions. The extension to vector modes is then discussed, first by demonstrating the symmetry connection between the ez and hz components and then including the full set of in-plane components. It is shown that each component of the electric and magnetic fields in a mode has the symmetry characteristics of a scalar wave, but different components have different irreducible representations. In the third subsection the relationship between the modes of a fibre with CMv symmetry and those of a cylindrically symmetric fibre is discussed.

2.1. Symmetry of solutions of a scalar wave equation

The eigenstates, ψ, of a two-dimensional scalar wave equation with an underlying CMv symmetry can be characterised by the way in which they are transformed under the elements of the point group. If an initial ψi(r, θ) is rotated anti-clockwise by an angle ϕ relative to fixed coordinates then it transforms to ψf (r, θ) = ψi(r, θϕ). Similarly, for reflection in a mirror line at an angle ϕ, ψf (r, θ) = ψi(r, θ). Each eigenstate transforms according to one of the irreducible representations of the point group, and character tables of the irreducible representations are given in many textbooks. A useful web resource that includes data for CMv groups up to M = 32 is [24]. The approach taken by McIsaac [1, 2] is followed here, in which explicit forms for the irreducible representations are constructed from basis functions of the form fn(r) cos and gn(r) sin , where f and g are arbitrary functions of r and only certain integers n are allowed in each irreducible representation. As discussed by Fini [9] (see also Ferrando et al. [7]) an exponential form can be used instead of cosine/sine functions; this can be advantageous for degenerate modes in terms of the minimum sector required for a full numerical solution. In terms of symmetry, the approaches are essentially equivalent. Most of the papers referred to in the introduction use McIsaac’s notation to label the different mode classes. A more standard Mulliken notation for irreducible representations is used here; this notation was also used, for example, by Aghaie et al. [12] and Wolff et al. [14].

All CMv point groups have non-degenerate A1 and A2 representations for which n is an integer multiple of M and which can be expressed in the form

A1:j=0fjM(r)cos(jMθ)andA2:j=1gjM(r)sin(jMθ)
respectively, where j is an integer. Under rotation by 2/M both of these generic forms transform into themselves (i.e. ψf = ψi) while under reflection in a mirror line at p′π/M, states of A1 symmetry transform to themselves and states of A2 symmetry have ψf = −ψi. If M is even there is a second set of non-degenerate representations, labelled by B1 and B2, which are expanded as
B1:j=0fjM+M2(r)cos((jM+M/2)θ)andB2:j=0gjM+M2(r)sin((jM+M/2)θ).

Under rotation by 2/M both of these transform as ψf = cos() ψi and under reflection in a mirror line at p′ π/M, B1 states transform as ψf = cos(p′π) ψi and B2 states transform as ψf = − cos(p′π)ψi.

For all CMv points groups with M > 2 there are doubly degenerate representations labelled by Eq, with q = 1, 2,. The number of Eq representations increases with M; for even and odd values of M there are M/2 − 1 and (M − 1)/2 distinct representations respectively. Within each doubly degenerate pair there are two states that can be constructed from cosine and sine basis functions and which are denoted by Eqc and Eqs. These are expanded as

Eqc:j=f|jM+q|(r)cos((jM+q)θ)andEqs:j=f|jM+q|(r)sin((jM+q)θ)
where states labelled by c and s have a mirror and anti-mirror respectively along the x axis; this choice of functions ensures that c and s states are orthogonal. A general function that belongs to the Eq representation can be expressed as ψ = αc Eqc + αs Eqs where αc and αs are arbitrary constants. Under the symmetry operations of the point group this general function transforms into a different linear combination of Eqc and Eqs. For a rotation by 2/M this can be expressed as
(αcαs)f=(cos2pqπMsin2pqπMsin2pqπMcos2pqπM)(αcαs)i
where the i and f labels indicate the coefficients before and after the rotation. The equivalent expression for reflection in a mirror line at p′π/M is
(αcαs)f=(cos2pqπMsin2pqπMsin2pqπMcos2pqπM)(αcαs)i.

The matrices in Eqs. (4) and (5) are consistent with the characters of the symmetry elements for Eq irreducible representations [24], as are the transformations given above for states with A-and B-type symmetry.

To make the analysis more concrete, explicit forms for states of different symmetries for point groups C7v and C8v are given below; any eigenstate ψ will be expressible in one of these forms:

C7vA1:f0(r)+f7(r)cos7θ+f14(r)cos14θ+A2:g7(r)sin7θ+g14(r)sin14θ+E1c:f1(r)cosθ+f6(r)cos6θ+g8(r)cos8θ+f13(r)cos13θ+E1s:f1(r)sinθf6(r)sin6θ+f8(r)sin8θf13(r)sin13θ+E2c:f2(r)cos2θ+f5(r)cos5θ+f9(r)cos9θ+f12(r)cos12θ+E2s:f2(r)sin2θf5(r)sin5θ+f9(r)sin9θf12(r)sin12θ+E3c:f3(r)cos3θ+f4(r)cos4θ+f10(r)cos10θ+f11(r)cos11θ+E3s:f3(r)sin3θf4(r)sin4θ+f10(r)sin10θf11(r)sin11θ+
C8vA1:f0(r)+f8(r)cos8θ+f16(r)cos16θ+A2:g8(r)sin8θ+g16(r)sin16θ+B1:f4(r)cos4θ+f12(r)cos12θ+B2:g4(r)sin4θ+g12(r)sin12θ+E1c:f1(r)cosθ+f7(r)cos7θ+f9(r)cos9θ+f15(r)cos15θ+E1s:f1(r)sinθf7(r)sin7θ+f9(r)sin9θf15(r)sin15θ+E2c:f2(r)cos2θ+f6(r)cos6θ+f10(r)cos10θ+f14(r)cos14θ+E2s:f2(r)sin2θf6(r)sin6θ+f10(r)sin10θf14(r)sin14θ+E3c:f3(r)cos3θ+f5(r)cos5θ+f11(r)cos11θ+f13(r)cos13θ+E3s:f3(r)sin3θf5(r)sin5θ+f11(r)sin11θf13(r)sin13θ+

The patterns seen here can easily be extended to other values of M. In the expressions for the functions of Eq symmetry, the alternating pattern of signs in the Eqs expansion arises from the negative values of j in Eq. (3).

The expansions given in Eq. (3) for doubly degenerate modes are a little different from those of McIsaac [1] and other authors. McIsaac uses two semi-infinite sums rather than the j sums in Eq. (3) running from −∞ to ∞. More importantly, he does not highlight the fact that the radial terms in the Eqc and Eqs expansions must be related to one another, as seen in Eqs. (3), (7) and (8). This connection is implicit in previous work (e.g. [3, 7, 9]) but does not appear to have been brought out explicitly.

As well as the basic transformation properties of the states, two extensions are required for the analysis in subsection 2.2. First, the symmetries of the derivatives of states are considered. Differentiation with respect to r results in a function with the same irreducible representation as the original state, but differentiation with respect to θ changes the representation, as shown in the first line of Table 1. It is straightforward to show from Eqs. (1) and (2) that the non-degenerate representations, A1/A2 and B1/B2 transform into a function with the symmetry of the other representation in the pair, i.e. 1 ↔ 2. The doubly degenerate Eq representations Eq. (3) transform into a pair of functions that also have Eq symmetry. For an initial function ψ = αc Eqc + αs Eqs the differential with respect to θ can be written in the form ψ=αcEqc+αsEqs, where Eqc and Eqs are a pair of functions that have the standard form given by Eq. (3) and the coefficients αc,s are related to the original coefficients by

(αcαs)=(0110)(αcαs).

The transformation matrix here is included in the first line of Table 1.

Tables Icon

Table 1. Effect of differentiation with respect to θ and multiplication by functions with A1 and A2 symmetry on states with given irreducible representations. The top row gives the original representation and the other rows the representation resulting from the given operation.

The symmetry properties of the product of a function that has A1 or A2 symmetry with states of different irreducible representations are also required. These can be determined by direct multiplication of the standard forms given in Eq. (1) with each of Eqs. (1), (2) and (3), followed by identification of the symmetry type of the resulting expression. The results are shown in Table 1; multiplication by a function of A1 symmetry does not change the representation, while multiplication by a function of A2 symmetry changes the representation in an equivalent way to differentiation with respect to θ. The transformation matrices given for the Eq representation are to be interpreted in the same way as in Eq. (8).

2.2. Modal symmetry in fibres with point group CMv

To extend these properties of scalar solutions to the vector modes of a fibre, an analysis of the symmetries of the field components er(r, θ), eθ (r, θ), ez (r, θ), h˜r(r,θ), h˜θ(r,θ) and h˜z(r,θ) is required (a scaled magnetic field defined by h˜(r,θ)=(μ0/ϵ0)1/2h(r,θ) is used to simplify the notation, and lower case letters are used for the fields to avoid any confusion between the electric field and the E symmetry type). All field components are assumed to have a z dependence of the form exp(iβz), where β is the propagation constant. The components are closely inter-related and any two of them are sufficient to generate the full set. For example, if ez and h˜z are known, the remaining components are given by

er(r,θ)=ik2[βezr+k0rh˜zθ]
eθ(r,θ)=ik2[βrezθk0h˜zr]
h˜r(r,θ)=ik2[βh˜zrϵk0rezθ]
h˜θ(r,θ)=ik2[βrh˜zθ+ϵk0ezr].
where k0 is the free space wavevector and k the transverse wavevector
k(r,θ)=(ϵ(r,θ)k02β2)1/2.

These equations are well known in the analysis of circularly symmetric fibres where ϵϵ(r) [25], but they are also valid for a general dielectric profile ϵ(r, θ). Similarly, ez and h˜z can be expressed in terms of the transverse field components as

ez(r,θ)=ik0ϵ1r[r(rh˜θ)h˜rθ]
h˜z(r,θ)=ik01r[r(reθ)erθ].

Coupled wave equations that depend only on ez and h˜z are derived by substituting Eq. (9) into Eq. (11), yielding

(2+k2)ez1ϵϵr[β2k2ezr+βk0k2rh˜zθ]1ϵϵθ[β2k2r2ezθβk0k2rh˜zr]=0
(2+k2)h˜zϵr[k02k2h˜zrβk0k2rezθ]ϵθ[k02k2r2h˜zθ+βk0k2rezr]=0.
where ∇2 is the two-dimensional Laplacian operator. These equations are versions in cylindrical polar coordinates of the general equations given, for example, in [25]. A smoothly-varying dielectric function is implied by the differentials in Eqs. (12), but the modal symmetries that emerge from Eqs. (12) will be retained in the limiting case of a piecewise constant dielectric function (as is usually assumed to be the case for microstructured fibres). Eqs. (12) form the basis for the analysis; once the symmetries of ez and h˜z have been determined, Eqs. (9) can then be used to find the symmetries of the transverse components. Instead of starting from Eqs. (12), McIsaac [1, 2] assumes a piecewise dielectric function and makes the connection between the symmetries of ez and h˜z through the boundary conditions. He states in a footnote that his results also apply in waveguides with isotropic media where the dielectric function is transversely inhomogeneous; the analysis presented here shows explicitly how the different symmetries of the longitudinal components arises from the governing equations Eqs. (12) for this case.

We first note that all the functions, other than the field components, that appear in Eqs. (9) and (12) have either A1(ϵ,1/ϵ, ∂ϵ/∂r, k2, 1/k2) or A2(∂ϵ/∂θ) symmetry. These functions always appear as multipliers of the field terms, and the symmetry effect of such multiplications is given in Table 1. Where there are repeated products (for example, in the term 1/ϵ × ∂ϵ/∂θ × 1/k2) the symmetry is found by repeated application of Table 1 (thus the overall symmetry of the example product is A2).

The key point in determining the allowed symmetries of vector modes is that each term in Eq. (12a) (and separately in Eq. (12b)) must transform under the same irreducible representation, so that the whole equation transforms consistently under rotations and reflections. For example, if ez has A1 symmetry then, from Table 1, ∂ez /∂θ has A2 symmetry. Every term in Eq. (12a) must then have A1 symmetry (because the (∇2 + k2)ez, ∂ez /∂r, and ∂ez /∂θ terms do) and every term in Eq. (12b) must have A2 symmetry (because the ∂ez /∂r and ∂ez /∂θ terms do). It follows that, in this case, h˜z must have A2 symmetry to keep the symmetries of the two coupled equations consistent. The symmetries of the other field components can then be found from Eqs. (9). The result is given in the first line of Table 2, where this overall symmetry type is labelled as A(1). An equivalent analysis can be applied to the other non-degenerate solutions. By starting with the symmetry of ez being A2, B1 and B2 in turn, the symmetries of the other field components can be determined, with the resulting modal symmetries being labelled as A(2), B(1) and B(2) in Table 2. In each case, it can be seen that every field component has a symmetry that is characteristic of a solution of a scalar wave equation, but components with the given symmetries must be combined to create a complete vector mode.

Tables Icon

Table 2. Irreducible representations of the components of the electric and magnetic fields for modes of A(1), A(2), B(1), B(2) and Eq(1,2)) symmetry (see text for details).

The doubly degenerate modes are treated in a similar way. All the terms in both Eqs. (12a) and (12b) must be expressible in the form αc Eqc + αs Eqs where Eqc and Eqs are functions of the type given by Eq. (3) and the constants αc and αs are the same for each term within the separate equations. This will ensure that both equations transform correctly under the symmetry elements of the point group. If it is assumed that ez has this standard form with αc = α1 and αs = α2 then all the terms in Eq. (12a) must have these same constants multiplying the Eqc and Eqs parts of each term. The effects of the products and differentiations in each term can be determined by repeated application of the matrices given in Table 1. It is straightforward to show that Eqs. (12a) and (12b) have the correct overall form if and only if the constants in the expansion of h˜zhave the form αc = α2 and αs = −α1. It then follows from Eqs. (9), in the same way as for the non-degenerate modes, that er and h˜θ have the same form as ez, while eθ and h˜r have the same form as h˜z. This overall symmetry type is labelled as Eq(1,2) in Table 2, where the column vectors for each component are to be interpreted in the same way as in Eqs. (4) and (5). Orthogonal modes within the doubly degenerate pair for this overall symmetry type can be obtained by choosing particular values for the constants α. The final two lines of Table 2 give a pair of modes: Eq(1) with α1 = 1 and α2 = 0, and Eq(2) with α1 = 0 and α2 = 1. It should be emphasised that the signs given for the field components for Eq(1) and Eq(2) do not imply any particular sign relationship between different components; they give only the sign relationship between the two modes for a given component. For example, if ez for modes 1 and 2 within a degenerate pair has a Eqc and Eqs form respectively, then h˜z for modes 1 and 2 will have Eqs and Eqc forms respectively, but with opposite signs.

Previous authors (e.g. [11]) have given expressions for transverse field components based on McIsaac’s expansions, but Table 2, combined with Eqs. (1), (2) and (3), provides a complete description of the symmetry for every mode class in fibres with a CMv point group.

2.3. Relationship with modes of a fibre of cylindrical symmetry

Modes in fibres with cylindrical symmetry (which corresponds to a point group C∞v) are conventionally labelled as TE0m, TM0m, HEnm and EHnm, where n describes the azimuthal dependence of the field components and m labels the different modes for each n. All field components in a given mode vary as cos() or sin() [25] and thus there is a clear connection with the generic scalar functions given in Eqs. (1) to (3). The difference is that in C∞v the sums over M are absent and only one term is present in each case. A1 corresponds to n = 0 and fields with this symmetry are constant in θ. The A2, B1 and B2 functions do not exist in C∞v. Functions of the Eq type retain only the cos()/sin() term, and so q can be identified with the azimuthal label n in cylindrically symmetric fibres.

TE0m modes have ez=er=h˜θ=0 and can therefore be identified with the general A(2) mode in Table 2. Similarly, TM0m modes have h˜z=h˜r=eθ=0 and are of the general A(1) type. HEnm and EHnm modes consist of a degenerate pair with the ez components in the two members of the pair having the form f(r) cos and f(r) sin , and the h˜z components −g(r) sin and g(r) cos , where f and g are functions of the radial variable [25]. Thus both HE and EH modes have the generic Eq(1,2) form of Table 2; there is no underlying symmetry difference between HE and EH solutions.

The modes of a hollow core fibre with point group CMv are closely related to those of a hollow tube with C∞v symmetry, but the reduction in symmetry caused by the microstructure can lead to changes in the modal symmetries, including degeneracy breaking. This is well known for point group C6v (e.g. [6, 12]) and in Table 3 the analysis is extended to point groups from C1v to C10v. The symmetry properties of the 12 lowest order modes, grouped by their value of n, are considered and the Table shows which generic symmetry type given in Table 2 each group falls into. For these modes, microstructured fibres with M ≥ 9 have essentially the same symmetry characteristics as a hollow tube, but for M = 8 and lower, there are differences. For example, in a hexagonal fibre with C6v symmetry, the HE31 mode is split into two non-degenerate modes, and the HE41 mode is symmetrically equivalent to the group of modes with n = 2. In contrast, in a fibre with C7v symmetry the HE31 mode remains doubly degenerate, while HE31 and HE41 share the same symmetry type which is distinct from the modes with n = 2. Fibres with C8v symmetry are similar to those with a C7v point group except that the eight-fold rotation symmetry splits the HE41 mode into two non-degenerate modes. It is straightforward to extend the entries in Table 3 both to higher order modes and to larger values of M.

Tables Icon

Table 3. Symmetry types for the 12 lowest modes in hollow-core, microstructured fibres with point groups from C1v to C10v. The modes are grouped and labelled according to the equivalent modes in a cylindrically symmetric hollow tube. Where the degeneracy is broken, the symmetries of the two resulting modes are separated by a semi-colon.

3. Conclusions

The key result of this paper is the description of the symmetries of all field components for each class of mode that exists in fibres with a CMv point group, as given in Table 2 and Eqs. (1) to (3). As mentioned in the Introduction, it is not claimed that this work is entirely new, or that previous work is incorrect. However by grounding the symmetry analysis in the governing equations for the longitudinal field components, Eqs. (12a) and (12b), and by using a matrix-based approach for the degenerate modes, a more complete and systematic picture of the symmetry emerges. With the growing interest in negative curvature fibres with relatively unfamiliar point groups, it is anticipated that Table 3 will provide a useful reference for the allowed symmetries of the lower order modes.

References

1. P. R. McIsaac, “Symmetry-induced modal characteristics of uniform waveguides – I: Summary of results,” IEEE Trans. Microw. Theory Tech. 23(5), 421–429 (1975). [CrossRef]  

2. P. R. McIsaac, “Symmetry-induced modal characteristics of uniform waveguides – II: Theory,” IEEE Trans. Microw. Theory Tech. 23(5), 429–433 (1975). [CrossRef]  

3. M. J. Steel, T. P. White, C. Martijn de Sterke, R. C. McPhedran, and L. C. Botten, “Symmetry and degeneracy in microstructured optical fibres,” Opt. Lett. 26(8), 488–490 (2001). [CrossRef]  

4. T. P. White, B. T. Kuhlmey, R. C. McPhedran, D. Maystre, G. Renversez, C. Martijn de Sterke, and L. C. Botton, “Multipole method for microstructured optical fibers. I. Formulation,” J. Opt. Soc. Am. B 19(10), 2322–2330 (2002). [CrossRef]  

5. B. T. Kuhlmey, T. P. White, G. Renversez, D. Maystre, L. C. Botten, C. Martijn de Sterke, and R. C. McPhedran, “Multipole method for microstructured optical fibers. II. Implementation and results,” J. Opt. Soc. Am. B 19(10), 2331–2340 (2002). [CrossRef]  

6. R. Guobin, W. Zhi, L. Shuqin, and J. Shuisheng, “Mode classification and degeneracy in photonic crystal fibers,” Opt. Express 11(11), 1310–1321 (2003). [CrossRef]   [PubMed]  

7. A. Ferrando, M. Zacarés, P. Fernández de Córdoba, D. Binosi, and J. A. Monsoriu, “Vortex solitons in photonic crystal fibers,” Opt Express 12(5), 817–822 (2004). [CrossRef]   [PubMed]  

8. M. J. Steel, “Reflection symmetry and mode transversality in microstructured fibers,” Opt. Express 12(8), 1497–1509 (2004). [CrossRef]   [PubMed]  

9. J. M. Fini, “Improved symmetry analysis of many-moded microstructure optical fibers,” J. Opt. Soc. Am. B 21(8), 1431–1436 (2004). [CrossRef]  

10. P. Boyer, G. Renversez, E. Popov, and M. Nevière, “Improved differential method for microstructured optical fibres,” J. Opt. A: Pure Appl. Opt. 9, 728–740 (2007). [CrossRef]  

11. F. Poletti and P. Horak, “Description of ultrashort pulse propagation in multimode optical fibers,” J. Opt. Soc. Am. B 25(10), 1645–1654 (2008). [CrossRef]  

12. K. Z. Aghaie, V. Dangui, M. J. F. Digonnet, S. Fan, and G. S. Kino, “Classification of the core modes of hollow-core photonic-bandgap fibers,” IEEE J. Quantum Electron. 45(9), 1192–1200 (2009). [CrossRef]  

13. F. Zolla, G. Renversez, A. Nicolet, B. Kuhlmey, S. Guenneau, D. Felbacq, A. Argyros, and S. Leon-Saval, Foundations of Photonic Crystal Fibres, 2nd Edition (World Scientific, 2012). [CrossRef]  

14. C. Wolff, M. J. Steel, and C. G. Poulton, “Formal selection rules for Brillouin scattering in integrated waveguides and structured fibers,” Opt. Express 22(26), 32489–32501 (2014). [CrossRef]  

15. F. Yu and J. C. Knight, “Negative curvature hollow-core optical fiber,” IEEE J. Sel. Topics Quantum Electron. 22(2), 4400610 (2016). [CrossRef]  

16. C. Wei, R. J. Weiblen, C. R. Menyuk, and J. Hu, “Negative curvature fibers,” Adv. Opt. Photonics 9(3), 504–561 (2017). [CrossRef]  

17. M. Michieletto, J. K. Lyngsø, C. Jakobsen, J. Lægsgaard, O. Bang, and T. T. Alkeskjold, “Hollow-core fibers for high power pulse delivery,” Opt. Express 24(7), 7103–7119 (2016). [CrossRef]   [PubMed]  

18. F. Yu, M. Xu, and J. C. Knight, “Experimental study of low-loss single-mode performance in anti-resonant hollow-core fibers,” Opt. Express 24(12), 12969–12975 (2016). [CrossRef]   [PubMed]  

19. J. R. Hayes, S. R. Sandoghchi, T. D. Bradley, Z. Liu, R. Slavík, M. A. Gouveia, N. V. Wheeler, G. Jasion, Y. Chen, E. N. Fokoua, M. N. Petrovich, D. J. Richardson, and F. Poletti, “Antiresonant hollow core fibre with an octave spanning bandwidth for short haul data communications,” J. Lightwave Technol. 35(3), 437–442 (2017). [CrossRef]  

20. A. D. Pryamikov, A. S. Biriukov, A. F. Kosolapov, V. G. Plotnichenko, S. L. Semjonov, and E. M. Dianov, “Demonstration of a waveguide regime for a silica hollow-core microstructured optical fibre with a negative curvature of the core boundary in the spectral region > 3.5µ m,” Opt. Express 19(2), 1441–1448 (2011). [CrossRef]   [PubMed]  

21. F. Yu, W. J. Wadsworth, and J. C. Knight, “Low loss silica hollow core fibers for 3–4 µ m spectral region,” Opt. Express 20(10) 11153–11158 (2012). [CrossRef]   [PubMed]  

22. M. R. Abu Hassan, F. Yu, W. J. Wadsworth, and J. C. Knight, “Cavity-based mid-IR fiber gas laser pumped by a diode laser,” Optica 3(3), 218–221 (2016). [CrossRef]  

23. B. Debord, A. Amsanpally, M. Chafer, A. Baz, M. Maurel, J. M. Blondy, E. Hugonnot, F. Scol, L. Vincetti, F. Gérôme, and F. Benabid, “Ultralow transmission loss in inhibited-coupling guiding hollow core fibers,” Optica 4(2), 209–217 (2017). [CrossRef]  

24. G. Katzer, “Character tables for point groups used in Chemistry,” http://gernot-katzers-spice-page.com/character_tables.

25. A. W. Snyder and J. Love, Optical Waveguide Theory(Springer, 1983).

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 (1)

Fig. 1
Fig. 1 Schematic diagram of the cross section of hollow-core, anti-resonant fibres consisting of thin-walled capillaries uniformly distributed around an outer jacket. Air regions are white and glass regions are shaded. (a) and (b) have point groups C7v and C8v respectively. Mirror lines are shown by dashed and dotted lines; in (a) all mirror lines are equivalent, while in (b) there are two distinct sets of mirror lines represented by the two line types.

Tables (3)

Tables Icon

Table 1 Effect of differentiation with respect to θ and multiplication by functions with A1 and A2 symmetry on states with given irreducible representations. The top row gives the original representation and the other rows the representation resulting from the given operation.

Tables Icon

Table 2 Irreducible representations of the components of the electric and magnetic fields for modes of A(1), A(2), B(1), B(2) and E q ( 1 , 2 )) symmetry (see text for details).

Tables Icon

Table 3 Symmetry types for the 12 lowest modes in hollow-core, microstructured fibres with point groups from C1v to C10v. The modes are grouped and labelled according to the equivalent modes in a cylindrically symmetric hollow tube. Where the degeneracy is broken, the symmetries of the two resulting modes are separated by a semi-colon.

Equations (17)

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

A 1 : j = 0 f j M ( r ) cos ( j M θ ) and A 2 : j = 1 g j M ( r ) sin ( j M θ )
B 1 : j = 0 f j M + M 2 ( r ) cos ( ( j M + M / 2 ) θ ) and B 2 : j = 0 g j M + M 2 ( r ) sin ( ( j M + M / 2 ) θ ) .
E q c : j = f | j M + q | ( r ) cos ( ( j M + q ) θ ) and E q s : j = f | j M + q | ( r ) sin ( ( j M + q ) θ )
( α c α s ) f = ( cos 2 p q π M sin 2 p q π M sin 2 p q π M cos 2 p q π M ) ( α c α s ) i
( α c α s ) f = ( cos 2 p q π M sin 2 p q π M sin 2 p q π M cos 2 p q π M ) ( α c α s ) i .
C 7 v A 1 : f 0 ( r ) + f 7 ( r ) cos 7 θ + f 14 ( r ) cos 14 θ + A 2 : g 7 ( r ) sin 7 θ + g 14 ( r ) sin 14 θ + E 1 c : f 1 ( r ) cos θ + f 6 ( r ) cos 6 θ + g 8 ( r ) cos 8 θ + f 13 ( r ) cos 13 θ + E 1 s : f 1 ( r ) sin θ f 6 ( r ) sin 6 θ + f 8 ( r ) sin 8 θ f 13 ( r ) sin 13 θ + E 2 c : f 2 ( r ) cos 2 θ + f 5 ( r ) cos 5 θ + f 9 ( r ) cos 9 θ + f 12 ( r ) cos 12 θ + E 2 s : f 2 ( r ) sin 2 θ f 5 ( r ) sin 5 θ + f 9 ( r ) sin 9 θ f 12 ( r ) sin 12 θ + E 3 c : f 3 ( r ) cos 3 θ + f 4 ( r ) cos 4 θ + f 10 ( r ) cos 10 θ + f 11 ( r ) cos 11 θ + E 3 s : f 3 ( r ) sin 3 θ f 4 ( r ) sin 4 θ + f 10 ( r ) sin 10 θ f 11 ( r ) sin 11 θ +
C 8 v A 1 : f 0 ( r ) + f 8 ( r ) cos 8 θ + f 16 ( r ) cos 16 θ + A 2 : g 8 ( r ) sin 8 θ + g 16 ( r ) sin 16 θ + B 1 : f 4 ( r ) cos 4 θ + f 12 ( r ) cos 12 θ + B 2 : g 4 ( r ) sin 4 θ + g 12 ( r ) sin 12 θ + E 1 c : f 1 ( r ) cos θ + f 7 ( r ) cos 7 θ + f 9 ( r ) cos 9 θ + f 15 ( r ) cos 15 θ + E 1 s : f 1 ( r ) sin θ f 7 ( r ) sin 7 θ + f 9 ( r ) sin 9 θ f 15 ( r ) sin 15 θ + E 2 c : f 2 ( r ) cos 2 θ + f 6 ( r ) cos 6 θ + f 10 ( r ) cos 10 θ + f 14 ( r ) cos 14 θ + E 2 s : f 2 ( r ) sin 2 θ f 6 ( r ) sin 6 θ + f 10 ( r ) sin 10 θ f 14 ( r ) sin 14 θ + E 3 c : f 3 ( r ) cos 3 θ + f 5 ( r ) cos 5 θ + f 11 ( r ) cos 11 θ + f 13 ( r ) cos 13 θ + E 3 s : f 3 ( r ) sin 3 θ f 5 ( r ) sin 5 θ + f 11 ( r ) sin 11 θ f 13 ( r ) sin 13 θ +
( α c α s ) = ( 0 1 1 0 ) ( α c α s ) .
e r ( r , θ ) = i k 2 [ β e z r + k 0 r h ˜ z θ ]
e θ ( r , θ ) = i k 2 [ β r e z θ k 0 h ˜ z r ]
h ˜ r ( r , θ ) = i k 2 [ β h ˜ z r ϵ k 0 r e z θ ]
h ˜ θ ( r , θ ) = i k 2 [ β r h ˜ z θ + ϵ k 0 e z r ] .
k ( r , θ ) = ( ϵ ( r , θ ) k 0 2 β 2 ) 1 / 2 .
e z ( r , θ ) = i k 0 ϵ 1 r [ r ( r h ˜ θ ) h ˜ r θ ]
h ˜ z ( r , θ ) = i k 0 1 r [ r ( r e θ ) e r θ ] .
( 2 + k 2 ) e z 1 ϵ ϵ r [ β 2 k 2 e z r + β k 0 k 2 r h ˜ z θ ] 1 ϵ ϵ θ [ β 2 k 2 r 2 e z θ β k 0 k 2 r h ˜ z r ] = 0
( 2 + k 2 ) h ˜ z ϵ r [ k 0 2 k 2 h ˜ z r β k 0 k 2 r e z θ ] ϵ θ [ k 0 2 k 2 r 2 h ˜ z θ + β k 0 k 2 r e z r ] = 0.
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.