Diffraction decorating caustics in gravitational lensing...1 Diffraction decorating caustics in...

28
1 Diffraction decorating caustics in gravitational lensing M V Berry H H Wills Physics Laboratory, Tyndall Avenue, Bristol BS8 1TL, UK [email protected] Abstract The spacings of interference fringes in an observation plane, and the accompanying intensity amplifications, are calculated for the gravitational lensing of a distant source by an isolated star and near the fold and cusp singularities from a binary star lens. In the astronomically natural transitional approximations, the familiar wavelength scalings associated with short-wave asymptotics are accompanied by a variety of dependences on disparate astronomical lengths, such as the Schwartzschild radius, separation of binary stars, and distance to the lens. Also discussed briefly are the images seen in a telescope looking in the direction of the source; if the aperture is modelled by Gaussian apodization, the image is a complexification of the wave in the observation plane. Keywords: asymptotics, diffraction catastrophes, scaling, focusing, apodization, cusps Submitted to: J.Optics, January 2021 Orcid id: 0000-0001-7921-2468 “...it would not be surprising if the wave optical features of gravitational lensing emerge observationally during the next century.”[1]

Transcript of Diffraction decorating caustics in gravitational lensing...1 Diffraction decorating caustics in...

  • 1

    Diffraction decorating caustics in gravitational lensing

    M V Berry

    H H Wills Physics Laboratory, Tyndall Avenue, Bristol BS8 1TL, UK

    [email protected]

    Abstract

    The spacings of interference fringes in an observation plane, and the

    accompanying intensity amplifications, are calculated for the gravitational

    lensing of a distant source by an isolated star and near the fold and cusp

    singularities from a binary star lens. In the astronomically natural transitional

    approximations, the familiar wavelength scalings associated with short-wave

    asymptotics are accompanied by a variety of dependences on disparate

    astronomical lengths, such as the Schwartzschild radius, separation of binary

    stars, and distance to the lens. Also discussed briefly are the images seen in a

    telescope looking in the direction of the source; if the aperture is modelled by

    Gaussian apodization, the image is a complexification of the wave in the

    observation plane.

    Keywords: asymptotics, diffraction catastrophes, scaling, focusing,

    apodization, cusps

    Submitted to: J.Optics, January 2021

    Orcid id: 0000-0001-7921-2468

    “...it would not be surprising if the wave optical features of gravitational lensing emerge observationally during the next century.”[1]

  • 2

    1. Introduction

    With gravitational lensing based on the deflection of light rays now established

    as a valuable tool in astronomy [1-4], there is growing anticipation of effects of

    wave interference, promising to add interferometric precision to inferences

    about distant objects. It is well understood that lensing is dominated by events

    and places where focusing occurs, that is, by caustics, and that the decoration of

    these singularities by ‘diffraction catastrophes’ [5, 6] will play a major role in

    the envisaged wave effects [7]. On astronomically relevant scales, the

    associated diffraction integrals are highly oscillatory, raising challenges for their

    numerical evaluation; an important recent advance, especially for

    multidimensional integrals, has been the development and application of

    sophisticated contour deformation techniques [8, 9].

    My aim here is to calculate the fringe spacings and intensity

    amplifications associated with several kinds of focusing, in an observation plane

    (Earth-based for example). For gravitational lensing, geometrical optics almost

    suffices, so the relevant regime is extreme short-wave asymptotics: small

    wavelengths l, i.e. large wavenumbers k=2p/l. The scalings with k are well

    understood [10, 11]: for example near a fold caustic the spacing is O(l2/3) –

    much larger than the familiar interference fringe spacing O(l) away from

    caustics – and the intensity amplification is O(l–1/3). (This explains, for

    example, why the rainbow – which is a fold caustic – is bright, and why our

    naked eyes can resolve supernumerary rainbows decorating the main bow [12],

    even though the wavelengths in sunlight are much smaller than raindrops.) But

    in addition to the wavelength dependence, the fringe spacings and

    amplifications are strongly influenced by the enormously varied relevant

    astronomical lengths: Schwartzschild radii, binary star separations, lens-

    observer distances. It is these additional dependences that will be calculated.

  • 3

    Section 2 revisits the basic wave theory for the wave in the observation

    plane. I will also briefly discuss the images in a telescope looking in the

    direction of the lensing objects; this is the intensity of the Fourier transform of

    this wave, windowed by the telecope aperture, and if this is modelled by

    Gaussian apodization the images are complexifications of those in the

    observation plane. The main results of the paper are explicit calculations for

    three simple lens models: the focal spot from a single star (section 3), and

    across the fold and cusp caustics from a binary star (section 4). The concluding

    section 5 summarises the main results, in terms of physical variables rather than

    the scaled variables convenient for calculation.

    In short-wave asymptotics there are four different regimes, and it is worth

    pinpointing the one relevant to the anticipated wave phenomena in gravitational

    lensing. Regime 1 is the coarsest scale, in which the wave intensity is simply

    the sum of intensities of contributing rays: inverse Jacobians between source

    and observer. This is the regime currently employed in conventional

    gravitational lensing: geometrical optics, with singularities at the caustics. Next

    is regime 2, in which the intensity is the square of the sum of interfering wave

    amplitudes, each with a phase (optical path length) decorating the square root of

    the inverse Jacobian, This is the ‘geometrical wave’; it gives an accurate

    description of the wave everywhere except very near the caustics, where the

    light is brightest and wave effects strongest – the very features of interest here.

    We need to describe the waves very close to the geometrical caustics. This is

    regime 3, where the lensed wave is described by canonical special functions -

    diffraction catastrophes [13, 14] – each representing a different type of caustic.

    This ‘transitional asymptotics’ does not match smoothly onto the geometrical

    wave of regime 2. The two are unified in the ‘uniform asymptotics’ of regime 4,

    in which the variables in the diffraction catastrophes are stretched to represent

    the exact optical path lengths of the contributing rays so as to be valid far from

  • 4

    caustics as well as close to them [5, 6, 13, 15-17]. In gravitational lensing, it is

    only very close to caustics that wave effects are likely to be observable, so the

    additional sophistication of regime 4 is unnecessary: regime 3 suffices, and has

    the advantage of leading to simple explicit formulas.

    2. Reprise of basic wave theory

    For present purposes, the simplest formulation suffices: paraxial diffraction

    from a distant point source, in the thin screen approximation [7, 18], in which

    the lenses are projected along the propagation direction, leading to a phase

    distribution W(r) in the lens plane r=(x,y) (figure 1). Mass endows space with a

    refractive index, for which an adequate approximation is

    , (2.1)

    where f(r,z) is the Newtonian gravitational potential. The lens plane phase is

    then

    , (2.2)

    in which ‘regularised’ refers to ignoring irrelevant infinite constants.

    n r, z( ) = 1− 2φ r, z( )c2

    W r( ) = dz−∞

    ∫ n r, z( )−1( )⎛

    ⎝⎜⎞

    ⎠⎟ regularised

  • 5

    Figure 1. Lens and observation planes

    Elementary Fresnel-Kirchhoff diffraction gives the wave in the

    observation plane R=(X,Y), distant Z from the phase screen, normalised to unit

    strength for the wave arriving at the screen:

    . (2.3)

    The intensity across the observation plane is |y|2. If the source is at a finite

    distance Zs from the screen, the only effect on the intensity is that R and Z are

    scaled by 1+Z/Zs: the wave beyond the screen is stretched and further away.

    Like all wave patterns, the interference detail will depend on k; this

    colour dependence will be a characteristic signal of wave lensing, distinguishing

    it from geometric gravitational lensing, the latter being achromatic because the

    index (2.1) is wavelength-independent. Spectral distortion of the Airy pattern

    observation plane R

    r

    Z

    distant source

    wavefront -W(r)

    lens plane

    ψ R,Z( ) = −ik2πZ

    dr exp ik W r( )+ r − R( )2

    2Z⎛

    ⎝⎜⎞

    ⎠⎟⎛

    ⎝⎜

    ⎠⎟

    r plane

    ⌡⎮⎮⌠

    ⌡⎮⎮

  • 6

    near a fold caustic was discussed elsewhere [19]. Simply put: caustics arising

    from polychromatic sources will be reddened on the dark side.

    It is not my purpose here to give a detailed discussion of the wave

    features of images seen when looking in directions W=(qX,qY) from the point

    R,Z with a telescope of finite aperture. Elsewhere [19] I have elaborated some

    of the peculiarities associated with looking at images when there are caustics in

    the telescope aperture, including the connection with Husimi functions. But it is

    worth drawing attention to one aspect.

    The wave in direction W is the windowed Fourier transform of the wave

    at R,Z in the observation plane. If the telescope aperture with radius L is

    modelled by Gaussian apodization rather than the more realistic sharp edges,

    this is

    . (2.4)

    A straightforward calculation using (2.3) leads to

    . (2.5)

    Thus the wave corresponding to looking in direction W is a complexification of

    the wave in the observation plane. The corresponding complexifications of the

    Airy and Pearcey functions were discussed in [19].

    An easy connection with the aperture-blurred geometrical image is

    obtained by approximating (2.5) by including only the terms that do not vanish

    for large k:

    ψ Ω,R,Z( ) = 12π L2

    d ′RR plane∫∫ exp −

    ′R − R( )22L2

    ⎝⎜⎞

    ⎠⎟exp ikΩ⋅ ′R( )ψ ′R ,Z( )

    ψ Ω,R,Z( ) = exp − 12 k 2L2Ω2( )exp ikΩ⋅R( ) ψ R,Z( )⎡⎣ ⎤⎦R→R+ikL2Ω,Z→Z−ikL2

  • 7

    (2.6)

    The saddle points r of the fast-varying first exponential factor in the integral are

    the lens plane points from which rays emerge in the viewing direction W, and

    with these r the second exponential factor gives the geometric aperture blurring

    (the diffraction blurring O(1/(kL)) requires considering both factors).

    3. Focal diffraction from a single star

    For a single star with mass M, the phase W in (2.3) depends on its

    Schwartzschild radius RS:

    . (3.1)

    This case has been studied extensively, and the diffraction integral evaluated

    exactly [7, 20], but we need to extract the relevant asymptotics. It is convenient

    to scale observation plane distances R, Z and wavenumbers k with RS, and lens

    plane distances r differently:

    . (3.2)

    In astronomical lensing, the parameters k and z greatly exceed unity. With these

    scalings, (2.3) becomes

    , (3.3)

    ψ Ω,R,Z( ) = exp −ikZΩ2( )

    2π L2

    × drr plane∫∫ exp ik W r( )+Ω⋅ r( )( )exp − r − R −ΩZ( )

    2

    2L2⎛

    ⎝⎜⎞

    ⎠⎟.

    W r( ) = −2RS logr, RS =2GMc2

    k ≡ κRS, R ≡ Rsρ,θR( ), Z ≡ RSζ , r ≡ 2Zk t,φr

    ⎝⎜

    ⎠⎟

    ψ ρ,ζ ,κ( ) = 2 dtt exp i −2κ log t + t2( )( ) J00

    ∫ tρ2κζ

    ⎝⎜⎞

    ⎠⎟

  • 8

    in which here and hereafter we neglect all external phase factors irrelevant to

    the wave intensity.

    The integral can be expressed in terms of confluent hypergeometric

    functions [7, 20]:

    . (3.4)

    The intensity falls from the initial value, which for k>>1 is

    , (3.5)

    to the asymptotic value unity for large r.

    Of the variety of asymptotic approximations for confluent

    hypergeometric functions [13] (or equivalently Whittaker functions [21]), the

    transitional asymptotics (regime 3) relevant to astronomy is obtained by

    applying large k stationary-phase approximation to the exponential factor in

    (3.3) and using this saddle value t=√k in the Bessel factor. Thus

    . (3.6)

    As the dashed curves in figure 2 show, the transitional approximation accurately

    describes more interference oscillations as k increases, that is, as l decreases for

    fixed Z.

    ψ ρ,ζ ,κ( ) = 2πκ1− exp −2πκ( ) 1F1 iκ ,1,i

    ρ2κ2ζ

    ⎛⎝⎜

    ⎞⎠⎟

    ψ 0,ζ ,κ( ) 2 ≈ ψ 0 ζ ,κ( ) = 2πκ

    ψ ρ,ζ ;κ( ) 2 ≈ 2πκ J02 ρκ2ζ

    ⎝⎜⎞

    ⎠⎟

  • 9

    Figure 2. Focal spot intensity pattern for the indicated parameter values. Full red curves: exact intensity (3.4); dashed curves: Bessel transitional approximation (3.6) (regime 3); dotted curves: geometric wave approximation (regime 2); the full line is the asymptotic limiting intensity 1

    The canonical star-lensing function is shown in figure 3a (figures

    3b,3c show the fold and cusp canonical functions to be calculated in section 4).

    Figure 3. Canonical functions representing diffraction intensity, with indicated principal

    fringe separations X for (a) focal spot: ; (b) fold caustic: ; (c) axial section of cusp |P(c)|2

    A measure of the fringe spacing is the distance Xstar between the focus and the

    nearest maximum. Thus, from the argument of J0, the spacing in scaled units is

    0 1 2 3 4 50

    2

    4

    6

    0 1 2 30

    4

    8

    12

    0 1 20

    10

    20

    30

    0.0 0.4 0.8 1.20

    20

    40

    60

    |ψ|2

    κ=1 κ=2

    κ=5 κ=10

    ρ/√ζ

    J02 c( )

    -5 0 50.0

    0.4

    0.8

    -6 -4 -2 0 2 40.0

    0.1

    0.2

    0.3

    -8 -6 -4 -2 0 2 4 60

    2

    4

    6a b cXstar Xfold Xcusp

    c

    |ψ|2

    J02 c( ) Ai c( )2

  • 10

    . (3.7)

    The intensity amplification factor is 2pk.

    The transitional approximation (3.6) is different from the geometrical

    wave approximation; this is obtained from (3.3) by using the asymptotic

    approximation for J0 in the integrand, in the form of a cosine, separated into two

    exponentials, and then evaluating the integral by stationary phase [20]. The

    resulting intensity is shown as the dotted curves in figure 2. Of course the

    approximation fails near the focus at r=0, but it agrees with the exact intensity

    very well everywhere else. In particular, it reaches the necessary large r limit

    unity (as an elementary exercise confirms) – unlike the transitional

    approximation (3.6), which by design works from r=0 out to a finite value r(k)

    that increases with k.

    4. Caustic diffraction from a binary star

    4.1. Diffraction integral, caustic, geometrical optics

    For an equal-mass binary with separation 2a, the phase in (2.3) is

    . (4.1)

    It is convenient to scale distances in the observation and lens planes by a,

    distances Z by a2/RS, and wavenumber k by RS:

    . (4.2)

    Astronomically, k>>1 always. We will often specialise formulas for general z

    to the asymptotic regime z>>1, because this gives relatively simple formulas

    Δρstar =X starκ

    ζ2, X star = 3.832

    W r( ) = −RS log x − a( )2 + y2( ) x + a( )2 + y2( )⎡⎣ ⎤⎦

    k ≡ κRS, x, y( ) ≡ a u,v( ), X ,Y( ) ≡ a ξ,η( ), Z ≡ a

    2

    RSζ

  • 11

    and corresponds to many cases of interest (for a binary consisting of two solar-

    mass stars (RS~3km) separated by 1au (i.e a=0.5au), z>1 corresponds to

    Z>65pc). With the scaling (4.2), the binary star lensed wave becomes

    (4.3)

    For large k, the wave is dominated by the caustics, where real saddles of

    the phase S coincide. As is known [1], for constant z sections the caustic

    consists of fold curves and cusp points forming closed astroids: two for z

  • 12

    Figure 4. Caustics from a binary-star lens, for the distances from the observation plane (a) z =2.05, (b) z=4; the red dots indicate points on the fold and cusp near which the local approximations are calculated in sections 4.2 and 4.3.

    The saddle-points of the integrand in (4.3) determine the ray (x, h) in the

    observation plane that originates from the point (u,v) in the lens plane, and, by

    inversion, the lens plane points from which rays reach given observation points:

    -1 0 1-4

    -2

    0

    2

    4

    -1 0 1-8

    0

    8

    4

    -4

    a b

    ξ

    fold

    cusp1cusp2

    cusp3

    η

  • 13

    (4.4)

    There are five saddles, all real within the astroids and three real and two

    complex outside. The caustic is determined by their coalescence, where the

    determinant of the Hessian matrix

    (4.5)

    vanishes:

    . (4.6)

    Explicitly, the caustic is given parametrically in terms of w by the real values of

    (4.7)

    The transitional approximations (regime 3) that will be calculated in the

    following sections are those associated with the singularities illustrated in figure

    4: the fold and cusp1 and cusp2 along the h axis, and cusp3 along the x axis.

    Explicitly, these are located at

    ∂u S = 0⇒ ξ = u 1−4ζ w2 −1( )1+ w2( )2 − 4u2

    ⎝⎜⎜

    ⎠⎟⎟,

    ∂v S = 0⇒η = v 1−4ζ w2 +1( )1− w2( )2 + 4v2

    ⎝⎜⎜

    ⎠⎟⎟, w = u2 + v2

    ⇒ u = uj ξ,η,ζ( ),u = uj ξ,η,ζ( ){ }, j = 1…5( ).

    M =∂uu S ∂uv S∂vu S ∂vv S

    ⎝⎜⎜

    ⎠⎟⎟

    detM = 0⇒w2 −1( )2 + 4u2( )w2 +1( )2 − 4u2( )2

    =w2 +1( )2 − 4v2( )w2 −1( )2 + 4v2( )2

    = 116ζ 2

    u± w,ζ( ) = ± ζ 32 1+ w4 + 2ζ 2( ) + 1+ w2( )2 + 8ζ 2 ,v± w,ζ( ) = ± ζ 32 1+ w4 + 2ζ 2( ) − 1− w2( )2 − 8ζ 2 .

  • 14

    (4.8)

    where in each case we also give the limiting forms: near where the caustic is

    born (z=1/4 for the fold, z=2 for cusp1 and cusp2, and z=0 for cusp3), and far

    way, i.e. z>>1. We will also need the points u, v in the lens plane from which

    these caustics originate:

    (4.9)

    In all cases, the procedure will be the same: locally expanding the phase S in

    (4.3) about the caustic points x, h, u, v.

    Although we are aiming for the transitional approximations, in

    subsequent sections it will be convenient also to use the geometrical wave

    (regime 2), as an useful check on the calculations. This is

    ηfold ζ( ) = 2 ζ ζ + 2( ) −1− 2ζ ζ ζ + 2( ) −1+ζ( ) = 163 3 ζ −

    14( )3/2 +O ζ − 14( )5/2( ) = 2ζ −1+ 14ζ −1 +O ζ −2( ),

    ηcusp1 ζ( ) = −2 ζ ζ − 2( ) −1+ 2ζ ζ ζ − 2( ) −1+ζ( ) = 3 + 2 23 ζ − 2 +O ζ − 2( ) = 2ζ −1− 34ζ −1 +O ζ −2( ),ηcusp2 ζ( ) = 2 ζ ζ − 2( ) −1+ 2ζ − ζ ζ − 2( ) −1+ζ( ) = 3 − 2 23 ζ − 2 +O ζ − 2( ) = 1/ ζ +O ζ −3/2( ),ξcusp3 ζ( ) = 2 ζ ζ + 2( ) +1+ 2ζ − ζ ζ + 2( ) +1+ζ( ) = 1−ζ +O ζ 3/2( ) = 1/ ζ +O ζ −3/2( ),

    ufold = 0, vfold ζ( )= − 2 ζ ζ + 2( ) −1− 2ζ ,

    ucusp1 = 0, vcusp1 ζ( )= − −2 ζ ζ − 2( ) −1+ 2ζ ,

    ucusp2 = 0, vcusp2 ζ( )= − 2 ζ ζ − 2( ) −1+ 2ζ ,

    vcusp3 = 0, ucusp3 ζ( )= 2 ζ ζ + 2( ) +1+ 2ζ .

  • 15

    , (4.10)

    in which

    (4.11)

    Even though the geometric wave fails on the caustics, it gives an easy global

    picture of the interference structure, as figure 5 illustrates.

    Figure 5. Geometrical wave (regime 2) (4.10) for k=200, z=5, showing the central astroid

    It is also interesting to consider the geometric optics limit, i.e. regime 1.

    Here there is a slight subtlety, arising from the fact that along symmetry axes

    two or more rays may contribute with the same phase and so their contributions

    ψ geom ξ,η,ζ ,κ( ) =exp iκ S j + 14 iπµ j( )

    ζ detM jj=1,real rays

    5

    S j = S uj η,ζ( ),v j η,ζ( ),ξ,η,ζ( ), M j = M u=u j ξ ,η,ζ( ),v=v j ξ ,η,ζ( )( ),µ j = sgn eigenvalues M j( )( )

    1,2∑ .

    -0.6 0.60-0.6

    0.6

    0

    ξ

    η

  • 16

    must be added coherently, with each intensity multiplied by a degeneracy factor

    dj (1 or 2 in our examples. The geometric intensity is thus

    . (4.12)

    Figure 6 illustrates the accuracy of the geometric wave approximation (regime

    2), and how geometric optics (regime 1) describes the mean intensity,

    everywhere except near the caustics, to the precise description of which we now

    turn.

    Figure 6. Wave intensity for h=0, z=1, k=20, including two cusp caustics (cf. figure 5). Red

    curve: exact, calculated numerically from (4.3); full black curve: geometric optics

    approximation (4.10) (regime 1); dashed black curve: geometric wave approximation (4.8)

    (regime 2)

    4.2. Fold diffraction

    The calculation will correspond to the red dot with the largest value of h on the

    fold caustic in figure 4a. As z increases, this caustic originates at z=1/4 and then

    Igeom ξ,η,ζ ,κ( ) =d j

    ζ 2 detM jj=1,real rays

    5

    -2 -1 0 1 20

    10

    20

    30

    40

    ξ

    |ψ|2

  • 17

    recedes (figure 4b). To get the transitional approximation we expand the phase

    in (4.3) about the caustic points in the lens and observation planes in (4.8) and

    (4.9):

    (4.13)

    in which K is an irrelevant constant and

    . (4.14)

    The u integral is Gaussian, and the contribution to the diffraction integral close

    to the fold at hfold, from the dv integral in the neighbourhood of ufold=0, v=vfold,

    is the anticipated Airy function [13, 22]

    , (4.15)

    where

    (4.16)

    and

    , (4.17)

    with k1/3Efold(z)2 giving the intensity amplification at the fold caustic.

    The separation of the first two intensity maxima (cf. figure 3b) is thus

    . (4.18)

    S u,vfold ζ( )+δv,0,ηfold ζ( )+δη,ζ( ) = K − δvδηζ −13δv

    3A ζ( )+ u2

    ζ!

    A ζ( ) = 12ζ 2

    2 ζ 2 +ζ( ) −1− 2ζ ζ 2 +ζ( ) 1+ζ( )+ζ 2 +ζ( )( )

    ψ fold δη,ζ ,κ( ) =κ 1/6Efold ζ( )Ai δηκ 2/3Ffold ζ( )( )

    Ffold ζ( ) =1

    ζ A ζ( )1/3= 23 ζ − 14( )( )1/6

    1+O ζ − 14( )( ) = 1ζ 1+O1ζ

    ⎛⎝⎜

    ⎞⎠⎟

    ⎛⎝⎜

    ⎞⎠⎟

    Efold ζ( ) =πζ

    1A ζ( )1/3

    = π3 ζ − 14( )( )1/6

    1+O ζ − 14( )( ) = πζ 1+O1ζ

    ⎛⎝⎜

    ⎞⎠⎟

    ⎛⎝⎜

    ⎞⎠⎟

    Δηfold =X foldκ

    −2/3

    D ζ( ) ≈ζ >>1X foldκ−2/3ζ , X fold = 2.2294

  • 18

    This shows that in the short-wave limit the fold intensity grows as k1/3, and the

    fringes get smaller as k–2/3 , as anticipated from catastrophe optics [5]; the

    fringes also expand linearly as the lens distance z increases. For z>2 this fold

    lies on an astroid (figure 4) that gets smaller as z increases: its size (distance

    from the fold down to cusp 1) is close to 1/z (cf. (4.8). The local approximation

    (4.15) and (4.16) is valid only while the fold and cusp are well separated, i.e. the

    distance between these singularities exceeds the fringe separation. This requires

    . (4.19)

    These are the main results of this section. The Airy function emerges

    clearly only for very large k. This is because there is an additional and non-

    negligible contamination from the three geometrical rays that are separate from

    the two that coalesce on the fold. This contamination gets slowly weaker as

    k increases, but is still visible as very fast oscillations even for k=108 (figure

    7a). In any practical situation these are likely to be averaged away by

    decoherence, leaving the Airy approximation emerging clearly; numerical

    calculations (not shown) confirm this. This phenomenon, of diffraction

    catastrophes being contaminated by fast oscillations that are often only

    imperfectly resolved in practice, is familiar in several contexts, including the

    quantum scattering of atoms and molecules [23, 24] and optical rainbows [25,

    26].

    ζ < κ1/3

    X fold

  • 19

    Figure 7. Wave intensity close to (a) fold; (b) cusp1; (c) cusp2; (d) cusp3, corresponding to red dots in figure 4a for z=2.05, k=108; the red curves are the transitional approximations (regime 3) of (a) sections 4.1, and (b,c,d) section 4.2; the black curves show the geometric

    wave approximation (regime 2). In (a) the inset shows the finest two interference scales of the rays contaminating the fold caustic.

    4.3. Cusp diffraction

    The procedure for all three cusps is the same; we give the details for cusp3, i.e.

    the cusp on the x axis (figure 4a), and then state the results for cusp1 and cusp2.

    To get the transitional approximation for cusp3, we expand the phase in (4.3)

    about its points x=xcusp,h=0, in the observation plane (equation (4.8)), and

    u=ucusp3, v=0 in the lens plane (equation (4.9)):

    (4.20)

    in which the second equality follows after integrating over du, and

    a

    3.2003 3.2005 3.20070

    50

    100

    2.140 2.150 2.1600

    4000

    8000

    1.400 1.405 1.4100

    10000

    20000

    0.5548 0.5552 0.5556 0.55600

    40000

    80000

    2.5x10-6b

    c d

    |ψ|2

    η

    η

    η

    ξ

    S ucusp3 ζ( )+δu,v,ξcusp3 ζ( )+δξ,0,ζ( ) = K + A3v

    4 +δu −δξζ

    +C3v2⎛

    ⎝⎜⎞⎠⎟+ δu

    2

    ζ+!

    = K + A3 − 14C32ζ( )v4 + 12δξC3v2 +!

  • 20

    (4.21)

    Thus the contribution to the diffraction integral close to xcusp3, from the

    neighbourhood of v=0, u=ucusp is

    . (4.22)

    P is the quartic-phase diffraction catastrophe integral (Pearcey function [13,

    27]) describing waves along the symmetry axis of the cusp and illustrated in

    figure 3c:

    (4.23)

    Here the z-dependent functions in the argument of P, and the prefactor, are

    (4.24)

    and

    A3 =18ζ 2

    1−ζ −ζ 2 +ζ ζ 2 +ζ( )( ),C3 =

    12ζ 2

    1+ 2ζ + 2 ζ (2 +ζ ζ 2 +ζ( ) ζ +1( )−ζ ζ + 2( )( ).

    ψ cusp3 δξ,ζ ,κ( ) = E3 ζ( )κ 1/4P δξκ 1/2F3 ζ( )( )

    P c( ) = dt exp i t 4 + ct2( )( )−∞

    = π c8exp − 18 ic

    2( ) exp 18 iπ( ) J− 14 18 c2( )− sgn c( )exp − 18 iπ( ) J 14 18 c2( )( ).

    F3 ζ( ) =C3

    2 A3 − 14C32ζ

    = 1ζ 3/2

    −ζ ζ + 2( )+ ζ +1( ) ζ ζ + 2( )( )× 1+14ζ + 22ζ 2 + 8ζ 3 + 2 2 + 7ζ + 4ζ 2( ) ζ ζ + 2( )= 2

    ζ1+O ζ( )( ) = 2 1+O ζ −1( )( ).

  • 21

    (4.25)

    The intensity amplification at cusp3 is k1/2E3(z)2. The separation of the

    first two intensity maxima (cf. figure 3c) is

    . (4.26)

    The k-1/2 dependence was anticipated from catastrophe optics, but the fact the

    for large z the fringe separation is independent of z was unexpected. An

    analogous phenomenon occurs in optical ‘diffractionless beams’ [28-31].

    For z>2, cusp3 lies on an astroid (figure 4) that gets smaller as z

    increases: its size (distance across to the corresponding cusp on the other side of

    the astroid, i.e. for x>112 X cuspκ

    −1/2 , X cusp = 3.3770

    2 / ζ

    ζ < 16κX cusp2

  • 22

    Figure 8. Wave intensity across cusp3 for z=1 and (a) k=20, (b) k=200, (c) k=1000, (d) k=106, showing emergence of the transitional approximation (4.23) (red curves) as k

    increases; the full black curves show the geometric optics approximation (regime 1), and the dashed black curves show the geometric wave approximation (regime 2)

    For cusp1 and cusp2, the procedure is the same, with h and x replacing

    x and h, and v and u replacing u and v. For the inward-pointing cusp1, the

    counterparts of (4.23)-(4.29) are

    (4.28)

    where

    (4.29)

    and

    0.0 0.2 0.4 0.6 0.8 1.00

    1

    2

    3

    4

    0.5 0.6 0.70

    0.05

    0.10

    0.60 0.62 0.64 0.66 0.680

    0.005

    0.010

    0.679 0.680 0.681 0.682 0.68300.678

    1x10-7

    2x10-7

    3x10-7

    0.70

    0.8

    |ψ|

    ξ

    b

    c d

    a

    ψ cusp1 δη,ζ ,κ( ) = E1 ζ( )κ 1/4P −δηκ 1/2F1 ζ( )( ),

    F1 ζ( ) =1

    ζ 3/2ζ ζ − 2( )+ ζ −1( ) ζ ζ − 2( )( )×

    −1+14ζ − 22ζ 2 + 8ζ 3 − 2 2 − 7ζ + 4ζ 2( ) ζ ζ − 2( )= 12 3 ζ − 2( ) 1+O ζ − 2( )( ) = 1ζ 1+O ζ −1( )( ),

  • 23

    (4.30)

    The intensity amplification at cusp1 is k1/2E1(z)2. The separation of the

    first two intensity maxima (cf. figure 3c) is

    . (4.31)

    This increases with z, and gets much bigger than the separation (4.26) for

    cusp3. The validity condition, analogous to (4.19) for the fold and (4.27) for

    cusp3, is that cusp1 is well separated from the fold on the same astroid:

    . (4.32)

    For the outward-pointing cusp2,

    , (4.33)

    where

    (4.34)

    and

    E1 ζ( ) =1

    π ζ 1− 4ζ + 2ζ 2 + 2 ζ −1( ) ζ ζ − 2( )( )( )1/4

    = 121/4 π

    1+O ζ − 2( )( ) = 12πζ −3/4 1+O ζ −1( )( ).

    Δηcusp1 =X cusp

    F1 ζ( )κ 1/2≈

    ζ >>1ζ 1/2X cuspκ

    −1/2 , X cusp = 3.3770

    ζ < κ1/3

    X cusp2/3

    ψ cusp2 δη,ζ ,κ( ) = E2κ 1/4P δηκ 1/2F2 ζ( )( )

    F2 ζ( ) =1

    ζ 3/2−ζ ζ − 2( )+ ζ −1( ) ζ ζ − 2( )( )×

    −1+14ζ − 22ζ 2 + 8ζ 3 + 2 2 − 7ζ + 4ζ 2( ) ζ ζ − 2( )= 12 3 ζ − 2( ) 1+O ζ − 2( )( ) = 2 1+O ζ −1( )( ),

  • 24

    (4.35)

    The intensity amplification at cusp2 is k1/2E2(z)2. The separation of the

    first two intensity maxima (cf. figure 3c) is

    . (4.31)

    This is asymptotically independent of z, like the separation (4.26) for cusp3.

    The validity condition for these oscillations is the same as for cusp3, on whose

    astroid it lies, namely

    . (4.32)

    Figures 7b and 7c illustrate the accuracy of the transitional approximation

    near cusp1 and cusp2. For all three cusps, the fast-oscillating contamination by

    interfering waves not associated with the caustic is much weaker than for the

    fold (figure 7a). This because there are only two such contaminating rays, rather

    than three.

    5. Concluding remarks

    The calculations reported here are intended to indicate the fringe separations

    and intensity amplifications that could occur in gravitational lensing. For

    convenience, Table I presents the results in terms of physical variables. The

    wavelength dependences are those anticipated from singularity theory, but the

    dependences on astronomical lengths could not have been anticipated without

    E2 ζ( ) =1

    π ζ 1− 4ζ + 2ζ 2 − 2 ζ −1( ) ζ ζ − 2( )( )( )1/4

    = 121/4 π

    1+O ζ − 2( )( ) = 2πζ 1/4 1+O ζ −1( )( ).

    Δηcusp2 =X cusp

    F2 ζ( )κ 1/2≈

    ζ >>112 X cuspκ

    −1/2 , X cusp = 3.3770

    ζ < 16κX cusp2

  • 25

    detailed calculation. In particular the dependences of fringe spacing on

    Schwartzschild radius, binary separation, and lens distance, are very different

    for the inward-pointing cusp1 and the outward-pointing cusp2 and cusp3.

    Further similar calculations can be envisaged, for example of the fringe

    spacings across a cusp, rather than along its symmetry axis; these spacings are

    much smaller: O(l3/4) rather than O(l1/2) [5] (an example of this diffraction

    anisotropy is the longitudinal stretching of fringes in cusps seen by people

    wearing eyeglasses when looking at street lights on rainy nights; the images are

    distorted by raindrop ‘lenses’ on the glass lenses [11, 14, 32]; the apparent

    oxymoron that these fringes appear black-and-white, i.e. not coloured, is a

    psychovisual effect [33]).

    Table 1. Fringe separations, intensity amplifications and validity conditions, expressed in

    physical variables, for Z>>RS for star and ZRS/a2>>1for fold and cusps; Xfold=2.2294,

    Xcusp=3.3770,

    Two main effects threaten the detection of wave effects such as those

    anticipated here. The fringes in the observation plane could be blurred by the

    finite size of the distant source; its angular size, seen from the lens plane, should

    be smaller than the angle subtended at the lens plane by these fringes, i.e.

    fringe size amplification validity

    star X starλ23/2π

    ZRS

    4π 2 RSλ

    fold X foldλ2/3RS

    1/3Z2π( )2/3 a

    π Aimax2 a2

    Z2πλRS

    2

    ⎛⎝⎜

    ⎞⎠⎟

    1/3

    Z < 2πa2

    X fold λRS2( )1/3

    cusp 1 X cuspZλ2π

    Pmax2 a3

    RS 2πλZ3

    Z < a2 2πX cusp

    2 RS2λ

    ⎝⎜⎞

    ⎠⎟

    1/3

    cusps 2,3X cuspa23/2

    λπRS

    2 Pmax2 RSa

    2Zπλ

    Z < 32πa2

    λ

    Aimax2 = 0.2869, Pmax

    2 = 6.9438

  • 26

    (spacings in Table I)/Z. Of course, the size of the aperture of the telescope

    directed at the source is also a limiting factor; I have not discussed this here,

    apart from briefly at the end of section 2 (see also [19]).

    Acknowledgements. My research is supported by a Leverhulme Trust Emeritus

    Fellowship.

    References

    1. Petters, A.O., H. Levine, and J. Wambsganss, 2001,Singularity theory and gravitational lensing. Boston: Birkhäuser.

    2. Treu, T., P.J. Marshall, and D. Clowe, 2012,Resource Letter GL-1: Gravitational Lensing. Am. J. Phys. 80: p. 753-763.

    3. Bartelmann, J. and P. Schneider, 2001,Weak Gravitational Lensing. Phys. Repts 340: p. 291-472.

    4. Schneider, P., C. Kochanek, and J. Wambganss, 2006,Gravitational Lensing: Strong, Weak and Micro. Saas-Fee Advanced Course 33. Springer.

    5. Berry, M.V. and C. Upstill, 1980,Catastrophe optics: morphologies of caustics and their diffraction patterns. Progress in Optics 18: p. 257-346.

    6. Kravtsov, Y.A. and Y.I. Orlov, 1993,Caustics, Catastrophes and Wave Fields. Springer Series on Wave Phenomena. Berlin & Heidelberg: Springer-Verlag.

    7. Nakamura, T.T. and S. Deguchi, 1999,Wave Optics in Gravitational Lensing. Proc. Theor. Phys. Supp. 133: p. 137-153.

    8. Feldbrugge, J. and N. Turok, 2020,Gravitational lensing of binary systems in wave optics. arXiv:2008.01154v1 [gr-qc].

    9. Feldbrugge, J., U.-L. Pen, and N. Turok, 2019,Oscillatory path integrals for radio astronomy. arXiv:1909.04632v1 [astro-ph.HE].

    10. Berry, M.V., 1977,Focusing and twinkling: critical exponents from catastrophes in non-Gaussian random short waves. J. Phys. A 10: p. 2061-81.

    11. Nye, J.F., 1999,Natural focusing and fine structure of light: Caustics and wave dislocations. Bristol: Institute of Physics Publishing.

    12. Lee, R. and A. Fraser, 2001,The rainbow bridge: rainbows in art, myth and science. Bellingham, Washington: Pennsylvania State University and SPIE press.

    13. http://dlmf.nist.govDLMF, 2010,NIST Handbook of Mathematical Functions, ed. F.W.J. Olver, et al. Cambridge: University Press.

  • 27

    14. Berry, M.V., 1976,Waves and Thom's theorem. Advances in Physics 25: p. 1-26.

    15. Chester, C., B. Friedman, and F. Ursell, 1957,An extension of the method of steepest descents. Proc. Camb. Phil. Soc. 53: p. 599-611.

    16. Berry, M.V., 1969,Uniform approximation : a new concept in wave theory. Science Progress (Oxford) 57: p. 43-64.

    17. Kravtsov, Y.A., 1968,Two new asymptotic methods in the theory of wave propagation in inhomogeneous media. Sov. Phys. Acoust. 14: p. 1-17.

    18. Nambu, Y., 2013,Wave optics and image formation in gravitational lensing. J. Phys. Conf. Ser. 410: p. 012036.

    19. Berry, M.V., 2007,Looking at colaescing images and poorly resolved caustics. J. Opt. A. 9: p. 649-657.

    20. Herlt, E. and H. Stephani, 1976,Wave Optics of the Spherical Gravitational Lens Part I: Diffraction of a Plane Electromagnetic Wave by a Large Star. Int. J. Theor. Phys. 15: p. 45-65.

    21. Dingle, R.B., 1973,Asymptotic Expansions: their Derivation and Interpretation. New York and London: Academic Press.

    22. Vallée, O. and M. Soares, 2010,Airy functions and applications to physics (2nd Ed.). London: Imperial College Press.

    23. Hundhausen, E. and H. Pauly, 1965,Der Regenbogeneffekt und Interferenzerscheinungen bei molekularen Stössen. Z. für Phys. 187: p. 305=337.

    24. Tabor, M., 1977,Fast oscillations in the semiclassical limit of the total cross section. J. Phys. B 10: p. 2649-2662.

    25. Laven, P., 2003,Simulation of rainbows, coronas and glories by use of Mie theory. Appl. Opt. 42: p. 436-444.

    26. Nussenzveig, H.M., 1992,Diffraction effects in semiclassical scattering. Cambridge: University Press.

    27. Pearcey, T., 1946,The structure of an electromagnetic field in the neighbourhood of a cusp of a caustic. Phil. Mag. 37: p. 311-317.

    28. Durnin, J., 1987,Exact solutions for nondiffracting beams. I. The scalar theory. J. Opt. Soc. Amer. 4: p. 651-654.

    29. Durnin, J., J.J. Miceli Jr, and J.H. Eberly, 1987,Diffraction-free beams. Phys. Rev. Lett. 58: p. 1499-1501.

    30. Siviloglou, G.A., J. Broky, A. Dogariu, and D.N. Christodoulides, 2007,Observation of accelerating Airy beams. Phys. Rev. Lett 99: p. 213901.

    31. Berry, M.V. and N.L. Balazs, 1979,Nonspreading wave packets. Am. J. Phys. 4: p. 264-7.

    32. Berry, M.V., Singularities in Waves and rays, in Les Houches Lecture Series Session 35, R. Balian, M. Kléman, and J.-P. Poirier, Editors. 1981, North-Holland: Amsterdam. p. 453-543.

  • 28

    33. Berry, M.V. and A.N. Wilson, 1994,Black-and-white fringes and the colours of caustics. Applied Optics 33: p. 4714-4718.