Open Access Paper
29 July 2016 Telescope polarization and image quality: Lyot coronagraph performance
Author Affiliations +
Abstract
In this paper we apply a vector representation of physical optics, sometimes called polarization aberration theory to study image formation in astronomical telescopes and instruments. We describe image formation in-terms of interferometry and use the Fresnel polarization equations to show how light, upon propagation through an optical system become partially polarized. We make the observation that orthogonally polarized light does not interfere to form an intensity image. We show how the two polarization aberrations (diattenuation and and retardance) distort the system PSF, decrease transmittance, and increase unwanted background above that predicted using the nonphysical scalar models. We apply the polarization aberration theory (PolAbT) described earlier (Breckinridge, Lam and Chipman, 2015, PASP 127, 445-468) to the fore-optics of the system designed for AFTA-WFIRST– CGI to obtain a performance estimate. Analysis of the open-literature design using PolAbT leads us to estimate that the WFIRST-CGI contrast will be in the 10-5 regime at the occulting mask. Much above the levels predicted by others (Krist, Nemati and Mennesson, 2016, JATIS 2, 011003). Remind the reader: 1. Polarizers are operators, not filters in the same sense as colored filters, 2. Adaptive optics does not correct polarization aberrations, 3. Calculations of both diattenuation and retardance are needed to model real-world telescope/coronagraph systems.

1.

INTRODUCTION

An optical system corrected for geometrical path different errors is a necessary but not sufficient condition for the perfect image formation needed to directly image terrestrial exoplanets. Geometric (trigonometric) path difference errors are controlled using adaptive optics (tip-tilt & wavefront), active metrology and precision pointing. However, image quality is also determined by several physical optics factors: diffraction, polarization, partial coherence, chromatism lead to poor image formation, and is not corrected through the control of geometric path difference. We show that the source of physical optics errors lies in the opto-mechanical packaging of optical elements, masks, stops and the thin film coatings needed to obtain high transmittance. Adaptive optics corrects wavefront errors described by geometric or optical path length errors but not those wavefront errors introduced by physical optics.

Breckinridge Kuper & Shack1 first (1984) analyzed the use of the Lyot coronagraph to image distant exoplanets and modeled a system using the scalar approximation to the vector electromagnetic wave. They showed how complex apodization of the exit pupil reveals an exoplanet in the presence of mirror fabrication errors. Earlier (1971) Breckinridge2 showed that polarization internal to instruments contribute to spectrophotometric errors. Breckinridge & Oppenheimer3 (2004) showed that internal polarization plays an important role in exoplanet coronagraphy. Carson, et. al.4 provided a measurement of the polarization dependent PSF. Breckinridge5 alerted the WFIRST-CGI science and technology development team to contrast degradation caused by internal polarization.

Today, ground and space exoplanet coronagraphs are designed and built under the assumption that the scalar wave approximation to the vector electromagnetic wave6, 7, is adequate. Shaklan ET. Al. 8 examined a terrestrial planet finder (TPF) coronagraph design using vector electromagnetic (E&M) waves and concluded that for TPF designs vector-waves were not necessary to develop a system to control scattered light to the level required at that time. Recently, Breckinridge9 and Chipman10 used vector E&M wave analysis with the polarization aberration tools developed by Chipman and others to model point-spread functions (PSF) for astronomical telescopes and discovered that several physical optics effects are in reality very important for the design of high performance coronagraphs.

In this paper we examine the role physical optics has in the direct imaging of exoplanets and suggest strategies to minimize the negative effects of aberrations introduced by physical optics. This text is divided into 5 sections. Section 1 is the introduction. Section 2 discusses image formation in the presence of polarization and reviews scalar wave image formation, shows a dramatic example of what happens to the PSF in an extreme case of polarization variations across an exit pupil, provides a review of the Fresnel polarization equations. Section 3 describes vector-wave image formation using the Fresnel-Kirchoff diffraction integral and shows the, in the presence of polarization aberrations the PSF is the linear superposition of four PSF’s. Section 4 summarizes polarization ray trace (PRT) and polarization aberration theory (PolAbT) and gives the detailed structure of the four PSF’s: two have shape similar to an Airy diffraction pattern, but larger in extent with one displaced relative to the other and the other two are fainter, but severely distorted. Conclusions from the analysis of a three-mirror bent Cassegrain are tabulated. In section 5 we apply our understanding of the physical optics properties of exoplanet coronagraphs to a discussion of how we might achieve maximum extinction at the focal plane by “impedance” matching the complex properties of the occulting mask to the E&M content of the complex field at the focal plane. Section six provides a brief polarization aberration analysis of the AFTA-WFIRST-CGI system.

2.

TELESCOPE/INSTRUMENT SYSTEM IMAGE FORMATION

Scalar wave image formation

In this section we provide a brief review of image formation modeled using scalar wave theory as a basis to extend the work into the more accurate vector representation.

Object space irradiance distribution can be decomposed into an ensemble of delta functions. The intensity or height of each delta function maps out the structure of the object. The optical system operates on the complex amplitude and phase associated with that intensity to form an image at the detector. Most astronomical sources in the visible region of the spectrum radiate broadband, incoherent thermal light.

The theory of image formation is developed using the schematic shown in Figure 1 below. The coordinate system we will use in our analysis is shown in Fig 1 below. This system is in standard use by modern textbooks11, 12 on the physics of image formation. The object plane (#1 in the system) is represented by Cartesian coordinates from the Latin alphabet x1 y1 the pupil plane (#2 in the System) is represented by

Cartesian coordinates from the Greek alphabet ξ2, η2 and the image plane (#3 in the System) is represented by Cartesian coordinates from the Latin alphabet x3, y3.

Figure 1

schematic of an optical system in the meridional plane. The object space complex field U1 (x1,y1) is shown as a delta function δ(x1,y1) to represent a star on-axis. This object space field is propagated by Fresnel diffraction to just to the left of the entrance pupil and is shown by 00030_psisdg9904_99041c_page_2_2.jpg. The pupil is shown to have a complex transmittance τ2(ξ2, η2) at plane 2. The optical system, shown here schematically as a lens of focal length f, provides the optical power to create the field U3 (x3, y3) at an image plane 3.

00030_psisdg9904_99041c_page_2_1.jpg

The scalar complex amplitude and phase across the image plane is found by standing at the image plane (#3) in Fig 1 and looking to the left, or back through the system toward the object. The Fresnel-Kirchoff diffraction integral is used to model the propagation of scalar electromagnetic waves through the optical system shown in Fig. 1. The image plane complex amplitude and phase field U3(x3, y3) at the image plane is given by

00030_psisdg9904_99041c_page_3_1.jpg

Where K is a constant, the integral is taken over the complex field across the exit pupil, 00030_psisdg9904_99041c_page_2_2.jpg of the optical system whose focal length is f, 00030_psisdg9904_99041c_page_3_5.jpg is the quasimonochromatic wavelength of light. The amplitude and phase complex properties across the exit pupil are contained in the scalar term,

00030_psisdg9904_99041c_page_3_2.jpg

Where A2 (ξ2, η2) varies between 0 and 1 and describes amplitude part of the complex wave as a function of position across the exit pupil. The phase properties at each point across the exit pupil are described by ϕ2 (ξ2, η2). Eq 1 is written for the scalar wave solution to Maxwell’s equations and the not vector wave

solution to Maxwell’s equation.

To the left in Fig 1, we have a point source represented by a delta function. This point source is mapped onto the image plane. We record intensity at the image plane and define the image plane irradiance distribution for this point source to be:

00030_psisdg9904_99041c_page_3_3.jpg

Next if we let the object space irradiance be represented by IObject(x1,y1) and the image space irradiance represented by IImage (x3, y3) and use the theoretical development of Goodman, we can write,

00030_psisdg9904_99041c_page_3_4.jpg

Where the symbol ⊗ denotes the convolution operator.

To understand the need for vector-wave physical optics, we need to review the source of polarized light within an optical system and understand the complex (amplitude and phase) wavefront at the focal plane U3(x3, y3)where coronagraphers place the occulting mask to control scattered light to one part in 1011.

Role of vector waves in image formation

An experiment using linear orthogonal polarizers and a telescope shows the role of vector waves in image formation. Figure 2 shows the effects of adding polarizers to an optical system: Top left shows an open, unmasked exit pupil of a telescope with zero geometric wavefront error. Top right shows the shape of the PSF recorded with the pupil on the top left. Bottom left shows the same telescope pupil as that shown in the upper left, with two linear polarizers over the top, one aligned orthogonally to the other. Horizontally polarized light is admitted to the left-hand side of the pupil and vertically polarized light is admitted to the right-hand side of the pupil. The bottom right shows the PSF recorded using the pupil on the bottom left. Note that with no polarizer the angular resolution is not position angle dependent, however, with the polarizer the angular resolution is angle dependent.

Figure 2

PSF’s shown for a telescope with zero geometric wavefront aberration without (upper and with (lower) polarizers.

00030_psisdg9904_99041c_page_4_1.jpg

Orthogonally polarized white light does not interfere to create an image. In Fig 2, the lower left image of the exit pupil the polarized radiation from the left portion of the exit pupil does not interfere with the orthogonally polarized radiation from the right portion of the exit pupil. Therefore the PSF is elongated in the horizontal direction. In this case the PSF is the scalar sum (linear superposition) of two images of a “D” shaped aperture, not the vector sum across the circular aperture shown in the upper right panel in Fig 2.

Although this is a rather dramatic example and no one would intentionally place orthogonal linear polarizers over their telescope pupil, this shows that any source of polarization change across the exit pupil will result in distortion of the PSF at some level.

In the next section we identify sources within an optical system that polarize light. Current astronomical science measurement objectives requires high transmittance optical systems, which in turn require high reflectivity broad-band optical thin films. As the white-light electromagnetic wave propagates through the optical system it becomes partially polarized. The orthogonal components of the partially polarized beam do not interfere to contribute to the image, but rather flood the focal plane with unwanted background radiation. The Fresnel polarization equations give the magnitude and sign of this polarization and are described in the following section.

Fresnel polarization

Here we examine the source of phase and amplitude changes within astronomical telescopes and instruments. Systems require mirrors coated with metals (e.g. Al or Ag) to give high surface reflectance and thus maximize system transmittance. These mirrors are overcoated with a dielectric material that serves two purposes: 1. Transparent mechanical barrier coat to inhibit oxidation and surface abrasion. 2. Enhance the reflectivity at select wavelengths. In the next section we show that these metal surfaces partially polarize light.

In figure 2, above we saw that the change in polarization across the exit pupil affects image quality. Broadband unpolarized white-light is a characteristic of nearly all-astronomical sources and is divided equally into two orthogonally polarized beams for the derivation below. We represent unpolarized light by two orthogonal Eigenvector states and, for this example we select linearly polarized Eigenvector states. The Fresnel equations1314 are used to model the behavior of a vector electromagnetic complex wave interacting with a metal or dielectric surface (mirror).

A-J Fresnel in 1823 described the theory for interactions of electromagnetic radiation with dielectrics and metals. These relationships were developed further15 and are the basis of the commercial field of ellipsometry16.

Consider incoherent white-light incident at angle θ0, onto a metal mirror with isotropic

properties. This metal mirror has a wavelength dependent complex index, N1 (λ) = n1 (λ) + ik1 (λ). The Eigenstates of reflection are the s (perpendicular) and p (parallel) polarized components. A portion of the beam reflects at the incidence angle θ0 (Snell’s Law) and another portion (a damped evanescent wave)penetrates a short distance into the metal at the complex refraction angle of θ1 given by Snell’s law17 and is absorbed to heat the metal. This complex angle is given by

00030_psisdg9904_99041c_page_4_2.jpg

The complex reflectivities for light in the p and s polarizations are given by33

00030_psisdg9904_99041c_page_5_1.jpg

Two polarization effects occur. 1. There is a phase shift between the waves associated with each of the two polarizations, Ψ, called retardance which is given by

00030_psisdg9904_99041c_page_5_2.jpg

Equation 6 gives us the retardance Ψ for a single ray propagating through the system. An image requires

an array of rays. When we trace multiple rays from a single point in object space, the tangent of the retardance becomes

00030_psisdg9904_99041c_page_5_3.jpg

Where (ξ, η)are coordinates across the pupil.

The reflectivity is polarization dependent, with the result that reflection acts as a partial polarizer and the diattenuation; D at each point (ξ, η) across the pupil is given by

00030_psisdg9904_99041c_page_5_4.jpg

Where r is the complex reflectivity given in Eq 5 above. Metallic reflection acts as a weak polarizer, and D varies from zero (nonpolarizing) to one for ideal polarizers. Astronomical optical systems require large Etendu (area times solid angle), which requires large optics. However, the volume for spacecraft bus is required to be compact to fit inside launch shrouds. These two requirements conflict and often lead to many fold mirrors in the instrument which, unless designed properly will, in turn, lead to large internal polarization with the concomitant loss in transmission and image quality.

Note there are two polarization aberrations: 1. Diattenuation is commonly used to model a polarization dependent reflectivity and 2. Retardance is used to model a polarization dependent change in the phase of the complex wave upon reflection.

3.

IMAGE FORMATION IN THE PRESENCE OF POLARIZATION

Vector wave image formation

In the previous section we examined the scalar diffraction equation 3, above. If the source is partially polarized and the telescope has a polarization dependent transmission, the Eq. 3 is written as:

00030_psisdg9904_99041c_page_5_5.jpg

The complex wavefront for unpolarized, incoherent astronomical sources becomes partially polarized upon propagation through an astronomical telescope and instrument that contains metal mirrors and dielectric surfaces. A Jones matrix whose values change for different ray-paths across the exit pupil describes this wavefront. This spatially dependent matrix is called the Jones pupil18, we write this short hand for the complex field at each point (ξ, η)across the exit pupil:

00030_psisdg9904_99041c_page_6_1.jpg

On the LHS of Eq. 10, A is amplitude and ϕ is phase of the electric field for each of the 4

component waves in an arbitrarily selected X,Y Eigen basis-set at points ξ, η across the exit pupil. Subscript XX refers to the complex field exiting polarized in X resulting from the incident field with X polarization, as matrix multiplication would imply. A similar convention extends to the subscripts YY, YX and XY. Ideally, the Jones pupil would be the identity matrix for all ray paths and no undesired polarization change would occur. 19 That is, the off-diagonal elements in the matrix shown in Eq.10 would be zero. During image formation with incoherent light, none of these four Jones pupil components form interference fringes with each other20,21. Each is diffracted separately by scalar diffraction theory to calculate the four components of the amplitude response matrix, which is the generalization of the amplitude response function of diffraction theory22.

The vector transmittance of the telescope 00030_psisdg9904_99041c_page_6_5.jpg is written,

00030_psisdg9904_99041c_page_6_2.jpg

The telescope/coronagraph system complex transmittance across the exit pupil depends on the vector of the electromagnetic field at point ξ, η within the exit pupil. The complex electric field u3 (x3, y3) at the image plane, for an on-axis unpolarized star of unit brightness follows from the Fresnel Kirchoff diffraction integral and is written:

00030_psisdg9904_99041c_page_6_3.jpg

Where K is a constant and we assume that the optical power of the system is not vector (polarization) dependent. To calculate the intensity I3(x3,y3) that we will measure, we take

00030_psisdg9904_99041c_page_6_4.jpg

And from Eq 12, we find the intensity to be

00030_psisdg9904_99041c_page_7_1.jpg

Since we are observing a star, which is a thermal broadband white-light source, it is reasonable to assume that the complex electric fields given by JXX, JYY, JYX and JXY are statistically uncorrelated and therefore incoherent. Consequently, Eq. 11 can be expanded to gives,

00030_psisdg9904_99041c_page_7_2.jpg

And we see that we have four point-spread functions, one each for the uncorrelated fields JXX, Jrr, JYX and JXY and the image plane point spread function is the linear, incoherent superposition

of four PSF’s as shown in Eq. 13 below.

00030_psisdg9904_99041c_page_7_3.jpg

4.

POLARIZATION RAY TRACE (PRT) AND POLARIZATION ABERRATION THEORY (POLABT)

Polarization ray trace (PRT)

Polarization ray tracing (PRT) is a technique for calculating the polarization matrices for ray paths through optical systems23,24,25,26,27,28. Polaris-M29 was built from the ground up to calculate polarization effects in optical systems. It is based on a 3x3 polarization ray tracing calculus46. Diffraction image formation of PolAb beams is then handled by vector extensions to diffraction theory30, 31, 32,33. A calculation of the polarization point spread matrix and optical transfer matrix can be seen in Section 4 of Ref. 42.

Polarization aberration theory (PolAbT)

Polarization aberration theory (PolAbT) describes the polarization effects of diattenuation, retardance, and apodization in a series expansion, where a cascade of terms separate mathematically the largest effects, from smaller effects and associate these polarization related image defects with constructional parameters and coating performance metrics.34,35 For example one term, retardance tilt, is strongly associated with fold mirrors and causes the XX and YY image components to shift with respect to each other, making the PSF slightly elliptical1. Another term retardance-defocus causes astigmatism from primary and secondary mirrors, which is polarization dependent; the orientation of the retardance rotates with the orientation of an incident linear polarization.36.

PRT generates very large files of numbers, at least eight times more than a conventional ray trace, leaving the designer and management a substantial data interpretation task of the aberrations represented in a higher dimensional polarization space. PolAbT is more difficult analytically than PRT, but it simplifies the ray tracing results into a small number of “terms” which are understood and addressed in an uncoupled manner. This enables us to manage polarization aberrations in more complicated systems, such as WFIRST-CGI.

A distinction between the two is seen in the comparison between classical geometric aberration ray trace (analogue to PRT) and the structural aberration coefficients37 (analogue to PolAbT) used by advanced designers to arrive quickly at near-optimized designs. Thus using PolAbT together with PRT is far more powerful than either method alone.

Polarization ray-trace

The output of a CAD ray-trace computer program is combined with Fourier optics to calculate point spread functions. Figure 3 shows a side view of a typical optical system with a fan of rays originating from a point on the object and passing through an optical system with k surfaces to the system exit pupil. Each ray strikes a real physical surface at a known angle of incidence (no paraxial approximation).

Figure 3

A fan of rays is shown passing from the object plane through an optical system with k surfaces before the exit pupil.

00030_psisdg9904_99041c_page_8_1.jpg

We know the physical properties of each surface that is a reflecting metal or dielectric. Using the Fresnel equations, discussed in section TBD we calculate values for each of the four complex entries in Eq. 10, for each ray intercept through the system. We compute the multiplicative amplitude and cumulative phases for both perpendicular and parallel light and map these into four arrays of complex numbers across the exit pupil. We then take a digital FFT to calculate the four PSF’s.

Three-mirror imaging telescope polarization ray trace (PRT)

The PolAb and diffraction image formation of a 3-mirror telescope system (fig 3 ref. 1) classical Cassegrain were analyzed with the Polaris-M software to evaluate the order of magnitude of polarization effects, which would be expected in a telescope/coronagraph. Polarization artifacts were discovered with “ghost PSFs” about twice the size of the Airy disk at 105 of the peak intensity. This 3-mirror system is a much simpler system than the 18-mirror WFIRST-CGI. PRT for an 18-mirror system is very labor intensive and we developed our theory based on the much simpler 3-mirror system and derived the more general PolAbT and used that to predict performance of the 18 mirrors (before the image plane stop) WFIRST-CGI.

For the three-mirror bent Cassegrain telescope, we adjusted the surface figures and vertex intervals so that geometric aberrations were zero for the on-axis PSF. All residual aberrations were therefore caused by polarization, not geometric aberrations. The primary mirror is 2.4 m dia. and F/# 1.2, the curvature on the secondary was adjusted to give a system F#/8. The position of the secondary along the system axis was adjusted so the F/# 8 converging beam reflected from a flat 45-degree mirror to deviate the beam by 90-degrees to a focal plane underneath the primary. Our calculations assumed bare Al at λ = 800 nm, where the complex index of refraction, N1 is N = 2.8 + 8.45i. Each surface contributes its own polarization for each ray that strikes it. The cumulative complex transmittance for each point in the exit pupil is found by multiplying the 3x3 polarization ray tracing matrices at each of the 3 surfaces38. This provides in 3D the exiting polarization state for arbitrary incident states. The Jones field at the exit pupil is Fourier transformed to find the E&M field at the focal plane for all incident polarization states. The modulus squared of this field contains four PSF components39.

This Cassegrain telescope is illuminated by a thermal white-light star on-axis. The image plane irradiance is found to be the linear superposition of the four amplitude PSF’s40 seen in Fig 4. Not visible at this scale: the centroid of each of the four are shifted slightly, one with respect to the other.

Figure 4,

a single star on axis in object space produces a four-element amplitude response matrix (ARM). Here we plot the amplitude for these complex elements of the 2x2 ARM, as they appear nearly super-posed at the focal plane of the Cassegrain telescope. The scale or size of each of these ARM elements is wavelength dependent.

00030_psisdg9904_99041c_page_9_1.jpg

The spatial extent of each of the off-diagonal ghost PSFs, IXY and IYX, are twice as large as the spatial extent of the on-diagonal images. The detailed analysis of these images shows several issues of concern for the WFIRST-CGI, which contains 15 more additional mirrors. The two principal images IXX and IYY are shifted by 0.625 mas with respect to each other due to a linear variation of retardance at the mirror – additional mirrors (like we have in CGI) in the path will increase this separation. Each of the principal images, XX and YY, is slightly astigmatic, but with opposite astigmatism sign (rotated 90°). The PolAb cause polarization crosstalk between X and Y polarized light, the off-diagonal elements, which although weak, 0.0037 in amplitude, 10-5 in flux, has a much larger extent than the Airy disks of the principal components - additional mirrors (like we have in CGI) will increase the flux in these “ghost PSF’s”

Primary and secondary mirror coatings induce astigmatism on-axis41 which couples light into the orthogonal polarization state in Maltese cross type patterns, yielding the ghost PSFs. Focusing through fold mirrors introduces a linear variation of retardance, putting a different linear phase shift on the two principal components, shifting them in opposite directions! We derived eighteen scaling relations or design rules for these system parameters using PolAb theory.

Fig 5 (top) below shows a plot of the log10 of the irradiance in the meridional plane at the image for IXX (the solid line) and IYX (the dotted line) while Fig 5 (lower) is a “face-on” map of IYX with the classic Airy diffraction rings superposed.

Figure 5

(upper) plots (solid line) of the logio of the irradiance in the meridional plane at the image for the irradiance distribution. In the upper part of the figure, the lower dotted line shows the logio irradiance for the irradiance distribution. The lower image in this figure is the “face-on” appearance of Iyx or the PSF Image of the off diagonal terms shown in Figs 4 and Eq. 10.

00030_psisdg9904_99041c_page_10_1.jpg

Note that the positions of the zeros in the PSF associated with IXX are not superposed on those associated with Iyx. Therefore ghost images could be misinterpreted as a “false alarm” candidate exoplanet.

Polarization changes across the PSF

To first order the optical system of a coronagraph is isoplanatic and the fore-optics PSF for the star is the same as that for the planet. The image plane mask suppresses the field from the star out to a distance of ∼3λ/D. In the case of the 2.4 m Cassegrain system at λ=800nm where 3λ/D ≈ 0.2 asec we find that for both the star and its displaced planet the radius r of the 90% encircled energy for rXX = rYY = 0.15 asec and rYY = rYX = 0.36 asec. Radiation from the cross-product terms extends well beyond the 0.2-asec coronagraph dark hole to create an irregular background pattern that will confuse exoplanet measurements.

Conclusions from the analysis of the 3 mirror bent Cassegrain

Breckinridge, Lam and Chipman (2015)42 and Chipman, Lam and Breckinridge (2015)43 used the POLARIS-M software to provide a detailed complex vector polarization analysis of a typical astronomical bent Cassegrain telescope comprised of an F/# 1.2 primary and a secondary mirror whose optical power is sufficient to give an F/#8 beam converging past a 45-degree flat mirror to a focal plane. The curvatures on the primary and secondary were designed to give zero geometric aberrations on axis. Coatings were assumed to be bare Al and wavelength 800nm. Several important facts about image quality in astronomical telescopes based on this this simple configuration were found910 to affect high-contrast exoplanet coronagraphy.

  • 1. The image plane PSF is the incoherent sum of 4 point nearly super-posed spread functions: two normalized near 1, but distorted PSF’s and two highly distorted “Ghost PSF’s”, with long “tails”, but a factor, γ fainter at the peak by approximately 10-5.

  • 2. The magnitude of this factor depends on the angles of incidence and the sign of the rays on each mirror in the path as well as the highly reflective coatings on the mirrors.

  • 3. The number of fold mirrors in the optical system determines the factor γ which increases as the square of the number of “aligned” fold mirrors. Crossed fold mirrors (s-p at first becomes p-s at the next) at the same angle of incidence can compensate polarization. “Aligned” here means the number of mirrors which share their s-p orientation minus the number with the opposite p-s orientation.

  • 4. The radius of the spatial extent of the 90% encircled energy of these two ghost PSF images is approximately twice as large as the radius of the two primary diffraction patterns.

  • 5. The PSF images for two orthogonal linearly polarization components are shifted with respect to each other, causing the PSF image for un-polarized point sources to become slightly elongated (elliptical) with a centroid separation on the order of 0.6 mas.

Applications

At least two astronomical science measurement objectives are affected by internal polarization. These are: 1 Exoplanet characterization using coronagraphy and 2 Astrometry. The astrometric implications are discussed by Safonov44 and will not be repeated here. Rather we will concentrate on applications to characterize exoplanets and discuss this topic in more detail in the next section.

5.

LYOT CORONAGRAPH TO CHARACTERIZE EXOPLANETS

Optical schematic

Figure 6 shows the optical schematic for a typical exoplanet Lyot coronagraph. Several physical optics phenomena associated with image formation contribute to modify the vector complex electromagnetic field incident onto the focal plane, and degrade image quality. These are caused by the interaction of light and matter: polarization induced by mirrors, windows & stops and diffraction produced by masks and stops and chromatic aberration which results from the wavelength-dependent indices of refraction of materials required to reflect (metal mirrors) and transmit (dielectrics, filters, prisms, etc.) light.

Figure 6.

Optical schematic for a typical Lyot coronagraph. The Star/planet complex electromagnetic field enters the Lyot coronagraph system from the left and is focused onto a complex occulting mask located at an image plane stop. This image plane stop is at the front focus of a collimator lens. The optical power on the collimator is such that an image of the entrance pupil is formed on the Lyot stop.

00030_psisdg9904_99041c_page_11_1.jpg

The three-dimensional electric field from exoplanets is thermal white-light broadband, either reflected from the planetary system’s parent star or emission from the planet or a mixture. The electric fields associated with the star and the planet are both spatially and spectrally incoherent45, 46, 47 at the source. This radiation travels through space, enters the telescope/coronagraph system and reflects from several mirrors to strike the stop at an image plane. The field is partially coherent at this stop. The stop must block by absorption or reflection almost all of the incident electric field from the star while passing as much as possible of the exoplanet field. Starlight diffracts around the entrance pupil and this electric field is scattered into the coronagraph to be blocked by the Lyot stop shown in Fig 6. The detector at the focal plane samples the modulus squared of the electromagnetic complex field, which appears as a speckle pattern cased by partial coherence of the wave-fields. This intensity speckle pattern is then digitized at an optimum dynamic range (number of bits) to obtain a high enough SNR for exoplanet characterization.

This intensity distribution contains information about the characteristics of the exoplanet as viewed through the “filter” of the telescope/coronagraph complex-vector transfer function (CVTF). This function, represented by the amplitude response matrix (ARM)48 varies significantly with the physical optics and the opto-mechanical design implementation. It is currently unknown for space coronagraph optical systems.

Maximum extinction

To minimize scattered light, the complex vector field at the image plane where the occulting mask is located (plane 3 in Fig. 6) must match the complex vector filter of the occulting mask. We use the formalism developed for optical processing of images49. From Eq. 12 we find the expression for the incoherent superposition of the four complex fields representing each of the 4 incoherent elements of the ARM to be:

00030_psisdg9904_99041c_page_12_1.jpg

Where 00030_psisdg9904_99041c_page_12_3.jpg is the Fourier transform operator, each of the four terms is complex and the superscript - sign on u refers to the complex electric field infinitesimally (ε) to the left (-) of the image plane. Note that it has been shown10 mathematically, confirmed observationally50 and relationships verified51 that the centroids of the terms are displaced slightly one from the other.

Let the transmittance of the image-plane occulting-mask be represented by the complex transmittance, τ(x3,y3), then the electric field infinitesimally (ε) to the right (+) of the image plane is given by

00030_psisdg9904_99041c_page_12_2.jpg

To achieve the contrast levels needed for terrestrial exoplanet spectrometry we need to minimize u+ (x3, y3), over the wavelength bands of interest while maximizing the transmittance at the position angle and separation of the exoplanet in orbit around the star.

To further control the unwanted radiation, this process needs to be repeated at the exit pupil to create the optimum complex Lyot stop. We need to satisfy the joint condition. This is a form of E&M impedance matching both at the image plane occulting mask and at he exit pupil Lyot stop under the conditions to maximize the field of the exoplanet at the entrance aperture to the spectrometer, to maximize the intensity at the focal plane of the spectrometer.

6.

PRELIMINARY POLARIZATION-ANALYSIS OF WFIRST-CGI

The WFIRST-CGI optical system52 contains two deformable mirrors to tune the geometric wavefront error phase53 in order to obtain a dark hole at the image plane. Mirrors 1 and 2 are shared with the WFIRST-WFI. Mirrors 3 and 4 fold the light to the Tertiary followed by 13 additional mirrors, two of them separated deformable mirrors before the coronagraph image plane mask, for a total of 18 mirrors between the exoplanet and the principal image plane mask. These 18 mirrors are tilted to the propagation axis and will partially polarize and retard the incident light. The exo-planet science path signal passes around this mask to three more mirrors before the pupil mask, and then follows 9 more tilted mirrors to the detector. The science path from the exoplanet to the focal plane appears to have a total of 33 reflections, at incidence angles between 15o and 20o.

We now extrapolate the polarization properties we find in our reference 3-mirror Cassegrain telescope system to the WFIRST-CGI. In WFIRST-CGI, our reference system includes the primary and secondary curved mirrors, but the number of reflections is increased from 1 in the 3-mirror reference system to over 30 and the incident angle changed from 45 degrees in the reference system to an average of about 15-20°. Since diattenuation and retardance are quadratic in the AOI, one mirror at 45° is approximately equal to 20 mirrors at 15-20°. Without a detailed modeling, we can expect the PSF for WFIRST-CGI to be similar to that shown in Fig. 4. This will give a contrast of the space-based WFIRST-CGI of approximately 10-6.

For54 33 bare aluminum mirrors at 15O AOI an unpolarized star becomes about 15% polarized with the weakest image in light polarized parallel to the plane of incidence. The requirement55 to observe in unpolarized light cannot be met with the system as designed using only geometric aberration considerations.

Polarization mitigation technologies are needed for AFTA-WFIRST-CGI to increase its threshold raw contrast and increase the system scientific yield to increase the success of the mission.

REFERENCES

1 

J. B. Breckinridge, T. G. Kuper and R. V. Shack, “Space telescope low scattered light camera: a model,” in Optical Engineering, 23, 816-820. also Proc. SPIE, 395 –403 (19841982). Google Scholar

2 

J. B. Breckinridge, “Polarization Properties of a Grating Spectrograph.,” Applied Optics, 10 286 –294 (1971). https://doi.org/10.1364/AO.10.000286 Google Scholar

3 

J. B. Breckinridge and B. Oppenheimer, “Polarization Effects in Reflecting Coronagraphs for White Light Applications in Astronomy,” ApJ, 600 1091 –1098 (2004). https://doi.org/10.1086/apj.2004.600.issue-2 Google Scholar

4 

Joseph Carson, James Breckinridge, John Trauger..., “The effects of instrumental elliptical polarization stellar point spread function fine structure,” in Proc. of the SPIE 6265-3M, (2006). https://doi.org/10.1117/12.672518 Google Scholar

5 

WFIRST-CGI science and technology definition team (STDT) 2nd meeting Spring 2011 GSFC, Google Scholar

6 

N. Zimmerman, A. J. E. Riggs, N. J. Kasdin,, “Shaped pupil coronagraphs: high contrast solutions for restricted focal planes,” JATIS 2, 011012, (2016). Google Scholar

7 

Hong Tang, M. Rud, R. Demers, R. Goullioud,, “The WFIRST/AFTA Coronagraph Instrument Optical Design,” in Proc. SPIE, 960504 (2015). Google Scholar

8 

K. Balasubramanian, S. Shaklan, A. Give’on, E. Cady and L. Marchen, “Deep UV to NIR space telescopes and exoplanet coronagraphs: a trade study on throughput, polarization, mirror coating options and requirements,” in Proc. SPIE 8151 51511G, (2011). Google Scholar

9 

Breckinridge, J., W.S.T. Lam and R. A. Chipman, “Polarization aberrations in astronomical telescopes,” 445 –468 (2015). Google Scholar

10 

R. A. Chipman, Wai Sze T. Lam and J. Breckinridge, “Polarization aberrations in astronomical telescopes,” in Proc SPIE 9613, (2015). Google Scholar

11 

Joseph W. Goodman, “Fourier Optics,” Roberts and Company book ISBN 9747077-2-4, (2005). Google Scholar

12 

J. D. Gaskill, “Linear Systems, Fourier Transforms and Optics,” John Wiley & Sons, NY 400 pages., (1978). Google Scholar

13 

Born, M. & E. Wolf, “Principles of optics,” Pergamon Press-Fresnel Equations are derived on, 627 –633 (1993). Google Scholar

14 

L. Ward, “Optical Constants of Bulk Materials and Films,” Adam Hilger, Bristol and Philadelphia, 245 page book., (1988). Google Scholar

15 

A. Thelen, “Design of Optical Interference Coatings,” McGraw-Hill, New York (1988). Google Scholar H. A. Macleod, “Thin-Film Optical Filters,” 185 –186 Institute of Physics,2002). Google Scholar J.A. Dobrowolski, “Coatings and Filters,,” Handbook of Optics, 8.1 –8.124 McGraw-Hill, New York (1978). Google Scholar

16 

R. M. A. Azzam and N. M. Bashara, “Ellipsometry and Polarized Light,” North Holland -Elsevier ISBN 0 4444 87016 4 ; 527 pages., (2003). Google Scholar

17 

Ward, L., “Optical Constants of Bulk Materials and Films,” Adam Hilger, p245, (1988). Google Scholar

18 

R. C. Jones, JOSA, 31, 488, (1941). Google Scholar

19 

20 

E. Wolf, “Theory of Coherence and Polarization of Light,” Chapter 8, 225 Cambridge University Press,2007). Google Scholar J. W. Goodman, “Statistical Optics,” Wiley,2015). Google Scholar J. D. Strong, “Classical Optics,” Freeman, 178 ,1958). Google Scholar

21 

E. Wolf, “Theory of Coherence and Polarization of Light,” Chapter 8, 225 Cambridge University Press,2007). Google Scholar J. W. Goodman, Statistical Optics, Wiley,2015). Google Scholar J. D. Strong, Classical Optics, Freeman, 178 ,1958). Google Scholar

22 

J.P. McGuire, Jr., R.A. Chipman, “Diffraction image formation in optical systems with polarization aberrations I: Formulation and example,” Journal of the Optical Society of America A., 7 (9), 1614 –1626 (19901990). https://doi.org/10.1364/JOSAA.7.001614 Google Scholar

23 

Bruegge, T. J., in Proc. SPIE, (1989). Google Scholar

24 

Chipman, R. A., “Polarization analysis of optical systems,” in Proc. SPIE, (1989a). Google Scholar

25 

Wolff, L. B. and D. J. Kurlander, “1990, IEEE Computer Graphics and applications,” 10 (6), 44 –55 Google Scholar

26 

J.D. Trolinger Jr., R.A. Chipman, D.K. Wilson, “Polarization ray tracing in birefringent media,” Optical Engineering, 30 (4), 461 –466 (19911991). https://doi.org/10.1117/12.55815 Google Scholar

27 

Yun, G., Crabtree, K. & Chipman, R., Appl. Opt. 50, 2855 & 2011b, Appl. Opt. 50, 2866, (2011a). Google Scholar

28 

S.C. McClain, L.W. Hillman, R.A. Chipman, “Polarization ray tracing in anisotropic optically active media I, algorithms, & II, theory and physics,” Applied Optics, 10 (11), 2371 –2393 (1993). Google Scholar

29 

R. A. Chipman and W. S. T. Lam, “The Polaris-M ray tracing program,” in Proc. SPIE, 9613 –18 (2015). Google Scholar

30 

Kuboda, H. & Inoue, S., JOSA, 49, 191, (1959). Google Scholar

31 

Urbanczyk, W., Opt. Acta., 33 (53), (1986). Google Scholar

32 

McGuire, J. P. & Chipman, R. A., JOSA A, 7, 1614 & 1991, JOSA A, 8, 6, (1990). Google Scholar

33 

Tu, Y., Wang, X., Li, S. & Cao, Y., Opt. Lett., 37 (2061), (2012). Google Scholar

34 

R.A. Chipman, L.J. Chipman, “Polarization aberration diagrams, Optical Engineering,” 28 (2), 100 –106 (1989). Google Scholar

35 

J.P. McGuire, Jr., R.A. Chipman, “Polarization aberrations I: Rotationally symmetric optical systems and Polarization aberrations II: Tilted and decentered optical systems,” Applied Optics, 33 (2), 5080 –5107 (1994). https://doi.org/10.1364/AO.33.005080 Google Scholar

36 

D.J. Reiley, R.A. Chipman, “Coating-induced wave-front aberrations: On-axis astigmatism and chromatic aberration in all-reflecting systems,” 33 (10), 2002 –2012 (1994). Google Scholar

37 

I. C. Gardner, “Application of the Algebraic Aberration Equations to Optical Design,” NBS Scientific Papers, 22 (550), (1927). Google Scholar

38 

G. Yun, K. Crabtree, and R. Chipman, “Three-dimensional polarization ray-tracing calculus I: definition and diattenuation”, Appl. Opt. 50, 2855-2865 (2011). AND G. Yun, S. McClain, and R. Chipman, “Three-dimensional polarization ray-tracing calculus II: retardance,” Appl. Opt, 50 2866 –2874 (2011). https://doi.org/10.1364/AO.50.002866 Google Scholar

39 

Figure 7: Breckinridge, J., W.S.T. Lam and R. A. Chipman, “Polarization aberrations in astronomical telescopes,” Publ. Astron Soc. Of the Pacific, 127 445 –468 (2015). https://doi.org/10.1086/681280 Google Scholar

40 

Figure 7: Breckinridge, J., W.S.T. Lam and R. A. Chipman, “Polarization aberrations in astronomical telescopes,” Publ. Astron Soc. Of the Pacific, 127 445 –468 (2015). https://doi.org/10.1086/681280 Google Scholar

41 

D.J. Reiley, R.A. Chipman, “Coating-induced wave-front aberrations: On-axis astigmatism and chromatic aberration in all-reflecting systems,” Applied Optics, 33 (10), 2002 –2012 (1994). https://doi.org/10.1364/AO.33.002002 Google Scholar

42 

J. Breckinridge, Wai Sze T. Lam and R. A. Chipman, “Polarization Aberrations in Astronomical Telescopes: The Point Spread Function,” Publications of the Astronomical Society of the Pacific (PASP), 127 445 –468 (2015). https://doi.org/10.1086/681280 Google Scholar

43 

R. A. Chipman, Wai Sze T. Lam and J. Breckinridge, “Polarization aberrations in astronomical telescopes,” in Proc SPIE 9613, (2015). Google Scholar

44 

B. Safonov, On-sky demonstration of optical polarostrometry, MNRAS, 451 3161 –3172 (2015). Google Scholar

45 

M. Born and E. Wolf, “Principles of Optics,” Chapter 10: Partially coherent light, Cambridge University Press, (1999). Google Scholar

46 

L. Mandel & E. Wolf, “Coherence Properties of Optical Fields,” Rev. Mod. Phys, 37, 231 –280 (1965). https://doi.org/10.1103/RevModPhys.37.231 Google Scholar

47 

J. Goodman, “Statistical Optics,” Ch 5: Coherence of Optical Waves, (2015). Google Scholar

48 

Breckinridge, Lam and Chipman, “Polarization Aberrations in Astronomical Telescopes: The Point Spread Function,” Publications of the Astronomical Society of the Pacific (PASP), 127:445468. Equation 8, (2015). Google Scholar

49 

Joseph Goodman, “Fourier Optics,” Chapter on optical matched filters, Roberts and Company book ISBN 9747077-2-4, (2005). Google Scholar

50 

B. Safanov, On-sky demonstration of optical polaroastrometry, NMRAS, 451, 3161 –3172 (2015). Google Scholar

51 

B. Safanov, private communication, (2016). Google Scholar

52 

T. Hong, M. Rud, R. Demers, R. Goullioud, J. Krist, F. Shi and F. Zhao, The WFIRST/AFTA coronagraph instrument optical design SPIE vol, 9605 (2015). Google Scholar

53 

S. Shaklan & K. Balasubramanian,, in Proc. SPIE, (2011). Google Scholar

54 

Breckinridge, W.S.T. Lam and R. A. Chipman, Polarization aberrations in astronomical telescopesPubl. Astron Soc. Of the Pacific, 127 445 –468 (2015). Google Scholar

55 

WFIRST-AFTA 2015 SDT report Table F-2 lists the requirement to observe in unpolarized light in filter bands 3,6, & 7, Google Scholar
© (2016) COPYRIGHT Society of Photo-Optical Instrumentation Engineers (SPIE). Downloading of the abstract is permitted for personal use only.
J. B. Breckinridge and R. A. Chipman "Telescope polarization and image quality: Lyot coronagraph performance", Proc. SPIE 9904, Space Telescopes and Instrumentation 2016: Optical, Infrared, and Millimeter Wave, 99041C (29 July 2016); https://doi.org/10.1117/12.2231242
Lens.org Logo
CITATIONS
Cited by 6 scholarly publications.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Polarization

Mirrors

Point spread functions

Dielectric polarization

Telescopes

Exoplanets

Coronagraphy

Back to Top