arXiv:1105.2615v1 [astro-ph.IM] 13 May 2011 bainuista r eindt ii h nominal the mimic to cal- designed optical are polarized of that use units and ibration stars, standard po- and larized unpolarized observing include Cali- techniques bration techniques. rapid differential of- emphasize common-path and telescope and optics chopping and non-moving of instrument use minimize, involve the ten to calibrate Strategies of- and are noise. photon that stabilize the errors than larger ten induce all transparency sensitiv- can and variations optical non-uniform seeing, atmospheric as Moving nonlinearity, or such ity effects observation. al. po- detector an et the elements, of 2008; Harrington enhances precision and 2007, errors larimetric systematic these 2003, Kuhn minimizes & Harrington 2009a,b). al. et Kuhn 2010, Infra- Hodapp & al. is Harringtion et and 2009, Thornton polarization Visible 2006, (cf. Resolution before (HiVIS) are 2000, High There Spectropolarimeter reflections our red oblique . at polarized ex- polarized transformed analyzed highly linearly Our be circularly circumstances, 100% that 7 2008; some nearly System is in 2010. to even, Electro-Optical telescope can al. Advanced light al. et 2005; 3.67m et Ichimoto the Elmore (AEOS) Socas-Navarro with 2009; Selbing 2003; 2006; perience 2005; Snik & tech- al. Keller et 1992; al. Giro al. of et et Socas-Navarro Socas-Navarro al. 1993; repretoire et (Elmore 2005a,b; a al. telescopes coud´e, et developed calibrating a Kuhn of for have astronomers analyzer niques Solar particular the polarization reaches polarimeter. in input it Gregorian or the before Nasmyth, light multiple scramble the with to of common telescopes reflections is altitude- mirror it modern Unfortunately, telescope-detector for for calibration. techniques polarization efficient by enhanced EIIGTLSOEMELRMTIE SN ATM K POLA DAYTIME USING MATRICES MUELLER TELESCOPE DERIVING [email protected] [email protected] [email protected] rpittpstuigL using 2018 typeset 14, Preprint November of Version ffcieclbaino h eecp n instrument and telescope the of calibration Effective be may which tool powerful a is Spectropolarimetry ehdfrrcvrn h nu oaiainsaeo h ih n t and light the of matrix state Mueller polarization system input optical fro the different the recovering signficantly by for be defined method can is signal “cross-talk”, optical ization the of state output fiain bevtoso h rgt ihyplrzddyiesyus sky tec daytime the headings: external polarized demonstrate Subject without coud´e highly spectropolarimeter percent a bright, few and the a telescope of of Observations accuracy an ification. with telescope the eecpsotnmdf h nu oaiaino ores htt that so source a of polarization input the modify often Telescopes 2 1. nttt o srnm ai nvriyo aai 4Ohia 34 Hawaii, of University Maui, Astronomy for Institute 1 nttt o srnm,Uiest fHwi,28 Woodla 2680 Hawaii, of University Astronomy, for Institute A 3 INTRODUCTION T eateto srnm,WimnClee 4 oe v. W Ave., Boyer 345 College, Whitman Astronomy, of Department E tl mltajv 11/10/09 v. emulateapj style X srnmclTcnqe,Atooia ntuetto,Data Instrumentation, Astronomical Techniques, Astronomical niques ai .Harrington M. David eso fNvme 4 2018 14, November of Version 1 ABSTRACT .R Kuhn R. J. , lsl olw olt19 n lre20.I the de- In is light 2009. of 4-vector: state Clarke a as polarization and noted the 1992 formalism, Stokes Collet follows closely on telescope 3.7m AEOS technique Force the Air Maui. illustrates Haleakala the paper and using This zenith different the at observing positions. Sun by azimuth the realized with is sky This illuminated percent. solar ac- few an the with a recovered of pointing, be curacy can telescope polarization matrix each linear Mueller at multiple telescope full detected as are long states As observing input modifications telescope. dark calibration the hardware precious of no using requires source. and without calibration time observable characterized is well polarized, relatively It highly a observable, day- is easily The and normally linearly . is Rayleigh the sky solar of time to percent. observations due more few sky or telescope- a polarized two of many only accuracy requires of an It calibration with systems matrix polarimeter Mueller full the obser- target during optics telescope vations. the of illumination oiinage 0 angles position and oe as noted oe aeadaedfie as: defined are and case lower South-West. and while North-East North-South aligned be + to direction vibration field aie nest.Nt htacrigt hsdefinition, 90 this angles to along according polarization that linear Note intensity. larized and beam, light the infratooia oaier sfrte+ the for is polarimetry astronomical for tion U h olwn icsino oaiainformalism polarization of discussion following The yield to designed is here presented technique new The nti formalism, this In h omlzdSoe aaeesaedntdwith denoted are parameters Stokes normalized The 2 U a h lcrcfil irto ieto lge to aligned direction vibration field electric the has n hno Hall Shannon and , h ieryplrzditniyaogpolarization along intensity polarized linearly the hnique. − nDie oouu I 96822 HI, Honolulu, Drive, wn uS. uaai I 66 and 96768 HI, Pukalani, St., Ku Q h nu.Ti iig rpolar- or mixing, This input. the m ak rtlsoehrwr mod- hardware telescope or masks and laWla A 99362 WA, Walla, alla edsrb eea efficient 4 an full here he describe We . ◦ n h aekl .mAEOS 3.7m Haleakala the ing emaue iclro linear or circular measured he − n 45 and S 1.1. U V i I epciey h yia conven- typical The respectively. , [ = stergthne iclrypo- circularly right-handed the is ersnstettlintensity, total the represents Polarization IAINOBSERVATIONSRIZATION ◦ ,Q ,V U, Q, I, × 3 ntepaepredclrto perpendicular plane the in ule arxof matrix Mueller 4 nlssadTech- and Analysis ◦ ] T n 135 and ◦ ilb de- be will Q electric (1) Q 2 Harrington et al.

is also available using industry standard atmospheric ra- [1,q,u,v]T = [I,Q,U,V ]T /I (2) diative transfer software packages such as MODTRAN (cf. Fetrow et al. 2002, Berk et al. 2006). The degree of polarization can be defined as a ratio of polarized light to the total intensity of the beam:

Q2 + U 2 + V 2 P = = q2 + u2 + v2 (3) p I For details on polarization ofp light and stellar spec- tropolarimetry, see Collet 1992 and Clarke 2009. To describe how polarized light propagates through any optical system, the Mueller matrix is constructed which specifies how the incident polarization state is transferred to the output polarization state. The Mueller matrix is a 4 × 4 set of transfer coefficients which when S multiplied by the input Stokes vector ( iinput ) gives the S output Stokes vector ( ioutput ): S M S ioutput = ij iinput (4) If the Mueller matrix for a system is known, then one inverts the matrix and deprojects a set of measurements to recover the inputs. One can represent the individ- ual Mueller matrix terms as describing how one incident state transfers to another. In this work we will use the notation: Fig. 1.— The celestial triangle representing the geometry for the sky polarization at any telescope pointing. γ is the angu- II QI UI VI lar distance between the telescope pointing and the sun. θs is the solar angular distance from the zenith. θ is the angular M IQQQUQVQ ij =  IUQUUUVU  (5) distance between the telescope pointing and the zenith. φ is IV QV UV VV the angle between the zenith direction and the solar direction   at the telescope pointing. ψ is the angle between the tele-   1.2. Deriving Telescope Mueller Matrix Elements scope pointing and the solar direction at the zenith. The law of Cosines is used to solve for any angles needed to compute Typical calibration schemes on smaller telescopes of- degree of polarization and position angle of polarization. ten use fixed polarizing filters placed over the telescope aperture to provide known input states that are de- tected and yield terms of the Mueller matrix. This ap- There are many atmospheric and geometric con- proach has been used for solar telescopes (Socas-Navarro siderations that determine the skylight polarization 2005a,b, Socas-Navarro et al. 2005, 2006). Another ap- at a particular observatory site. The linear po- proach, also used in the solar observations has been larization amplitude angle can depend on the so- to image known sources and use spectropolarimet- lar elevation, atmospheric aerosol content, aerosol ric data and Zeeman effect symmetry properties to vertical distribution, aerosol scattering phase func- isolate and determine terms in the Mueller matrix tion, wavelength of the observation and secondary (Kuhn et al. 1993, Elmore et al. 2010). Night-time sources of illumination (cf. Horv´ath et al. 2002a,b, observations can use unpolarized and polarized stan- Lee 1998, Liu & Voss 1997, Suhai & Horv´ath 2004, dard stars to measure polarization properties of tele- G´al et al. 2001, Vermeulen et al. 2000, Pomozi et al. scopes (cf. Hsu & Breger 1982, Schmidt et al. 1992, 2001, Cronin et al. 2005, 2006, Heged¨us et al. 2007). Gil-Hutton & Benavidez 2003, Fossati et al. 2007). Anisotropic scattered sunlight from reflections off land Many studies have either measured and calibrated tele- or water can be highly polarized and temporally scopes, measured mirror properties or attempted to variable (Litvinov et al. 2010, Peltoniemi et al. 2009, design instruments with minimal polarimetric defects He et al. 2010, Salinas & Liew 2007, Ota et al. 2010, (cf. S´anchez Almeida et al. 1991, Giro et al. 2003, Kisselev & Bulgarelli 2004). Aerosol particle op- Patat & Romaniello 2006,Tinbergen 2007, Joos et al. tical properties and vertical distributions also vary 2008, van Harten et al. 2009, Roelfsema et al. 2010). (cf. Wu & Jin 1997, Shukurov & Shukurov 2006, Vermeulen et al. 2000, Ougolnikov & Maslov 2002, 1.3. The Polarized Sky as a Calibration 2005a,b, 2009a,b, Ugolnikov et al. 2004). The polar- The observed polarization of the daytime sky is a ization can change across atmospheric absorption bands useful calibration source. Scattered sunlight is bright, or can be influenced by other scattering mechanisms highly (linearly) polarized and typically illuminates (cf. Boesche et al. 2006, Zeng et al. 2008 Aben et al. the telescope optics more realistically than calibration 1999, 2001). screens. A single-scattering atomic Rayleigh calcula- Deviations from a single scattering Rayleigh model tion (cf. Coulson 1980, 1998) is often adequate to de- grow as the aerosol, cloud, ground or sea-surface scat- scribe the skylight polarization. More realistic model- tering sources affect the telescope line-of-sight. Clear, ing, involving multiple scattering and aerosol scattering, cloudless, low-aerosol conditions should yield high lin- Mueller Matrices Using Sky Polarization 3 ear polarization amplitudes and small deviations in the polarization direction from a Rayleigh model. Obser- vations generally support this conclusion (Pust & Shaw 2005, 2006a,b, 2007, 2008, 2009, Shaw et al. 2010). The geometry of our Rayleigh sky model is seen in Fig- ure 1. The geometrical inputs are the observers location (latitude, longitude, elevation) and local time. The solar location and relevant angles from the telescope point- ing are computed from the spherical geometry in Fig- ure 1. The maximum degree of polarization (δmax) in this model occurs at a scattering angle (γ) of 90◦. The Rayleigh sky model predicts the degree of polarization (δ) at any telescope pointing as:

δ sin2γ δ = max (6) 1+ cos2γ Since the angle of polarization in the Rayleigh sky model is always perpendicular to the scattering plane, one can derive the angle of polarization with respect to Fig. 2.— This Figure shows the sky polarization modeled for the altitude and azimuth axes of the telescope. The law th of Cosines for the scattering plane breaks down at the Haleakala at 3:00 UT on January 27 2010. The degree of polar- ization (δ) at all telescope pointings with δmax scaled to 100% is Zenith where θ is 0◦ and one must simply calculate the in the top left panel. The scattering angle with respect to the sun is in the top right panel. The solar altitude is 15◦ and the solar az- difference in azimuth between the telescope pointing and ◦ the solar azimuth to find the orientation of the polariza- imuth is 243 . The calculated q and u inputs as seen by the AEOS altitude-azimuth telescope are on the bottom two panels. +Q is tion. An example of the model we compared with our defined as E-field vibrations along the +Altitude direction. +U is observations is shown in Figure 2 for Haleakala on Jan- defined as E-field vibrations along the +Altitude, +Azimuth di- uary 27th 2010. rection. All panels show stereographic projections to the observed One set of all-sky polarization measurements obtained sky with the Zenith in the center of the circle. North, East, South, West are on the perimeter of the circle with North up and West at the Mauna Loa Observatory, a nearby site and at right. very similar altitude to Haleakala (Dahlberg et al. 2009) are particularly relevant. The maximum degree of po- In general only two measurements with different in- larization (δmax) changed throughout the day in their observations from a value of ∼60% with the sun at put linear polarization orientations are sufficient to de- termine the telescope polarization properties. In prac- an elevation of 30◦, to 85% with the sun setting. Al- though this is a significant change in the maximum tice we define a least-squares problem that takes several input polarization measurements of the sky at different degree of polarization (δmax) presumably due to mul- tiple scattering, the polarization direction at all tele- times but with identical optical configurations to derive scope pointings is expected to be close to the Rayleigh telescope properties. model (cf. Dahlberg et al. 2009, Suhai & Horv´ath 2. POLARIZED SKY OBSERVATIONS 2004, Pust & Shaw 2009, Pomozi et al. 2001). In our analysis below we will use the Rayleigh model to define To demonstrate this technique, we collected sky obser- the input polarization angle. We believe our assumed vations with AEOS using the new low spectral resolution mode for our coud´espectrograph we call LoVIS. This input angle is accurate over most of the sky to <5◦ (eg. Suhai & Horv´ath 2004). new spectrograph has been characterized to verify the expected performance using standard stars, calibration optics and various tests outlined in the Appendix. The 1.4. Using the Sky for Full-Stokes Calibration LoVIS system polarization properties most relevant to Despite the fact that the Rayleigh sky only provides this work include the telescope and instrument induced linearly polarized light to the telescope, with reasonable polarization, depolarization, the polarimetric response of assumptions, we can recover the full Mueller matrix. We LoVIS. These are briefly described below and in the ap- find that the telescope and spectrograph only weakly po- pendix. larize unpolarized input light, and the optical system has The measured degree of polarization for the many Lo- only a weak depolarization effect on polarized input light. VIS sky observations we have obtained shows that there Thus, the first row and column of the system Mueller ma- is relatively low instrument depolarization. For exam- trix can be approximated by the corresponding row and ple, on January 27th 2010 UT we have twenty sky po- column of the identity matrix. Direct measurements of larization observations with the telescope pointed at the these terms using stellar observations with AEOS con- Zenith taken over three hours ending just after sunset. firm that over the visible spectrum these Mueller matrix The measured degree of polarization averaged over all terms are less than 5%. With this assumption the tele- wavelengths is shown in Figure 3 as the solid line. Fol- scope system scrambles the input polarization simply by lowing Dahlberg et al. 2009 Figure 7, we assume the rotating the input 4 element Stokes vector in the 3-space maximum degree of polarization (δmax) for the Rayleigh defined along the Q, U, and V polarization axes to yield sky as roughly 65% when the solar elevation is 30◦, 75% the output measured state. at 15◦ and 85% at 0◦. If we apply this simple linear rela- 4 Harrington et al.

Fig. 3.— This Figure shows the average sky polarization with Fig. 4.— This Figure shows the continuum polarization variation the telescope pointed at the Zenith measured by LoVIS on January from average as a function of wavelength for all twenty LoVIS 27th 2010 UT. The statistical error in the measurement is much Zenith measurements on Jan. 27th 2010. The average value of smaller than the width of the line. The solid line is the average the polarization, shown in Figure 3 has been removed from all polarization measured for all wavelengths. The dashed line is a observations to highlight the change with wavelength. The solid Rayleigh Sky model prediction for the polarization at the Zenith lines show observations taken 2-3 hours before sunset. The dot- following measurements from Mauna Loa in Dahlberg et al. 2009 dash lines show observations taken 0.2-1 hour before sunset. The Figure 7. The maximum degree of sky polarization in the 90◦ dashed line shows a single observation taken just after sunset. scattering plane (δmax) is assumed to rise linearly from 65% to 85% as the sun sets from elevation 30◦ to 0◦. The scattering angle (γ) between the sun and the Zenith telescope pointing is shown in R the right hand y-axis. is shortened to cγ. We specify the rotation matrix ( ij ) using the ZXZ convention as: tion for δmax and compute the Rayleigh sky polarization (δ) at the Zenith for our observations, we get the dashed cγ sγ 0 1 00 cα sα 0 Rij = −sγ cγ 0 0 cβ sβ −sα cα 0 curve in Figure 3. The Figure also shows the scattering − angle (γ) between the Zenith telescope pointing and the 0 01 ! 0 sβ cβ ! 0 01 ! sun. (7) The computed degree of polarization changes with With this definition for the rotation matrix, we solve wavelength for all twenty Jan 27th observations is shown for the Euler angles assuming a fully linearly polarized in Figure 4. The average polarization shown in Figure daytime sky as calibration input. If we denote the mea- S 3 was removed from the spectra leaving only residual sured Stokes parameters, i, as(qm,um, vm) and the in- put sky Stokes parameters, Rj , as (qr,ur, 0) then the chromatic changes. There is a mild trend with wave- × length for the measured polarization to decrease by about 3 3 QUV Mueller matrix elements at each wavelength 10% from 5500A˚ to 7500A.˚ The atmospheric oxygen are: absorption band shows a mild increase in polarization. There is also a substantial change seen the chromatic qm QQUQVQ qr trend seen in the final observation taken just after sun- Si = um = Mij Rj = QUUUVU ur set as shown by the dashed line of Figure 4 consistent vm ! QV UV VV ! 0 ! with studies of the occultation and changing illumina- (8) tion (cf. Ougolnikov & Maslov 2002, 2005a,b, 2009a,b, This set of equations has 6 variables and only 5 known Ugolnikov et al. 2004). quantities. We have no V input to constrain the VQ, VU Given that AEOS telescope and LoVIS spectropo- and V V terms. Nevertheless, two input and output vec- larimeter induces polarization of less than 3% and the tors on the Poincare sphere are sufficient to fully specify depolarization is of the same order, we can use a 3 × 3 the full 3 × 3 rotation matrix. Thus, we use the fact that Mueller matrix approximation. In this 3×3 QUV space, the sky polarization changes orientation with time and the degree of polarization is preserved for any input state take measurements at identical telescope pointings sep- and the 3x3 Mueller matrix would be approximated as a arated by enough time for the solar sky illumination to rotation matrix inside the Poincar´esphere. change. This yields an over-constrained solvable prob- lem for all six linear polarization terms in the Mueller 3. TELESCOPE POLARIZATION AS ROTATION matrix. To model the the 3 × 3 Mueller matrix as a rotation When using this rotation matrix approximation for the matrix, we will use Euler angles. We scaled all our mea- telescope Mueller matrix, the Rayleigh Sky input Stokes sured Stokes vectors to unit length in order to remove parameters multiply each term of the rotation matrix to the residual effects from changes in the sky degree of give a system of equations for the three Euler anlges (α, polarization, telescope induced polarization and depo- β, γ). This system of equations can be solved using a larization; and to put our measurements on the Poincare normal non-linear least-squares minimization by search- spher´e. We denote the 3 Euler angles as (α,β,γ) and use ing the (α, β, γ) space for minima in squared error. This a short-hand notation for sines and cosines where cos(γ) direct solution of this set of equations using standard Mueller Matrices Using Sky Polarization 5

Fig. 5.— This Figure shows the six Mueller matrix elements Fig. 6.— This Figure shows the variation in Euler angles with versus wavelength at a telescope pointing of 90◦ altitude, 225◦ az- wavelength at a telescope pointing of 90◦ altitude, 225◦ azimuth. imuth. These were derived from three sky measurements on Jan- The value for each Euler angle at 6560A˚ have been removed for clar- uary 27th 2010 using the two-step solution. The UV term grows ity [α=176.5◦, β=81.6◦, γ=98.8◦]. The change with wavelength is with increasing wavelength while the QU term decreases with wave- dominated by β which corresponds to the changes in the QU and length. The other Mueller matrix elements remain near 0 or 1 and UV Mueller matrix elements from Figure 5. The statistical noise show smaller chromatic variation. is smaller than the width of the line. minimization routines is subject to several ambiguities different days. For instance, we have two sets of sky ob- that affect convergence using standard minimization rou- th th tines. For example, Euler angles are symmetric under servations taken December 10 and 11 2009 each con- sisting of thee observation sets spaced roughly one hour the exchange of (α, β, γ) with ( -180◦+α,-β, -180◦+γ). Solutions are also identical with multiples of 2π. Asa apart. The Euler angles derived from on different days convenient approach to this least-squares problem, we agree to within 0.5◦ giving an estimate of the system- used a different set of equations in a two-step solution. atic error limits. The Mueller matrix elements derived We first perform simple least squares solution directly with observations from different days agree to better than for the six linear-polarization Mueller matrix elements. 0.01. The changes in β with telescope azimuth shown in In the second step we fit a rotation matrix to the six Figure 7 are well above the systematic noise of roughly Mueller matrix elements at each wavelength. As an ex- 0.5◦ and are repeatable even when using many different ample, Figure 5 shows the Mueller matrix elements de- combinations of calibration data on both days. Similar rived with the Jan. 27th 2010 observations. The details trends for Euler angle variation with telescope pointing of our methods for deriving Euler angles and an example are seen at other wavelengths. of how one could plan sky calibration observations are outlined in the Appendix.

3.1. Euler Angle Solution Properties The Euler angles are well behaved smooth functions of both wavelength and telescope pointing. For instance, the Euler angle variations with wavelength calculated from January 27th 2010 observations at a telescope point- ing of at 90◦ altitude and 225◦ azimuth are shown in Figure 6. The change with wavelength is dominated by β while smaller variation is seen in α and γ. For clarity in the Figure, the Euler angle values at 6560A˚ have been removed so all curves overlap. Similar curves are seen for other telescope pointings. The Euler angles are also smooth functions of telescope pointing. We have observations on multiple days taken on a grid of 8 and 3 different altitudes. Figure 7 shows the derived Euler angles at 6560A˚ as a functions of Fig. 7.— This Figure shows the variation in derived Euler angles azimuth. There are three different curves corresponding at 6560A˚ as a function of telescope pointing. The top panel shows α for all telescope azimuths at elevations of 90◦ (solid), 75◦ (dashed) to the three different telescope altitudes. We find that γ and 55◦ (dot-dashed). The middle panel shows β and the bottom effectively absorbs the changing azimuth while α strongly panel shows γ for all azimuths at the same elevations as α using varies with altitude. Since there is substantial telescope the same line scheme. This Figure clearly shows that α changes induced rotation, these curves show significant deviations the strongest with elevation while γ shows changes with azimuth. from a purely geometrical rotation. There is not much change seen in β. The derived Euler angles in Figure 6 are quite repeat- able when derived from calibration observations taken on 6 Harrington et al.

3.2. Calibrating Spectropolarimetric Observations Once telescope calibrations have been derived, observa- tions subsequently taken with the instrument at the same telescope pointing can be de-rotated. The calibrated po- larization measurements are then oriented in a reference frame on the sky corrected for all geometric, telescope and instrument induced rotation. We have done many experiments where sky observations on some days are used to calibrate sky observations taken on other days to examine the stability and repeatability of the calibra- tions. As an example, Figure 8 shows six individual sky spec- tropolarimetric measurements taken on December 10th and 11th 2009 at different times of the afternoon. All observations were taken at a telescope pointing of az- imuth 021◦ and altitude 70◦. The input linear polariza- tion orientation changes with time and there is significant Fig. 9.— This Figure shows the difference between the de-rotated chromatism and cross-talk observed in all Stokes param- measured Stokes parameters and the predicted sky input Stokes parameters at altitude 70◦ and azimuth 021◦. The residual errors eters. The three observations from each day are taken at are less than 0.01 showing very low error between measured and similar times and the observed polarization spectra are expected Stokes parameters. similar. We get consistent results for Euler angles using many different combinations of observations taken between Au- gust 2009 and March 2010. Over this time period there were changes to the optical configuration such as re- mounting of the retarders, switching gratings and chang- ing LoVIS / HiVIS modes as would be expected during normal observing operations. If one derives a residual ro- tation angle between the de-rotated measurements and the predicted sky, the residual rotation angles are al- most always less than 3◦ when using any calibration set against any other observation set. If one uses calibration data where there are no optical changes between calibra- tion and observation, this resulting angular error is much smaller, typically less than 0.5◦ rotation. This allows us to estimate systematic error from optical configuration changes as a few degrees residual rotation. When observ- Fig. 8.— This Figure shows the measured spectropolarimetry ing as calibrated, the errors are under 1◦ QUV rotation. with LoVIS while the telescope is fixed at altitude 70◦ and azimuth ◦ 021 scaled to 100% polarization. Three observations were taken 4. DISCUSSION December 10th 2009 (solid lines) and another three observations were taken at similar times on December 11th 2009 (dashed lines). We have demonstrated a simple method for deriving The statistical noise is smaller than the width of the line. telescope polarization properties and Mueller matrix el- ements. By using observations of the bright, highly po- larized daytime sky and taking observations at multiple In order to illustrate a general method of telescope times using identical optical configurations one can ex- calibration, imagine a scenario where we derive a set of tract the polarization response of the telescope. calibrations using observations from one day and then This method can be easily implemented via least use this telescope calibration to de-rotate observations squares methods. A direct least-squares solution for taken on another day at an identical telescope pointing. Euler angles and a two-step solution for the linear- We will treat the December 10th 2009 observations as a polarization sensitive Mueller matrix elements and a ro- calibration set and derive Euler anlges for all telescope tation matrix fit to these elements have been imple- pointings and wavelengths. With these Euler angles in mented. Examples of how to optimally prepare calibra- hand, we will then treat observations taken on December tion observations and solve for the telescope response 11th 2009 as our science targets. The de-rotated science properties are detailed in the Appendix. observations will be compared against the Rayleigh sky The AEOS telescope and LoVIS spectropolarimeter prediction. This scenario is exactly what a typical ob- have over 20 reflections, many at high incidence angles server would perform during routine operations. The before the polarization analyzer. However, the major po- difference between the de-rotated December 11th obser- larization cross-talk effect is a simple rotations in QUV vations and the expected sky input Stokes parameters space that is readily derived to an accuracy of a few per- are shown in Figure 9. The residual error is typically cent from sky observations. Observations taken on mul- less than 0.01 for any Stokes parameter showing that the tiple days show we can de-rotate LoVIS sky observations de-rotated observations match the expected sky model to the predicted Rayleigh sky with at least 0.5◦ accuracy. with very high accuracy. The polarized sky is a calibration source that is bright, Mueller Matrices Using Sky Polarization 7 easily observable, highly polarized, simple to model, well characterized and easy to verify with additional instru- mentation. Unpolarized and polarized standard stars have several disadvantages as calibration sources. Stars are not available at all telescope pointings and require us- ing night-time observing time for calibration. Telescopes with complex optical trains such as AEOS require cali- bration at many pointings making the unavailability of stars at all pointings a serious problem. Solar instru- ments are typically unable to observe stars at all but can easily utilize this sky-based technique. Polarized standard stars normally have low degrees of polarization which adds time or noise constraints to such calibration efforts. This technique uses sky-light to illuminate the telescope and instrument optics like the target observa- tions. Night-time observations benefit from the fact that day-time observations are used to calibrate, rather than using precious night telescope time for calibration. 8 Harrington et al.

5. APPENDIX: SOLUTIONS FOR EULER ANGLES fit a rotation matrix, we avoid over-polarizing. There are a number of methods available to solve for Provided the arrays are indexed properly, the normal Euler Angles and Mueller matrix elements given a set solution for Mueller matrix elements can be computed via the normal least-squares method. We re-arrange the of sky polarization observations. Suppose we have mea- R surements at different times but at identical telescope time-varying Rayleigh sky inputs to ( ij ) for i indepen- dent observations and j input Stokes parameters. The pointings. Multiplying out terms in equation 7 for the S ZXZ-convention we get: measured Stokes parameters ( i) become individual col- umn vectors. The unknown Mueller matrix elements are also arranged as a column vector by output Stokes pa- − cαcγ sαcβsγ sαcγ + cαcβsγ sβsγ rameter (Mj ). If we write measured Stokes parameters R = −cαsγ − sαcβcγ −sαsγ + cαcβcγ sβcγ (9) ij as (qmi ,umi , vmi ) and the Rayleigh input Stokes parame- sαsβ −cαsβ cβ ! ters as (qri ,uri ), we can explicitly write a set of equations for just two Mueller matrix elements: Then equating Mueller matrix elements to rotation matrix elements could write the system of equations for q q u the three Euler angles as: m1 r1 r1 QQ S q R M q u i = m2 = ij j = r2 r2 UQ (12) q 3 ! q 3 u 3 ! qm1 qr1(+cαcγ c sαsγ ) + ur1(+cγ sα + cαc sγ ) m r r   − β β um1 qr1( c cγ sα cαsγ ) + ur1(+cαc cγ sαsγ ) − β − β − vm1 qr1(+sαs ) + ur1( cαs ) We have three such equations for each set of Mueller β − β  qm2   qr2(+cαcγ c sαsγ ) + ur2(+cγ sα + cαc sγ )  − β β matrix elements. With this can express the residual er- um2 = qr2( cβcγ sα cαsγ ) + ur2(+cαcβ cγ sαsγ ) − − − ror (ǫi) for each incident Stokes parameter (Si) with an  vm2   qr2(+sαs ) + ur2( cαs )  β − β  qm3   qr3(+cαcγ c sαsγ ) + ur3(+cγ sα + cαc sγ )  implied sum over j as:    − β β  um3 qr3( c cγ sα cαsγ ) + ur3(+cαc cγ sαsγ )    − β − β −   vm3   qr3(+sαsβ ) + ur3( cαsβ )     −  ǫi = Si − Rij Mj (13)    (10) This system of equations can be solved using a nor- The normal solution of an over specified system of mal non-linear least-squares minimization by searching equations is easily derived in a least-squares sense us- the (α, β, γ) space for minima in squared error. With ing matrix notation. The total error E as the sum of all the measured Stokes vector (Si), the Rayleigh sky input residuals for m independent observations we get: vector (Ri) and a rotation matrix (Rij ) we define the error (ǫ) as: m 2 E = ǫi (14) i=1 2 3 X 2 2 ǫ (α,β,γ)= [Si − RiRij (α,β,γ)] (11) We solve the least-quares system for the unknown i=1 j=1 Mueller matrix element (Mj ) by minimizing the error X X with respect to each equation. The partial derivative of For n measurements, this gives us 3 × n terms in this Equation 13 for ǫi with respect to Mj is just the sky least squares problem. The Euler angles give identical input elements Rij . Taking the partial with respect to rotation matrix elements under the exchange of (α, β, γ) each input Stokes parameter we get: with ( -180◦+α,-β, -180◦+γ) as well as with additional multiples of 2π. This direct solution is easily solvable in principle but the ambiguities make implementing this ∂E ∂ǫi = 2 ǫi = −2 Rij Si − RikMk = 0 solution with existing software languages more cumber- ∂Mj ∂Mj i i k ! some than necessary. X X X As an alternative method to the direct least-squares so- (15) lution for Euler angles, we can do a two-step process that gives the same result but is much easier to implement. We have inserted a dummy sum over the index k. Mul- First we solve a system of equations for the Mueller ma- tiplying out the terms and rearranging gives us the nor- trix elements directly that are not subject to rotational mal equations: ambiguity. With the Mueller matrix elements in hand, R R M R S we can then perform a simple fit of a rotation matrix ij ik k = ij i (16) i k i elements to the derived Mueller matrix terms. This two- X X X step process allows us to have accurate starting guesses Which written in matrix notation is the familiar solu- to speed up the minimization process, resolve Euler angle tion of of a system of equations via the normal method: ambiguities and to estimate error propagation. When deriving the Mueller matrix elements of the RT S M = (17) telescope, one must take care that the actual derived RT R matrices are physical. For instance, there are various matrix properties and quantities one can derive to test This simple equation is very easy to implement with the physicality of the matirx (Givens & Kostinski 1993, a few lines of code provided your observation times are Kostinski et al. 1993, Takakura & Stoll 2009). Noise chosen to give you a range of input states for a well- and systematic uncertainty might give over-polarizing or conditioned inversion. The noise properties and inver- unphysical Mueller matrices such as the element above sion characteristics of this equation can be calculated in that is >1. By simply using Mueller matrix elements to advance of observations and optimized. We can write Mueller Matrices Using Sky Polarization 9 the matrix A with an implied sum over i observations estimated starting points but give the same best-fit Eu- for each term as: ler angles to better than one part per million. Note that by choosing Euler angle guesses in the proper quadrants, q q q u the derived rotation matricies avoid the ambiguity of (α, A = RT R = ri ri ri ri (18) qri uri uri uri β, γ) with ( -180◦+α,-β, -180◦+γ). The differences be-   tween rotation matrix elements and Mueller matrix ele- The solution to the equations for the three sets of ments for our observations typically much less than 0.1. Mueller matrix elements can be written as: The two-step solution gives us the same Euler angles as the direct least squares solution typically better than 0.2◦ QQ 1 qri qmi and allows for computationally inexpensive and easy im- Mj = = A− (19) UQ uri qmi plementation of the least-squares minimization.    

QU 1 qri umi Mj = = A− (20) UU ur um    i i 

QV 1 qri vmi Mj = = A− (21) UV ur vm    i i  As an example, if we compute the inverse of A and 1 multiply out A− for the QQ term we can write:

− (qri qmi )(uri uri ) (uri qmi )(qri uri ) QQ = − (22) (qri qri )(uri uri ) (qri uri )(qri uri ) In this manner, we can easily implement the usual ma- trix formalism with a time-series of daytime sky obser- vations to measure six Mueller matrix elements. With the Mueller matrix terms in hand, solving for accurate initial Euler angle guesses is straightforward. Fig. 10.— This Figure shows the squared error (ǫ2) for the two- As an example, by using QV and UV from one can solve step solution derived using Equation 26. Each panel shows a slice through the volume (α, β, γ) at four values of α. Error increases for α and β directly: from dark to light colors. Good solutions are found at α ∼ 0 and at α = ±180. QV tan(α)= − (23) UV In order to verify this method, we computed the error QV −UV sin(β)= = (24) functions E for both the direct solution and the two-step sin(α) cos(α) solution and verified the minima with a direct search of the entire Euler angle space. We created a grid of Euler QQ If one writes cγ as x then uses coefficients a = sαsβ angles from -180◦ to 180◦ (using 2 degree increments) and b = cα one can numerically solve a quadratic for and computed the error E as in Equation 11. Figure 10 sαsβ shows the derived errors as functions of β and γ for four and estimate of γ: different values of α. Each panel is color coded to show 2 2 − 2 − the square error E increasing from dark to light. The Eu- 0 = (b + 1)x (2ab)x + (a 1) (25) ler angles found using the two-step method are identical 1 to those found with the direct solve within the crude 2 The sign ambiguity in solving for γ=cos− (x) is re- ◦ solved by comparing the sign of the computed Mueller sampling we chose for this brute-force search. The error matrix elements with those of the rotation matrix de- functions are always continuous with only two minima in rived with these estimated Euler angles. The estimate the search volume showing convergence is easily achieved shown here only utilized QQ, QV and UV but provided using any minimization scheme. a guess for minimization routines typically accurate to better than 2◦ in our data set. In our two-step method, we next solved for the best fit Euler angles by doing a normal non-linear least-squares minimization by searching the (α, β, γ) space around our initial guesses for minima in the summed squared error:

2 3 2 E(α,β,γ)= [Mij − Rij (α,β,γ)] (26) i=1 j=1 X X By using the built-in IDL routines POWELL or AMOEBA, we can construct the error function and find the minima. Both of these minimization routines require 10 Harrington et al.

6. APPENDIX: PLANNING SKY OBSERVATIONS USING TIME DEPENDENCE FOR MODULATION 1 + 1 + 1 − 1 √3 √3 √3 In order to use this technique efficiently, one must con- 1 + 1 − 1 + 1 O =  √3 √3 √3  (30) sider the noise propagation when inverting a sequence 1 − 1 + 1 + 1 √3 √3 √3 of sky observations to derive telescope Mueller matrix  1 1 1   1 − − −  properties. There is an analogy between using the time-  √3 √3 √3  dependent Rayleigh sky to measure polarization proper-   ties of the a telescope and the retardances chosen to cre- One recovers the input Stokes vector from a series of intensity measurements by inverting the modulation ma- ate an efficient modulation scheme for polarization mea- O surements. The underlying mathematics is the same and trix ( ) provided it is a square and multiplying this in- one can easily derive the requirements on the observa- verse by the measured intensities. If the matrix is not tions needed to derive high-accuracy telescope Mueller square, one can simply solve the over-specified system of matrix measurements. Effectively, one is attempting equations via the normal least squares formalism: to derive properties of a matrix through what different OT I groups call demodulations, inversions or deprojections. S = (31) We outline here the analogy between polarimetric modu- OT O lation and calculating noise properties with the Rayleigh Even for non-square matrices, we can define a demod- sky to determine a quality observing sequence for deriv- ulation matrix that captures the transfer properties of ing telescope Mueller matrices. the modulation scheme: Polarimeters modulate the incoming polarization state T 1 T via retardance amplitude and orientation changes. This Dij = [O O]− O (32) retardance modulation translated in to varying inten- sities using an analyzer such as a polarizer, polariz- In our case, the Rayleigh sky input parameters become O R ing beam splitter or crystal blocks such as Wollaston the modulation matrix ( ij = ij ) and the formalism prisms or Savart plates. These modulation schemes for noise propagation developed in many studies such as can vary widely. For example, Compain et al. 1999, del Toro Iniesta & Collados 2000 apply. th del Toro Iniesta & Collados 2000, De Martino et al. As an example, we will use the Jan. 27 2010 Mueller 2003, Nagaraju et al. 2007 and Tomczyk et al. 2010 matrices derived in Figure 5. The sky polarization ro- overview optimal schemes, error propagation and out- tated by about 33◦ during this period and the three cal- line schemes to maximize or balance polarimetric effi- culated Rayleigh sky inputs scaled to 100% polarization ciency and create polychromatic systems. There have are: been many implementations of achromatic and polychro- matic designs in both stellar and solar communities (cf. qr1 ur1 +0.980 +0.201 Gisler et al. 2003, Hanaoka 2004, Xu et al. 2006). In Rij = qr2 ur2 = +0.847 +0.534 (33) the notation of these studies, the instrument modulates qr3 ur3 ! +0.716 +0.698 ! the incoming polarization information in to a series of measured intensities (Ii) for i independent observations If each measurement has the same noise σ and there via the modulation matrix (Oij ) for j input Stokes pa- are n total measurements then the noise on each demod- rameters (Sj ): ulated parameter (σi) becomes: I O S n i = ij j (27) 2 2 D2 σi = nσ ij (34) This is exactly analogous to our situation where we j=1 have changed the matrix indices to be i independent X Stokes parameter measurements for j different sky input And the efficiency of the observation becomes: Stokes parameters: 1 n − 2 2 S = R M (28) ei = n D (35) i ij j  ij  j=1 In most night-time polarimeters, instruments choose a X modulation matrix that separates and measures individ- For instance, the normal modulation sequence of equa- ual parameters of the Stokes vector: tion 29 used by most night time spectropolarimeters gives n=6 and e2 = [1, 1 , 1 , 1 ]. The efficiency balanced scheme 1+1 0 0 i 3 3 3 − gives the same relative efficiencies as the normal modu- 1 1 0 0 lation scheme but uses only 4 exposures instead of 6. In O  1 0+1 0  the case of our Jan. 27th observations considering only ij = 1 0 −1 0 (29)   qu terms our demodulation matrix is:  1 0 0+1   1 0 0 −1    −   D RT R 1RT +1.23 +0.16 0.48 Other instruments choose only four measurements: the ij = [ ]− = −1.49 +0.43 +1.54 (36) minimum number of exposures required to measure the   Stokes vector. In these schemes, one can balance the With this demodulation matrix the efficiency for u efficiency of the measurement to minimize the noise on is only 60% worse than q and we have efficiencies of each Stokes parameter: ei = [0.43, 0.26] computed with Equation 35. The Mueller Matrices Using Sky Polarization 11

Mueller matrix derived from these Rayleigh sky obser- vations will have similar noise properties between Q and U terms. One must take care with this technique to build up observations over a wide range of solar locations so that the inversion is well conditioned. The path of the sun throughout the day will create regions of little input sky Stokes vector rotation causing a poorly constrained in- version with high condition number. For instance, at our location in the tropics the sun rises and sets with- out changing azimuth until it rises quite high in the sky. We are constrained to observing in early morning and late evening with the dome walls raised since we may not expose the telescope to the sun. This causes input vectors at east-west pointings to be mostly q oriented with little rotation over many hours. Observations at other times of the year or at higher solar elevations are required to have a well conditioned inversion. One can easily build up the expected sky input polarizations at a given observing site with the Rayleigh sky polarization equations. Then it is straightforward to determine the modulation matrix and noise propagation for a planned observing sequence to ensure a well-measured telescope matrix with good signal-to-noise. 12 Harrington et al.

7. APPENDIX: LOVIS SPECTROPOLARIMETER CHARACTERIZATION In order to expand the capabilities of our spectropo- larimeter, we have adapted the optics for use at low spec- tral resolution. The polarimeter unit for the spectro- graph is a Savart plate providing dual-beam analyzing mounted behind the entrance slit with either two rotat- ing achromatic retarders or two liquid-crystal variable re- tarders (LCVRs) outlined in Harrington & Kuhn 2008; Harrington et al. 2010. We have modified the HiVIS echell´ehousing to allow a flat mirror to be mounted by- passing the echell´ewithout disrupting the optical path. We call this new mode LoVIS.

Fig. 12.— This Figure shows an example extracted spectrum of the spectrophotometric standard star HR7950 taken on Sept. th 4 2009. The Hα, Na D lines and the atmospheric A-band are identified.

This change in resolution allows us to observe much fainter targets or observe sources at a much higher ca- dence. When observing the daytime sky with AEOS and LoVIS, we can achieve a spectropolarimetric precision of 0.1% with a 5 second exposure time using the 1.5“ slit. The polarimetric images sequences finish in one minute and the bulk of the overhead is from rotating the achro- matic wave plates in the typical 6-exposure sequence. To calibrate this new polarimetric mode and the tele- Fig. 11.— This Figure shows example LoVIS data when ob- scope polarization we used LoVIS to gather a wide range th serving a spectrophotometric standard star on Sept. 4 2009. of sky and calibration observations from August 2009 to Wavelength increases from right to left. The Na D lines, Hα and the atmospheric A-band can be seen by eye near spectral pixels February 2010. All observations were taken with the 2300, 1400 and 900 respectively. Apogee detector described in Harrington et al. 2010. We have sky observations taken solely at the Zenith as well as observations on a grid of altitudes and azimuths. On several occasions we did observations for 3 or 4 hours With a flat mirror replacing the echell´e, the only dis- continuously before sunset to accumulate a wide range persive element in the spectrograph is the cross-disperser st of polarization inputs at multiple pointings. working in 1 order. Only a single order is imaged on In order to decouple the LoVIS and AEOS polariza- the detector making the illuminated region of the device tion measurements, we did a number of tests with our much smaller. Figure 11 shows a raw data frame from a polarization calibration unit mounted at the slit as well spectrophotometric standard star (HR7940) taken with as at the entrance port to our instrument at the calibra- LoVIS in spectropolarimetric mode using the Apogee de- tion lamp unit. We did tests both without and with the tector. There is a single spectral order seen with orthogo- image rotator inserted in the the path. This image rota- nally polarized spectra being displaced roughly 70 pixels tor is the source of most of the cross-talk in the LoVIS spatially by the Savart plate. Wavelength increases from instrument. Removing it makes the polarization prop- right to left with the Hα line seen near spectral pixel 1400 erties of LoVIS much more benign as shown in Figures and the 6870A˚ atmospheric band seen near pixel 900. We 15 and 16 of Harrington et al. 2010. There is a trans- have two separate cross-dispersers available on the rota- missive window that separates the coud´erooms from the tion stage blazed for 6050A˚ and 8600A˚ ( Hodapp et al. central optical path to the telescope. This window is 2000, Thornton et al. 2003). directly after the last fold mirror of the telescope, m7, The new LoVIS mode gives us much higher sensitivity and can be set to either BK7 or infrasil. This window and allows for much faster calibrations at spectral res- can also have a polarimetric effect on the beam and both olutions of 1,000 to 3,000 depending on the slit. The windows were also tested. Effectively we find that the 1.5“ slit had spectral resolutions of 850 and 990 derived polarization properties derived for HiVIS are similar to from Thorium-Argone (ThAr) lines at 5700A˚ and 7000A.˚ LoVIS and that we can efficiently reproduce polarization These lines were spectrally sampled with roughly 9.5 measurements with both rotating achromatic wave plates pixels per Gaussian Full-Width-Half-Max at each wave- and LCVR modulators. length. The IDL reduction scripts we wrote for HiVIS We also measured the induced polarization of the en- outlined in Harrington & Kuhn 2008 were adapted to tire system when illuminated by a point-source utilizing this new low-resolution data. Figure 12 shows the ex- unpolarized standard stars. As an example, we show ob- tracted spectrum from the raw data of Figure 11. servations of many unpolarized standard stars at many Mueller Matrices Using Sky Polarization 13

Fig. 13.— This Figure shows the telescope induced polarization. We measured quv spectra for many different unpolarized standard stars at many different pointings observed with LoVIS on Sept. 5th 2009. Each Stokes parameter is shown in percent. The induced polarization is at the few percent level with some dependence on wavelength. pointings from September 5th 2009 in Figure 13. The measured polarization is typical of results obtained on any individual night. Consistent with our previous HiVIS measurements from Harrington & Kuhn 2008, the in- duced polarization is around a few percent, has some mild chromatic properties and does depend on telescope focus, the seeing and stellar location on the slit. In summary, the polarization properties of LoVIS are similar to HiVIS. The induced polarization is below a few percent. The polarization calibration optics give very pure inputs that are reproduced with both achro- matic wave plate and LCVR modulators. This new low- resolution mode works efficiently and allows us to obtain calibration observations at a much faster cadence. 14 Harrington et al.

REFERENCES Aben I. et al., 1999, Geophysical Res. Lett., 26, 591 Litvinov P. et al., 2010, JQSRT, 111, 529 Aben I. et al., 2001, Geophysical Res. Lett., 28, 519 Liu Y. & Voss K., 1997, ApOpt, 36, 8753 Berk A. et al., 2006, SPIE, 6233, 49B Nagaraju K. et al., 2007, Bull. Astr. Soc. India, 35, 307 Boesche E. et al., 2006, ApOpt, 47, 3467 Ota Y. et al., 2010, JQSRT, 111, 878 Clarke D., Stellar Polarimetry, 2009, Wiley Patat F. & Romaniello R., PASP, 2006, 118, 146 Collett E., Polarized Light: Fundamentals and Applications, Peltoniemi J. et al., 2009, JQSRT, 110, 1940 1992, CRC Pomozi I. et al., 2001, J. Exp. Biology, 204, 2933 Compain E. et al., 1999, ApOpt, 38, 3490 Pust N.J. & Shaw J.A., 2005, SPIE, 5888, 295 Coulson K.L., 1980, ApOpt, 19, 3469 Pust N.J. & Shaw J.A., 2006a, ApOpt, 45, 5470 Coulson K.L., Polarization and Intensity of Light in the Pust N.J. & Shaw J.A., 2006b, SPIE, 2006, 6240, 1157 Atmosphere, A. Deepak Publishing, 1988, Hampton Virginia Pust N.J. & Shaw J.A., 2007, SPIE, 6682, 1159 Cronin T.W. et al., 2005, SPIE, 5888, 389 Pust N.J. & Shaw J.A., 2008, ApOpt, 47, 190 Cronin T.W. et al., 2006, ApOpt, 45, 5582 Pust N.J. & Shaw J.A., 2009, SPIE, 7461, 10P Dahlberg A.R. et al. 2009, SPIE, 7461, 6D Roelfsema R. et al., 2010, SPIE, 7735, 1559 del Toro Iniesta, J.C., and Collados M., 2010, ApOpt, 39, 10 Salinas S.V. & Liew S.C., 2007, JQSRT, 105, 414 DeMartino A. et al., 2003, Optics Lett., 28, 616 S´anchez Almeida J. et al., 1991, Solar Physics, 134, 1 Elmore D.F. et al., 1992, SPIE, 1746, 22 Selbing J., SST⁀ polarization model and polarimeter calibration, Elmore D.F. et al., 2010, SPIE, 7735, 1215 Arxiv 1010.4142v1 Fetrow M.P. et al., 2002, SPIE, 4481, 149F Schmidt G.D. et al., 1992, AJ, 104, 1563 Fossati L. et al., 2007, ASP Conf. Proc., 364, 436 Shaw J.A. et al., 2010, SPIE, 7672, 9S G´al J. et al. 2001, Proc. R. Soc. Lond. 457, 1385 Shukurov A.Kh. & Shukurov K.A., 2006, Atmosph. & Oceanic Gil-Hutton R. & Benavidez P., 2003, MNRAS, 345, 97 Phys., 42, 68 Giro E. et al., 2003, SPIE, 4843, 456 Socas-Navarro H. et al., 2005, SPIE, 5901, 1217 Gisler D. et al. 2003, SPIE, 4843, 45 Socas-Navarro H. et al., 2006, SoPh, 235, 55 Givens C.R. & Kostinski A.B., 1993, J. Mod. Opt., 40, 471 Socas-Navarro H., 2005a, JOSA, 22, 539 Hanaoka Y., 2004, SoPh, 222, 265 Socas-Navarro H., 2005b, JOSA, 22, 907 Harrington D.M. & Kuhn J.R., 2007, ApJL, 667, L89 Suhai B. & Horv´ath G., 2004, JOSA A, 21, 1669 Harrington D.M. & Kuhn J.R., 2008, PASP, 120, 89 Takakura Y. & Stoll M.-P., 2009, ApOpt, 48, 1073 Harrington D.M. & Kuhn J.R., 2009a ApJS, 180, 138 Thornton R.J. et al., 2003, SPIE, 4841, 1115 Harrington D.M. & Kuhn J.R., 2009b ApJ, 695, 238 Tinbergen J., 2007, PASP, 119, 1371 Harrington D.M. et al., 2006, PASP, 118, 845 Tomczyk S. et al., 2010 ApOpt, 49, 20 Harrington D.M. et al., 2009, ApJ, 704, 813 Ougolnikov O.S. & Maslov I.A., 2002, Cos. Research, 40, 224 Harrington D. M. et al., 2010, PASP, 122, 420 Ougolnikov O.S. & Maslov I.A., 2005a, Cos. Research, 43, 17 He X. et al., 2010, JQSRT, 111, 1426 Ougolnikov O.S. & Maslov I.A., 2005b, Cos. Research, 43, 404 Heged¨us R. et al. 2007, ApOpt, 46, 2717 Ougolnikov O.S. & Maslov I.A., 2009a, Atmosph. & Ocean. Opt, Hodapp K.W. et al., 2000, SPIE, 4008, 778 22, 192 Horv´ath et al., 2002a, ApOpt, 41, 543 Ougolnikov O.S. & Maslov I.A., 2009b, Cos. Research, 47, 198 Horv´ath et al., 2002b, JOSA A, 19, 2085 Ugolnikov O.S. et al., 2004, JQRST, 88, 233 Hsu J-C. & Breger M., 1982, ApJ, 262, 732 van Harten G. et al., 2009, PASP, 878, 377 Ichimoto K. et al., 2008, SoPh, 249, 233 Vermeulen A. et al., 2000, ApOpt, 39, 6207 Joos F. et al., 2008, SPIE, 7016, 48J Wu B. & Jin Y., 1997, ApOpt, 36, 7009 Keller C.U. & Snik F., 2009, ASP Conf Proc, 405, 371 Xu C. et al., 2006, ApOpt, 45, 8428 Kisselev V. & Bulgarelli B., 2004, JQSRT, 85, 419 Zeng J. et al., 2008, Geophysical Res. Lett, 35, L20801 Kostinski A.B. et al., 1993, ApOpt, 32, 1646 Kuhn J.R. et al., 1993, SoPh, 153, 143 Lee R.L. Jr., 1998, ApOpt, 37, 1465