Use of the surface PSD and incident angle adjustments to investigate near specular scatter from smooth surfaces

Similar documents
Optical Scattering. Analysis. Measurement and SPIE PRESS. John C. Stover THIRD EDITION. Bellingham, Washington USA

SWING ARM OPTICAL CMM

WORCESTER POLYTECHNIC INSTITUTE

High spatial resolution measurement of volume holographic gratings

Class 11 Introduction to Surface BRDF and Atmospheric Scattering. Class 12/13 - Measurements of Surface BRDF and Atmospheric Scattering

Chapter 36. Image Formation

Coupling of surface roughness to the performance of computer-generated holograms

ksa MOS Ultra-Scan Performance Test Data

MEMS SENSOR FOR MEMS METROLOGY

Freeform metrology using subaperture stitching interferometry

Polarized light scattering measurements of roughness, subsurface defects, particles, and dielectric layers on silicon wafers

Mu lt i s p e c t r a l

Contrast Optimization: A faster and better technique for optimizing on MTF ABSTRACT Keywords: INTRODUCTION THEORY

Condenser Optics for Dark Field X-Ray Microscopy

2/26/2016. Chapter 23 Ray Optics. Chapter 23 Preview. Chapter 23 Preview

Engineered Diffusers Intensity vs Irradiance

LIGHT SCATTERING THEORY

Null test for a highly paraboloidal mirror

SEMI M is due for 5 year review. The document was reviewed by the task force. The following changes are made.

specular diffuse reflection.

COMPARISON OF TWO INSTRUMENT DESIGNS FOR NON-CONTACT MEASUREMENT OF GOSSAMER MIRRORS

dq dt I = Irradiance or Light Intensity is Flux Φ per area A (W/m 2 ) Φ =

PHYSICS. Chapter 34 Lecture FOR SCIENTISTS AND ENGINEERS A STRATEGIC APPROACH 4/E RANDALL D. KNIGHT

Keywords: Random ball test, interferometer calibration, diffraction, retrace, imaging distortion

Geometrical modeling of light scattering from paper substrates

Lecture Outline Chapter 26. Physics, 4 th Edition James S. Walker. Copyright 2010 Pearson Education, Inc.

Measurement of Highly Parabolic Mirror using Computer Generated Hologram

SIMULATION AND VISUALIZATION IN THE EDUCATION OF COHERENT OPTICS

Chapter 36. Diffraction. Dr. Armen Kocharian

Council for Optical Radiation Measurements (CORM) 2016 Annual Technical Conference May 15 18, 2016, Gaithersburg, MD

ENHANCEMENT OF DIFFUSERS BRDF ACCURACY

NEW OPTICAL MEASUREMENT TECHNIQUE FOR SI WAFER SURFACE DEFECTS USING ANNULAR ILLUMINATION WITH CROSSED NICOLS

( ) = First Bessel function, x = π Dθ

Lecture 4. Physics 1502: Lecture 35 Today s Agenda. Homework 09: Wednesday December 9

Optical simulations within and beyond the paraxial limit

MODULE 3. FACTORS AFFECTING 3D LASER SCANNING

Using a multipoint interferometer to measure the orbital angular momentum of light

PHYSICS 1040L LAB LAB 7: DIFFRACTION & INTERFERENCE

Ch. 25 The Reflection of Light

PHYS 219 General Physics: Electricity, Light and Modern Physics

A Study of Scattering Characteristics for Microscale Rough Surface

The Law of Reflection

Optics II. Reflection and Mirrors

Light Tec Scattering measurements guideline

Philpot & Philipson: Remote Sensing Fundamentals Interactions 3.1 W.D. Philpot, Cornell University, Fall 12

M1 Microroughness and Dust Contamination

Chapter 32 Light: Reflection and Refraction. Copyright 2009 Pearson Education, Inc.

Analysis of Cornell Electron-Positron Storage Ring Test Accelerator's Double Slit Visual Beam Size Monitor

Light: Geometric Optics

Ray Optics I. Last time, finished EM theory Looked at complex boundary problems TIR: Snell s law complex Metal mirrors: index complex

Extreme Ultraviolet Radiation Spectrometer Design and Optimization

EO-1 Stray Light Analysis Report No. 3

Lecture Notes (Reflection & Mirrors)

Hyperspectral interferometry for single-shot absolute measurement of 3-D shape and displacement fields

Southern African Large Telescope. PFIS Distortion and Alignment Model

Formulas of possible interest

Applications of Piezo Actuators for Space Instrument Optical Alignment

Eric Lindmark, Ph.D.

Lecture 4 Recap of PHYS110-1 lecture Physical Optics - 4 lectures EM spectrum and colour Light sources Interference and diffraction Polarization

dq dt I = Irradiance or Light Intensity is Flux Φ per area A (W/m 2 ) Φ =

Chapter 26 Geometrical Optics

FRED Slit Diffraction Application Note

3/10/2019. Models of Light. Waves and wave fronts. Wave fronts and rays

D&S Technical Note 09-2 D&S A Proposed Correction to Reflectance Measurements of Profiled Surfaces. Introduction

At the interface between two materials, where light can be reflected or refracted. Within a material, where the light can be scattered or absorbed.

SUPPLEMENTARY INFORMATION

Conceptual Physics Fundamentals

Optical Topography Measurement of Patterned Wafers

PHYSICS. Chapter 33 Lecture FOR SCIENTISTS AND ENGINEERS A STRATEGIC APPROACH 4/E RANDALL D. KNIGHT

TEAMS National Competition Middle School Version Photometry Solution Manual 25 Questions

PHY 171 Lecture 6 (January 18, 2012)

College Physics 150. Chapter 25 Interference and Diffraction

Michelson Interferometer

Nicholas J. Giordano. Chapter 24. Geometrical Optics. Marilyn Akins, PhD Broome Community College

Assembly of thin gratings for soft x-ray telescopes

Lenses lens equation (for a thin lens) = (η η ) f r 1 r 2

SUPPLEMENTARY INFORMATION

The NA PV Materials TC Chapter reviewed and recommended to issue for reapproval ballot.

Calibration of a portable interferometer for fiber optic connector endface measurements

The roughness model in SHADOW

Chapter 26 Geometrical Optics

ratio of the volume under the 2D MTF of a lens to the volume under the 2D MTF of a diffraction limited

Chapter 36. Diffraction. Copyright 2014 John Wiley & Sons, Inc. All rights reserved.

Ray Optics. Lecture 23. Chapter 23. Physics II. Course website:

Optics Vac Work MT 2008

Advanced modelling of gratings in VirtualLab software. Site Zhang, development engineer Lignt Trans

MICROSPHERE DIMENSIONS USING 3D PROFILOMETRY

Textbook Reference: Physics (Wilson, Buffa, Lou): Chapter 24

Reflection & Mirrors

OPTI 513R / Optical Testing

X-ray Optics. How do we form an X-Ray Image?

Experiment 8 Wave Optics

Light Tec Scattering measurements guideline

Light Tec Scattering measurements guideline

TEAMS National Competition High School Version Photometry Solution Manual 25 Questions

arxiv: v1 [astro-ph.im] 3 Oct 2016

Fundamental Optics for DVD Pickups. The theory of the geometrical aberration and diffraction limits are introduced for

Study of Air Bubble Induced Light Scattering Effect On Image Quality in 193 nm Immersion Lithography

axis, and wavelength tuning is achieved by translating the grating along a scan direction parallel to the x

Lecture 16 Diffraction Ch. 36

Transcription:

Use of the surface PSD and incident angle adjustments to investigate near specular scatter from smooth surfaces Kashmira Tayabaly a, John C. Stover b, Robert E. Parks a,c, Matthew Dubin a, James H. Burge* a a College of Optical Sciences, University of Arizona, 1630 E University Blvd, Tucson, AZ 85721; b The Scatter Works, 2100 N. Wilmot, Suite 202, Tucson, AZ 85712; c Optical Perspectives Group, LLC, 7011 E. Calle Tolosa, Tucson, AZ 85750 ABSTRACT The Rayleigh Rice vector perturbation theory has been successfully used for several decades to relate the surface power spectrum of optically smooth reflectors to the angular resolved scatter resulting from light sources of known wavelength, incident angle and polarization. While measuring low frequency roughness is relatively easy, the corresponding near specular scatter can be difficult to measure. This paper discusses using high incident angle near specular measurements along with profile generated surface power spectrums as a means of checking a near specular scatter requirement. The specification in question, a BRDF of 1.0 sr -1 at 2 mrad from the specular direction and at a wavelength of 1µm, is very difficult to verify by conventional scatter measurements. In fact, it is impractical to directly measure surface scatter from uncoated Zerodur because of its high bulk scatter. This paper presents profilometer and scatterometer data obtained from coated and uncoated flats at several wavelengths and outlines the analysis technique used to check this tight specification. Keywords: BRDF, near specular scatter, surface power spectrum, Rayleigh Rice, vector perturbation theory, metrology 1. INTRODUCTION Limiting scattered light in an optical system is often essential but also, most of the time, challenging. Each and every media or surface in the light path is likely to introduce scattered light that will degrade the image quality. This is even more the case when the object is faint relative to its surroundings. This scattered light limitation is particularly important as it applies to the Advanced Technology Solar Telescope (ATST) [1]. A thorough analysis has established that microroughness of the ATST primary mirror is one the main sources of scattered light [2]. Therefore, in order to meet the science requirement for the ATST, it was critical to specify the primary mirror in terms of its scatter function or more precisely, BRDF. Specifically, this 4 m off-axis parabolic mirror, made of Zerodur, has to meet a BRDF of less than 1.0 sr -1 at 2 mrad from the specular direction at a wavelength of 1 µm. Several techniques may be used for measuring the BRDF of a surface. However, such a measurement is difficult when it applies to near specular measurements for large optics where the bare glass has a low reflectivity. However, making coated replicas of the mirror s surface would be a risky, time consuming and expensive approach to test the BRDF of this surface. It was therefore crucial to find an alternative method for determining its BRDF at near specular angles. * jburge@optics.arizona.edu, phone: (520) 621-8182, fax: (520)-621-3389

Fortunately, when considering optically smooth, clean, first surface mirrors such as the ATST primary, the widely used Rayleigh-Rice vector perturbation theory establishes a one-to-one relationship between the BRDF and the topography of a surface or, to be more specific, its power spectral density (PSD). This relationship is also known as the Golden Rule and is written below for isotropic surfaces (Eq 1.1): BRDF(f) = 16π2 cos θ λ 4 i cos θ s Q PSD 2D (f) Equation 1.1 where λ is the wavelength of the incoming light with units of length, θ i and θ s the incident and scatter angles; f is the spatial frequency in polar coordinates and is given in inverse unit of length and Q represents a unitless polarization factor. PSD 2D is the 2D power spectral density with units of length to the fourth power. We will most typically use µm 4, but it also common to find the results in Angstroms 2.µm 2 (10 8 µm 4 ). It is defined as: PSD 2D f x, f y = lim Area 1 Area F[Z(x, y)] 2 where Area represents the surface s area considered, Z(x,y) is the surface height fluctuation at each point (x,y) of the surface in unit of length. f x and f y are the spatial frequency in x and y in inverse unit of length, f = f x2 + f y2. F denotes the Fourier Transform taken of Z(x,y). As the measurement of low frequency roughness is relatively easy, this approach provides a simple alternative to BRDF measurements for surfaces meeting the Rayleigh-Rice requirements and allows reformulating the ATST BRDF specification in terms of PSD. This term characterizes the surface microroughness. The ATST primary mirror is a 4 m diameter off-axis parabola, with a 16 m radius of curvature and a 4 m offset from the parent optical axis. In this configuration, the incident light reaches the center of the mirror with an angle of incidence of approximately 14. Assuming a 1 µm wavelength with an incident beam of θ i = 14 from the normal to the surface at the center of the mirror, and the mirror surface being highly reflective (Q=1), the ATST primary mirror BRDF specification becomes: BRDF f specification = 2 cycles/mm 1.0 sr 1 Equation 1.2 From Eq 1.1 we obtain a specification on the mirror surface finish of: PSD f specification = 2 cycles/mm 0.0067 µm 4 Equation 1.3 This paper presents diverse measurements put in place to investigate the possibility to check the ATST tight specification. The results obtained confirm the feasibility of using surface PSD to investigate near specular scatter introduced by optically-smooth surfaces. Thus, Section 2 describes the scatterometer used for direct BRDF measurements and presents how the geometry of the instrument was used to achieve near specular scatter measurements; while Section 3 details the MicroFinish Topographer (MFT) [3], a temporal phase shifting interferometer used for profilometric measurements. Data obtained from the two approaches just mentioned, are processed and compared in terms of surface power spectral density (PSD) as presented in Section 4. 2. NEAR SPECULAR DIRECT SCATTER MEASUREMENTS The difficulty in measuring scatter near the specular beam is separating scatter from the sample (signal) from scatter created by the scatterometer optics (noise). This is typically done by measuring the no sample scatter, called an instrument signature, and then comparing the measured sample scatter to the signature multiplied by the sample specular reflectance. There is usually an obvious point where the two results separate. This angle depends on both the scatterometer and the sample and typically varies between 0.05 and 3.0 degrees from the center of the specular beam. Instrument signature depends on wavelength and instrument field-of-view as well as the quality of instrument optics [4]. Near specular sample scatter depends on incident angle as well as wavelength and sample scatter characteristics. In this particular case the mirror BRDF specification of 1.0 sr -1 at 2 mrad (0.11 ) is similar to the signature of many

scatterometers. Fortunately the sample (but not the signature) dependence on incident angle can help solve this problem for the current situation where scatter is primarily caused by mild surface roughness. To see how this can be done, consider the incident plane grating equation which relates grating frequency, f, to the wavelength, λ, the incident angle θ i and the scattering angle θ s. sin θ s sin θ i f = λ Equation 2.1 Substituting in the specification values (λ = 1.0 µm, θ s = 0.11 deg, and θ i = 0) gives a spatial frequency of 1.9 cycles/mm. So the specification is essentially one of limiting roughness having this spatial frequency. Now if the incident angle is increased to 80 degrees at the same wavelength then we see that the same spatial frequency scatters at 80.65 degrees or 0.65 degrees from specular, which is a much easier measurement to separate signal from the instrument signature. The location that a given spatial frequency scatters from specular also depends on wavelength, but changes in wavelength do not offer a big measurement advantage as both scattering angle and the diffraction limited spot size change. As pointed out in the introduction, the Rayleigh Rice vector perturbation theory can be used to account for changes in BRDF with wavelength at a specific spatial frequency. 2.1 Scatterometer description Scatter measurements were taken on several samples at two wavelengths to illustrate the advantage of incident angle changes. The instrument is a CASI Scatterometer [4] with a sample receiver distance of 294 mm and a small near specular detector aperture of 340 µm. Laser sources of 488 nm and 633 nm were used for these experiments. Detector step sizes are programmable and can be made as small as 0.01 degrees. Larger detector apertures are used farther from specular and are inserted at programmable locations. The system centers the small aperture on the specular reflection and then takes measurements relative to this location. Signature symmetry is a good indication of correct specular location. System setup and calibration are checked by measuring a diffuse Lambertian sample with a BRDF of 0.32 sr -1. 2.2 Measurement of the Instrument Signature The CASI instrument signature was measured for two wavelengths λ = 488nm and λ = 632.8 nm as shown in Figure 2.1. Figure 2.1.Instrument signature measured for two wavelengths (λ b = 488 nm and λ r = 632.8 nm) as a function of polar angles in degrees. The green area indicates the angular region [0-2] mrad from specular.

A slight asymmetry in the measurements taken at positive and negative angles from specular is also observable, and may be due to a slight misalignment in the setup when collecting the data. This is of little consequence as the angle of interest (2 mrad) for ATST-BRDF specification lies in the region where scatter introduced by the instrument signature is prevalent and clearly constitutes a problem for near specular measurements. However, this issue does not exist with a profilometer measurement for BRDF prediction using the Rayleigh-Rice theory. Thus, the MicroFinish topographer, an interferometric microscope was used to investigate the feasibility of measuring near specular scatter. 3. PROFILOMETRIC MEASUREMENTS- MICRO-FINISH TOPOGRAPHER (MFT) The roughness measurements were made with a temporal phase shifting interferometer called a MicroFinish Topographer (MFT) that used 4Sight [5] software for data capture and analysis. The MFT was used because it is designed to sit directly on the surface being measured and will be used to measure the roughness of the 4 m diameter ATST mirror without having to make replicas. Figure 3.1 shows the MFT as it was used to measure the roughness of an 8.4 m diameter mirror. Because the MFT is supported by the surface being measured, it is largely immune to vibration and once adjusted for piston and tilt, it can be moved to other locations on the surface without having to make further adjustments to piston and tilt. Other than the mechanical design feature of sitting on the surface being measured, the MFT is a conventional surface roughness interferometer that uses commercial interferometric objectives, a Kohler illumination system with a red LED and a CCD camera. Figure 3.1- MFT picture measuring the surface topography of one of GMT s (8.4 m in diameter) mirrors The MFT allows accurate roughness measurements on a wide range of spatial frequencies [6] and, in particular, at spatial frequencies corresponding to near specular directions. Therefore, as shown in the next section, it is possible to use profilometric measurements to predict near specular scatter of smooth surfaces when the corresponding scatter measurement is usually limited by the scatterometer instrument signature.

4. RESULTS Both scatter and surface measurements were taken on two isotropic surfaces meeting the Rayleigh-Rice requirements: a Silicon wafer as well as a Zerodur optical flat partially coated. The results from those two approaches were then compared in terms of PSD on a logarithmic scale. The PSD from surface topography measurements (PSD surface ) is calculated following the steps below [4] [7]: 1. Measurement of Z(x,y) representing the surface topography with piston, tip/tilt and power removed. The p x, p y are the pixel sizes in x and y directions. N x, N y are the surface map sizes in terms of pixels in x and y directions 2. Calculation of PSD 2D such that: PSD 2D (fx, fy) = p xp y FFT N x N 2D [Z(x, y)] 2 y 3. Change PSD 2D map coordinates from Cartesian to Polar coordinates 4. Azimuthal average of PSD 2D to obtain PSD surface. The PSD scatter from the scatterometer BRDF measurement is calculated using the Golden Rule (Eq 2.1). 4.1 Silicon wafer Silicon wafers were one of the best polished, cleanest, most isotropic, highly reflective and cheapest surfaces available to us. A silicon wafer was subjected to profilometer (Figure 4.1) and scatterometer measurements for wavelengths (λ r = 633 nm and λ b = 488 nm) and 5 incidence angle. As shown in figure 4.2 below, surface and scatter measurements are consistent over the spatial frequency range not affected by the scatterometer signature (f 6 cycles/mm). In other words, scatter introduced by surface roughness is correctly predicted by the profilometer measurement. Similarly, the near specular scatter prediction of Rayleigh-Rice for the surfaces were verified. In order to shift the frequency corresponding to the instrument signature limitation to a frequency below f specification for ATST primary, the scatterometer must be used at very high angles of incidence - 84 for both 633 nm and 488 nm wavelengths. A comparison of the data (Figure 4.3) still shows satisfying agreement between the prediction from profilometer measurements and real scatter data at frequencies corresponding to near specular angles, for a smooth optical surface meeting the Rayleigh-Rice requirements. S co a e su ace 5 object e s 0.18 200 0.175 400 600 0.17 800 0.165 1000 1200 1400 0.16 0.155 1600 0.15 200 400 600 800 1000 Figure 4.1- Silicon wafer surface measured with the MFT 2.5x objective (RMS=1.44nm). Piston, Tip/Tilt, Power, Astigmatism, Coma and Spherical are removed to account for the MFT partial calibration for low frequencies.

10 6 10 4 Instrument Signature zone 2D ATST spec scatter λ= 488 nm 5 surface measurement 10 2 PSD (µ m 4 ) 10 0 10-2 10-4 X: 0.001939 Y: 0.01209 10-6 10-8 10-3 10-2 10-1 f in 1/µm Figure 4.2- Silicon wafer: PSD2D from surface (MFT 2.5x objective) and scatter measurement (θi = 5, λb = 488 nm) comparison 10 2 10 0 Scatter positive f 84 λ=488nm Scatter negative f 84 λ=488nm ATST spec 2.5x surface measurement PSD (µ m 4 ) 10-2 10-4 X: 0.001939 Y: 0.01209 10-6 10-8 10-3 10-2 10-1 f in 1/µm Figure 4.3 Silicon wafer: PSD2D from surface (MFT 2.5x objective) and scatter measurement (θi= 84, λb = 488 nm) comparison

4.2 Zerodur Coated versus uncoated A Zerodur optical flat (λ/20) was measured with the MFT before and after coating. The coating introduced pinholes (Figure 4.4a and 4.4b) changing the surface topography and making it slightly rougher. Consequently, the PSD of the surface has been modified. 0.2 j 0.25 200 200 0.245 400 600 0.15 400 600 0.24 0.235 0.23 800 800 0.225 1000 1200 1400 0.1 1000 1200 1400 0.22 0.215 0.21 0.205 200 400 600 800 1000 0.05 200 400 600 800 1000 1200 0.2 Figure 4.4a - Uncoated Zerodur 1" optical flat RMS =2.1 nm measured with the MFT 2.5x objective over a 4.4x3.3 mm 2 FOV. Piston, Tip/Tilt, Power are removed. Surface height is in waves at 635 nm. Figure 4.4b - 50 nm nickel coated Zerodur, 1" optical flat RMS =3.2 nm measured with the MFT 2.5x objective over a 4.4x3.3 mm 2 FOV. Piston, Tip/Tilt, Power are removed. Surface height is in waves. A comparison of the PSD surface obtained from uncoated and coated Zerodur measurement shows that the nickel coated sample introduced a hump in its PSD from 5 cycles/mm where this was not the case for bare Zerodur surface (Figure 4.5). As expected the coated surface which was rougher has a PSD higher than for the uncoated area on mid-range frequencies. 10 6 10 4 MFT measured uncoated surface MFT measured coated surface ATST spec 10 2 PSD in µm 4 10 0 10-2 10-4 10-6 10-4 10-3 10-2 10-1 10 0 f in µm Figure 4.5- Zerodur: PSD surface for uncoated and coated surface measured with the MFT 2.5x objective over a 4.4x3.3 mm 2 FOV.

Scatter versus surface measurements The PSD surface of the coated surface was compared with PSD scatter similarly to the Silicon wafer measurements. As five different spots were measured on the coated area of the Zerodur with the MFT, the PSD surface considered was the averaged PSD adding the standard deviation (2σ) of those five measurements. In Figure 4.6 surface and scatter PSDs are plotted. The scatter PSDs presented on this graph were obtained at 5 incidence angle, using two different wavelengths λ r = 632.8 nm and λ b = 488 nm, on both sides of the specular beam. The PSD scatter wavelength scales confirming, that after coating, the Zerodur surface still obeys the Rayleigh-Rice theory. As scatter data starts being believable from approximately f = 10 cycles/mm, it is noticeable that there is always at least one of the PSD scatter curves passing through the PSD surface error bars for f between 10 and 200 cycles/mm. That shows that a profilometric measurement properly predicts the scatter function introduced by surface microroughness for this sample. For frequencies smaller than 10 cycles/mm, the instrument signature of the scatterometer is prevalent over the scatter introduced by the surface topography. However, contrary to the silicon wafer s case, the coated Zerodur is contaminated by pinholes. Their contribution to direct scatter measurement at high angles of incidence (85 ) is significant and overshadows the scatter introduced by surface topography. As such, in terms of PSD, scatter introduced by reflection on this surface looks about three orders of magnitude higher than what it would have been if scatter was only due to surface microroughness (Figure 4.7)- in other words, if all other sources of scatter contribution was negligible. Figure 4.6- Coated Zerodur: PSD 2D from surface (MFT 2.5x objective) with 2σ error bars based on spots measured and scatter measurements (θi= 5, λ r = 632.8 nm and λ b =488 nm) comparison.

10 6 10 4 10 2 Scatter measurement 5 λ=632.8nm Scatter measurement 5 λ=488nm Scatter measurement 85 λ=632.8 nm ATST spec MFT measured coated surface PSD in µm 4 10 0 10-2 10-4 10-6 10-8 10-4 10-3 10-2 10-1 10 0 10 1 f in 1/µm Figure 4.7- Zerodur nickel coated: PSD 2D from surface (MFT 2.5x objective) and scatter measurement (θ i = 5, λ b = 488 nm, θ i = 5, λ r = 632.8nm, θ i = 85, λ r = 632.8 nm) comparison. At high angles of incidence, direct scatter measurements show a significant contribution of particle contamination to scattering (dark red curve): the PSD at small frequencies is about 3 orders of magnitude higher than expected. 5. CONCLUSION This paper presents comparative results obtained from a scatterometer direct BRDF measurement and the BRDF prediction from a surface profile measurement for both a silicon wafer and a Zerodur optical flat (λ/20) coated with nickel. For surfaces meeting the Rayleigh-Rice requirements, this paper experimentally shows that a profilometric measurement of those surfaces allows, through their PSD calculation, accurate prediction of their BRDF over a wide range of directions, including angles very close to the specular direction, while those are usually inaccessible to scatterometers limited by their signature. Therefore, an easy profile measurement of optically smooth surfaces is a satisfying alternative to direct BRDF measurements. It has the advantage of predicting BRDF of almost any surface, regardless to the material, shape or direction to specify. This is of prime interest for measuring BRDF of optics such as the primary mirror of the ATST. However, for surfaces not meeting the Rayleigh-Rice requirements, a direct BRDF measurements remains the most reliable and straightforward approach. As regards the particular example discussed in this paper of a scatter specification very near to specular, we have also illustrated a technique for increasing the angle of reflected scatter relative to specular. At a particular spatial frequency and wavelength, increasing the angle of incidence to near grazing will move the reflected scattered light farther from specular than is the case when using near normal incidence. This moves the scatter signal away from scatter due to the instrument, or away from the instrument signature, increasing the precision of the measurement 6. ACKNOWLEDGEMENTS This work was funded by the Association of Universities for Research in Astronomy, Contract C22026SB.

REFERENCES [1] ATST, Science Goals of the ATST, Project Documentation (2004). [2] Hubbard, Rob, M1 Microroughness and Dust Contamination Project Documentation, Tech. note 0013 Rev A.(2002). [3] http://optiper.com/microfinish-topographer/ [4] Stover, J.C., [Optical Scattering Measurement and Analysis Third Edition], SPIE Press (2012). [5] http://www.4dtechnology.com/home/index2.php [6] Tayabaly K.., Prediction of the BRDF with the MicroFinish Topographer roughness measurements Master Report, College of Optical Sciences, University of Arizona (May 2013) [7] Dittman, M., Grochocki, F. and Youngworth, K., No such a thing as sigma: flowdown and measurement of surface roughness requirement Proc. SPIE, 6291, 62910P (2006).