Sei sulla pagina 1di 139

University of California

Los Angeles

High-Performance Control and Prediction for Adaptive Optics

A dissertation submitted in partial satisfaction of the requirements for the degree Doctor of Philosophy in Mechanical Engineering

by

Jonathan Andrew Tesch

2011

Copyright by Jonathan Andrew Tesch 2011

You never know whos swimming naked until the tide goes out. - Warren Bu ett

iii

Table of Contents

1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.1 1.2 The Adaptive Optics Problem . . . . . . . . . . . . . . . . . . . . Dissertation Outline and Contributions . . . . . . . . . . . . . . .

1 2 6 7 7 12 13 16 16 21 24 27 28 32 36 38 39 45

2 An Adaptive Optics Experiment . . . . . . . . . . . . . . . . . . . 2.1 2.2 2.3 Primary Components . . . . . . . . . . . . . . . . . . . . . . . . . Optical Layout . . . . . . . . . . . . . . . . . . . . . . . . . . . . Computer Control and Simulation . . . . . . . . . . . . . . . . . .

3 Phase Modication using Deformable Mirrors . . . . . . . . . . 3.1 3.2 3.3 Membrane Deformable Mirrors . . . . . . . . . . . . . . . . . . .

Frequency Weighted Control Modes . . . . . . . . . . . . . . . . . Characterization of Long Term Drift . . . . . . . . . . . . . . . .

4 Wavefront Sensing and Reconstruction . . . . . . . . . . . . . . . 4.1 4.2 4.3 Shack-Hartmann Wavefront Sensing . . . . . . . . . . . . . . . . . Southwell Phase Reconstruction . . . . . . . . . . . . . . . . . . . Performance Criteria . . . . . . . . . . . . . . . . . . . . . . . . .

5 Generating Disturbance Wavefronts from Aero-Optical Data . 5.1 5.2 Least squares Estimation of Missing Wavefront Data . . . . . . . Mapping Wavefront Images to DM Actuator Commands . . . . .

iv

6 Experimental Modeling . . . . . . . . . . . . . . . . . . . . . . . . . 6.1 6.2 6.3 6.4 6.5 Linear Model of the Adaptive Optics Problem . . . . . . . . . . . Classical AO Loop and Closed Loop Plant Model . . . . . . . . . Optimal Minimum Variance Wavefront Correction . . . . . . .

48 48 52 55 56 60 62 63 65 68 79 80 81 84

Identication of the Deformable Mirror Poke Matrix . . . . . . . . Determining the Number of Control Channels . . . . . . . . . . . . . . . . . . .

7 Optimal Wavefront Correction with LTI Control 7.1 7.2 7.3 7.4

Formulation as a Disturbance Rejection Problem . . . . . . . . . . Subspace Identication of the Minimum-Variance Predictor . . . . Experimental Results . . . . . . . . . . . . . . . . . . . . . . . . . Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

8 Optimal Wavefront Correction with Adaptive Control . . . . . 8.1 8.2 Formulation as an Adaptive Control Problem . . . . . . . . . . . Experimental Results . . . . . . . . . . . . . . . . . . . . . . . . . 8.2.1 Adaptive Control with Statistically Stationary Wavefront Turbulence . . . . . . . . . . . . . . . . . . . . . . . . . . 8.2.2 8.3

85

Adaptive Control with Non-Stationary Wavefront Turbulence 95

Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107

A Estimating Flow Velocity from Disturbance Models . . . . . . . 109 B Primary Component Specications . . . . . . . . . . . . . . . . . 117 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118

List of Figures
1.1 A typical directed-energy engagement featuring a high-energy laser (HEL), and a lower energy laser beacon to illuminate the target. 2.1 2.2 Schematic of the optical layout. . . . . . . . . . . . . . . . . . . . Photo of the experiment in operation, with the red line illustrating the beam path. Note mirrors M1 and M2 are covered with the yellow and grey cloths. This setup possesses a linear polarizer, not depicted in Fig. 2.1, between lenses L3 and L4 to prevent the laser intensity from saturating the WFS or target camera image. . . . . 2.3 Simulink diagram used to run the experiment. Numerous blocks are Embedded Matlab functions that accesss underlying m-les. . 3.1 (a) Exploded view of the typical membrane deformable mirror components, (b) actuator pattern for the 31 actuator DM used for control and (c) pattern for the 61 actuator DM used for disturbance generation. . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2 Inuence function 1 computed from the rst column of the poke matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3 Wavefront images of the frequency-weighted DM modes based on idealized actuator inuence functions . . . . . . . . . . . . . . . . 3.4 (top) Time series for several modal coe cients with zero command to either DM1 or DM2 over an extended period, (bottom) RMS wavefront error over the same period. . . . . . . . . . . . . . . . . 26 22 20 19 14 11 4 10

vi

4.1

(a) Image from a Shack-Hartmann array corresponding to zero disturbance and control command, and example subaperture image. The + indicates the reference centroid location. (b) Schematic of a Shack-Hartmann wavefront sensor. The grid of lenslets extends into the plane of the page. . . . . . . . . . . . . . . . . . . . . . . 29

4.2

Reference image using mirrors M1 and M2 in lieu of the deformable mirrors, superimposed on the generated grid of AOIs. . . . . . . 31

4.3

Slope measurement and phase reconstruction arrangement for a 44 subaperture grid. Dots symbolize phase reconstruction points, and vertical and horizontal lines represent x and y slope measurement locations. The actual lenslet grid used is 12 12. . . . . . . 33

4.4

(Top) A typical slope vector showing local x and y tilts of the wavefront prole, (bottom) resulting wavefront reconstructed using 4.8. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

5.1

The doughnut-shaped layout of the Notre Dame wavefront data, with regions used to ll the central, circular obscuration. The top and bottom halves of the obscuration are corrected individually. The approximate ow velocity, v , is estimated using an image cor relation analysis similar to that described in Appendix A, and is equal to -0.96 pixels per time step in the horizontal direction, and 0.46 pixels per time step in the vertical direction. . . . . . . . . . 40

5.2

(left) Original AAOL wavefront image containing the central obscuration (right) the same wavefront with the obscuration lled. . 43

5.3

Normalized power spectral densities for pixels in the AAOL data comparing numerous regions in Fig. 5.1 to a pixel p1 S. . . . . . 44

vii

5.4

(Left) Resized wavefront image w to be mapped to the DM, (center) projection of w onto R(L), and (right) measured wavefront after applying the disturbance command cw calculated from (5.9). 47

5.5

Normalized power spectral density for the time series of a single pixel in the original, resized wavefront images, and spectral density for the same pixel in the measured WFS sequence. . . . . . . . . . 47 50

6.1 6.2

Block diagram of the general closed loop experimental system. . . Sensitivity transfer function S(z) of the classical AO loop for a variety of controller gains K. For these plots = 0.95 and d = 1.

The sampling frequency is 16 kHz, the same as in the Notre Dame AAOL experiment. . . . . . . . . . . . . . . . . . . . . . . . . . . 6.3 Block diagram used for optimal control design. G is the closed loop plant containing the classical AO loop. . . . . . . . . . . . . 6.4 6.5 6.6 6.7 7.1 Measured deformable mirror modes . . . . . . . . . . . . . . . . . Measured actuator inuence functions for DM1 . . . . . . . . . . Measured actuator inuence functions for DM2 . . . . . . . . . . Modal power distribution for a typical disturbance sequence. . . . Block diagram of the system used for optimal LTI control design. G is the augmented plant containing the classical AO controller. . 7.2 7.3 20 step moving RMS of the spatial RMS of the residual wavefront. 50 step moving average of the maximum intensity of the target camera image. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.4 Mean target camera intensity image averaged over 1500 frames. . 72 73 63 71 54 58 58 59 60 53

viii

7.5

Scatter plot showing the number of time steps the maximum instantaneous intensity was located in each pixel . . . . . . . . . . . 74

7.6

(Top) temporal RMS values of individual modal time series. (Bottom) temporal RMS values of the modal time series for the optimal LTI and classical controller results, normalized by corresponding open-loop values. . . . . . . . . . . . . . . . . . . . . . . . . . . . 75 76 77 82 84 88

7.7 7.8 8.1 8.2 8.3 8.4

Power spectral density for a selection of modal coe cients. . . . . Time series for a selection of modal coe cients. . . . . . . . . . . Block diagram of the extended plant with the adaptive controller. The lattice lter implementation for an FIR lter of order N . . .

50 step moving RMS of the RMS wavefront error. . . . . . . . . . Mean target camera image over 4000 samples in response to a statistically stationary disturbance sequence. . . . . . . . . . . . .

89 89

8.5 8.6

25 step moving average of the maximum target camera intensity. . Scatter plot showing the number of time steps the maximum instantaneous intensity was located in each pixel . . . . . . . . . . .

90

8.7

Modal power. (Top) Temporal RMS values of each modal time series. (Bottom) Temporal RMS of each modal time series normalized by the corresponding open loop value. . . . . . . . . . . 91 92 93

8.8 8.9

Power spectral densities of select modal sequences . . . . . . . . . Coe cients for a selection of modal control channels . . . . . . . .

8.10 (a) 20 step moving RMS of the RMS wavefront error, (b) Di erence between the RMS wavefront error for the LTI and adaptive controllers, and the ideal benchmark case. . . . . . . . . . . . . . 99

ix

8.11 (a) Mean target camera image for the time period T1 , (b) Mean target camera image for the period T2 . . . . . . . . . . . . . . . . 100 8.12 (a) 25 step moving RMS of the maximum target camera intensity over time period T1 , (b) 25 step moving RMS of the maximum target camera intensity over T2 . . . . . . . . . . . . . . . . . . . . 101 8.13 Scatter plot showing the number of time steps the maximum instantaneous intensity was located in each pixel for during period T1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102 8.14 Scatter plot showing the number of time steps the maximum instantaneous intensity was located in each pixel for during period T2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103 8.15 Power spectral densities of select modal sequences over period T1 . 104 8.16 Power spectral densities of select modal sequences over T2 . . . . . 105 A.1 Graphical example of the calculation of the cross-correlation element c3,5 for images A, B R44 . . . . . . . . . . . . . . . . . . . 111 A.2 Surface plot of CW () constructed from (A.11) with the WaveTrain realization for time delay = 7 . . . . . . . . . . . . . . . . . . . 116 A.3 Location of the maximum of CW () along the j axis versus the time delay. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116

List of Tables
2.1 Imaging components of the experimental system and their e ective focal lengths. PCX lenses are plan-convex, and DCX lenses are biconvex. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.1 5.2 Turbulence parameters for the original AAOL aberration data. . . Parameters for the applied disturbance sequences derived from the AAOL data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.1
2 2 Target camera performance metrics. i and j are the variances of

13 39

44

the instantaneous maximum location in the horizontal and vertical directions. J1 and J2 are the far eld performance metrics dened by (4.11) and (4.12). . . . . . . . . . . . . . . . . . . . . . . . . . 7.2 RMS of the wavefront error over time and space for time steps 200 to 2000. RMS is the mean reduction in the temporal RMS versus open loop over the rst 15 modal coe cients. . . . . . . . . . . . 8.1
2 2 Target camera performance metrics. i and j are the variances of

78

78

the instantaneous maximum location in the horizontal and vertical directions. J1 and J2 are the far eld performance metrics dened by (4.11) and (4.12). . . . . . . . . . . . . . . . . . . . . . . . . . 8.2 RMS of the wavefront error over time and space. RMS is the mean reduction in the temporal RMS versus open loop over the rst 15 modal coe cients. . . . . . . . . . . . . . . . . . . . . . . . . . . 94 94

xi

8.3

Target camera performance metrics with the non-stationary distur2 2 bance sequence. i and j are the variances of the instantaneous

maximum location in the horizontal and vertical directions. J1 and J2 are the far eld performance metrics dened by (4.11) and (4.12).106 8.4 RMS of the wavefront error over time and space. RMS is the mean reduction in the temporal RMS versus open loop over the rst 15 modal coe cients. . . . . . . . . . . . . . . . . . . . . . . . . . . 106

B.1 Shack-Hartmann Wavefront Sensor Specications . . . . . . . . . 117 B.2 Target Camera Specications . . . . . . . . . . . . . . . . . . . . 117 B.3 Membrane Deformable Mirror Specications . . . . . . . . . . . . 117

xii

Acknowledgments

There are many individuals whove help me along the enlightened path of grad school over the last ve years. Their advice, encouragement, and sometimes discouragement, has been a continual source of inspiration. First, many thanks to my advisor, Steve Gibson. His unyielding support and guidance were vital to my success during the last ve years. I am also extremely grateful for the assistance of my committee members. Professors MCloskey, Tsao and Vandenberghe all provided their own perspective and insight to my research. I would also like to thank Professor Eric Jumper and his group at Notre Dame for generously sharing the results of their work on the Airborne AeroOptics Laboratory. Many thanks also, to the numerous professors and instructors during my years at UC Berkeley. It was their lecture halls and classrooms where my interest in engineering was forged. Also, Im extremely grateful to my friends from Westlake, Berkeley and UCLA for keeping me sane, and for showing me the ner points of grad student life. Special thanks to my mom and dad for always being there for me. Their bottomless supply of love, encouragement, and steak dinners were vital ingredients in making this long road a delight.

xiii

Vita

1983 2005-2006

Born, Los Angeles, California, USA Undergraduate Research Assistant, Sohn Research Laboratory, University of California, Berkeley

2006

B.S. Degree, Mechanical Engineering, University of California, Berkeley

2006

Graduate Fellowship, Department of Mechanical and Aerospace Engineering, University of California, Los Angeles

2007

M.S. Degree, Mechanical Engineering, University of California, Los Angeles

2007

Summer Internship, The Aerospace Corporation, El Segundo, California

2007-2008

Graduate Student Instructor, Department of Mechanical and Aerospace Engineering, University of California, Los Angeles

Publications

J. Tesch and J. S. Gibson,Optimal and Adaptive Correction of Aero-Optical Wavefronts in an Adaptive Optics Experiment, in Unconventional Imaging and Wavefront Sensing VII, (San Diego, CA), SPIE, August 2011.

xiv

J. Tesch, S. Gordeyev, S. Gibson, and E. Jumper, Identication and Prediction of Aero-Optical Wavefronts in Laser Beam Propagation, in 42nd Plasma Dynamics and Lasers Conference, (Honolulu, HI), AIAA, June 2011.

xv

Abstract of the Dissertation

High-Performance Control and Prediction for Adaptive Optics


by

Jonathan Andrew Tesch


Doctor of Philosophy in Mechanical Engineering University of California, Los Angeles, 2011 Professor James S. Gibson, Chair

Atmospheric turbulence presents a major obstacle to maintaining the delity of high-energy laser beams over extended distances. Classical control methods, such as proportional-integral controllers, often struggle to mitigate broadband atmospheric aberrations as they do not address the loop latency inherent to adaptive optics systems. Without a form of wavefront prediction, there is no method to increase the bandwidth of these controllers other than increasing the sampling frequency. This dissertation presents two adaptive optics methods to predict and control wavefront aberrations, and extend adaptive optics controller bandwidth without an increased sampling frequency. The rst is an optimal linear time-invariant controller constructed from a state-space model of the turbulence ow. The second control method is an adaptive controller, based on a recursive least squares lattice lter, that tracks the aberrations statistics during operation. Both control methods have the form of multichannel prediction lters that capture the statistics of the aberration turbulence to mitigate latency in the adaptive optics loop.

xvi

The performance of these control schemes versus classical integrator methods is investigated in an adaptive optics experiment that accurately mimics largescale adaptive optics systems. Similar methods have been successfully employed to mitigate long-range atmospheric turbulence with constant power spectrums, while the experiments here examine controller performance against disturbances derived from recent in-ight measurements of aero-optical turbulence. Experimental results show signicant improvement in wavefront correction achieved by both prediction methods versus classical control methods. Altering the ow characteristics of the disturbance wavefront during the control process illustrates the ability of the adaptive controller to track changes in the aberration statistics.

xvii

CHAPTER 1 Introduction
Atmospheric turbulence presents a major obstacle to maintaining the delity of high-energy laser beams over extended distances. Convection currents and pockets of varying density create local disparities in the atmospheres index of refraction, causing portions of an incident beam to traverse di erent optical path lengths. The result is a distorted wavefront and a signicant degradation in the maximum spot intensity at the target. Adaptive optics tackles this phenomenon by altering the wavefront phase prole of an outgoing beam to counteract any subsequent atmospheric distortion. These systems employ a combination of wavefront sensors and phase modication devices, linked by a control system that attempts to minimize the spatial and temporal variance of the wavefront. The wavefront prole is directly related to the optical path length traveled by the beam, thus, minimizing wavefront variance is equivalent to ensuring that all parts of the beam traverse the same optical path length. In the simplest case, wavefront data is fed back to a deformable mirror (DM) which adjusts the laser wavefront to be the conjugate of the measured atmospheric distortion.

1.1

The Adaptive Optics Problem

The wavefront of a beam traveling through a medium is a surface of constant phase over the cross section of the beam path. Light from a coherent illumination source, such as a laser beam, is initially in-phase and hence possesses a planar wavefront prole. However, in a non-vacuum environment, local inhomogeneities in the transmission medium cause portions of the beam to traverse farther optical lengths, which determines the phase of the beam at a later location. For a light ray traversing distance d in a medium with constant refractive index n, the optical path length is dened as OPL = dn. Thus light traveling a given OPL will arrive with the same phase as if it had traveled the equivalent distance in a vacuum. For a beam with an extended cross section, the optical path di erence (OPD) over the cross section is dened as OPD(r) = OPL(r) OPL(r) (1.1)

where r is a vector in the cross section of the beam orthogonal to the direction of propagation, and OPL is the mean optical path length over the cross section. Thus, the wavefront prole at a particular location can be considered a map of the OPD over the beam cross section. For a beam that traverses a uniform physical distance, the wavefront prole is entirely a consequence of refractive inhomogeneities in the transmission medium experienced by di erent parts of the beam. In atmospheric optics, the GladstoneDale relationship provides a direct coupling between the local density and refractive index [9, 40]. The density eld varies spatially and temporally, and because propagation distances are quite large in AO applications, these small variations may result in large spatially and temporally varying wavefront distortions in an

incident beam. The wavefront prole of a beam plays a vital role in directed energy systems since maximum spot intensity of a beam focused on a target is directly related to the degree of phase distortion. The performance of adaptive optics systems is commonly measured by the Strehl ratio, which is the ratio of the axial intensity at the target to the theoretical, di raction-limited intensity that would result with no turbulence [46, 55] . The ratio is in turn a consequence of the optical path di erence; for example when the aperture is much larger than the turbulence structures, the Large Aperture Approximation of the mean Strehl ratio is [40] SR = e
2

kOPDk2

(1.2)

where OPD denotes the spatial RMS of the mean optical path di erence. A typical adaptive optics system is depicted in Fig. 1.1, where the objective is to project a high-energy laser (HEL) through an evolving turbulent medium and onto a target object [22]. Wavefront aberrations are measured by illuminating the target with a low-energy laser beacon which acts as a point source emitting spherical wavefronts. In an ideal case, with no turbulence, the long distance to the target causes the spherical wavefronts emanating from the beacon to be approximately planar when they reach the wavefront sensor. However the wavefront becomes distorted in the presence of turbulence; the AO system uses wavefront sensor measurements of these disturbances to drive a deformable mirror such that the beacon wavefronts are planar after reecting o of the mirror surface. If the HEL travels through same isoplanatic patch1 as the beacon, then imaging the outgoing HEL o of the same DM will pre-compensate the beam for
An isoplanatic patch is a region of atmosphere over which wavefront errors are constant or highly correlated
1

Figure 1.1: A typical directed-energy engagement featuring a high-energy laser (HEL), and a lower energy laser beacon to illuminate the target. the subsequent atmospheric turbulence. Despite the detail in this procedure, Fig. 1.1 may represent any number of adaptive optics applications. The target, for example, may be a laser communications satellite at a distance of several hundred kilometers from the source, a traveling ballistic object tens of kilometers away, or the cross section of a retina in a human eye to be imaged from a few centimeters. The motivation for this dissertation is the recent need for high-performance wavefront compensation in adaptive optics systems. Historically, classical control methods have successfully mitigated the atmospheric turbulence encountered by optical telescopes, and astronomical imaging devices using proportional-integral controllers can achieve performance near their di raction limit in certain cases [47]. However the e cacy of these classical methods su ers when faced with broadband wavefront turbulence, such as when the target and illumination source are on moving platforms, or when the beam is subject to strong aero-optical aberrations. In these situations, the inherent latency between wavefront measurement and control actuation accounts for the majority of AO performance degradation.

Classical AO controllers do not directly address this delay, hence the only way to extend their limited bandwidth is to increase the sampling frequency. This dissertation considers a variety of AO control schemes which utilize wavefront prediction methods to extend the bandwidth of the AO system without increasing the sampling frequency. The fundamental approach in these methods is to exploit the strong temporal and spatial correlations between turbulence wavefronts, and predict oncoming disturbances to mitigate overall loop latency. The rst method is a multichannel, linear time-invariant (LTI) controller derived from an a priori characterization of the turbulence statistics. The controller has the form of a minimum-variance Kalman predictor, but it is not derived via standard Kalman ltering methods with assumed plant and noise models. A related approach has been demonstrated in [14, 15, 16] although several aspects of the controller design and system identication are di erent here, including the integration of optimal control with a classical AO loop, the frequency-weighted deformable mirror modes used here to reduce the number of control channels, and the lattice-lter based subspace system identication algorithm. A second approach is a multichannel adaptive controller that identies an optimal wavefront predictor in real-time from unknown disturbance statistics. In contrast to the LTI controller, a turbulence model is not identied beforehand. Instead, the adaptive lter converges to the optimal predictor iteratively, and can adjust to maintain optimality even if the turbulence statistics are non-stationary. The adaptive ltering algorithm here is a recursive least squares (RLS) lattice lter, which possesses superior numerical stability and e ciency compared to traditional RLS methods [17]. A central component of this dissertation is the design of an adaptive optics experiment mimicking the measurement and compensation of the low-energy

beacon in Fig. 1.1. Consisting of a coherent laser source, a wavefront sensor, target camera, and two deformable mirrors, this hardware simulation serves as a exible test platform for comparing controller performance under realistic disturbance scenarios. One deformable mirror is used to generate stochastic wavefront sequences that reproduce phase aberrations from experimental measurements. These disturbances are corrected by a second deformable mirror, which is adjusted based on measurements from the wavefront sensor.

1.2

Dissertation Outline and Contributions

This dissertation presents results for an adaptive optics experiment used to test wavefront prediction methods under realistic disturbance situations. Recent

adaptive optics research has shown improved wavefront correction using adaptive ltering and control [8, 21, 22, 30, 31, 45] and by optimal linear time-invariant (LTI) ltering and control [15, 16, 39, 38, 52, 53]. Chapters 2 through 4 describe the setup of the adaptive optics experiment in detail, and explain the operation of the primary hardware components. Chapter 5 presents the aero-optical aberration data used as disturbance wavefronts, and describes the process of reproducing the original ight data in the adaptive optics experiment. The mathematical model of the experiment used for controller design is presented in Chapter 6, which also discusses the least squares characterization of the deformable mirrors used for control. Finally, Chapters 7 and 8 detail the design of the optimal control loops, and presents experimental results comparing their performance in response to the aero-optical disturbance wavefronts.

CHAPTER 2 An Adaptive Optics Experiment


An adaptive optics experiment mimicking the low-energy beacon system in Fig. 1.1 is used to test the control algorithms discussed in subsequent chapters. Wavefront disturbances are placed on a continuous wave laser beam using a highresolution deformable mirror, while a second, lower resolution deformable mirror implements control commands. A Shack-Hartmann wavefront sensor provides estimated phase measurements, while an unmodied video camera serves as a target for measuring absolute performance. The high-energy laser in Fig. 1.1 is not involved in the control process, and is thus omitted from the experiment. However in an actual adaptive optics system it would simply reect o of the control mirror before entering the atmosphere.

2.1

Primary Components

It is helpful to describe the fundamental hardware used in the experiment before discussing the general optical layout. Following chapters examine the operation of the primary components in further detail. Laser Source: The source of illumination in the adaptive optics experiment is a 0.8 mW, continuous wave, helium-neon (HeNe) laser with a wavelength of 634 nm. Although any laser of su cient power is suitable, the long co-

herence length of gas lasers greatly simplies optical alignment since the collimation of the beam can be checked at will using a shear plate. Diode lasers also diverge asymmetrically, which can be problematic when attempting to sense and correct large aberrations. Relay telescopes: Simple two lens systems alter the beam diameter as it travels through the experiment. The telescopes used here consist of two positive lenses nominally separated by the sum of their focal lengths. The focal point of the rst lens acts as a point source for the second, and collimated input beams remain collimated upon exiting. Details on lens selection and design of relay telescopes can be found in [24]. Spatial Filter: A spatial lter removes undesired intensity variations in a beam by physically blocking high frequency components. The lter consists of a relay telescope using a microscope objective lens, with the addition of a pinhole placed at the internal focal point. The objective lens focuses the incoming beam to the focal plane, however because of di raction the result is an Airy-like pattern comprised of a central spot and numerous concentric rings [13]. The pinhole is positioned to pass only the central spot, resulting in a near-perfect gaussian beam. Light emerging from the pinhole acts as a point source, and the subsequent diverging beam can be collimated at a desired diameter by a converging lens of appropriate focal length. From a Fourier optics perspective, the spatial lter is a 4f correlator where the 2-dimensional Fourier transform of the incoming beam appears on the intermediate focal plane. The pinhole aperture acts as an amplitude lter that scrubs the spectrum of high frequency content before it is transformed back into the spatial domain by the second lens [10]. Deformable Mirror: A deformable mirror (DM) is a reecting object with a

variable surface geometry. Two are used in this experiment to generate control and disturbance phases. Manipulating the DM control voltages alters the displacement of the mirror surface, changing the optical path length traveled by a particular area of the incoming beam. In this way phase proles an be imparted into the wavefront as it reects o of the mirror surface. The DM surface is concave in its un-actuated state, producing an inherent static wavefront bias that can be corrected with supplemental optics. The control DM in this experiment is a 31 actuator, 25 mm diameter membrane deformable mirror, while the disturbance DM contains 61 actuators. Both are manufactured by Active Optical Systems. Their operation is discussed further in Chapter 3. Wavefront Sensor: A wavefront sensor (WFS) estimates a function of the the phase prole of an incoming beam. As the objective of AO is to correct for phase errors, the WFS provides the primary data that is used to control the DM. This experiment uses a Shack-Hartmann WFS, manufactured by Active Optical Systems, that measures the spatial derivatives of the wavefront phase. The computation of the phase from WFS data is considered in Chapter 4. Target Camera: The target camera is used to image the beam prole or focused spot intensity directly. While data from the WFS is frequently used as a controller input, the ultimate goal of AO is to maximize the intensity of the focused beam. Therefore the target camera can be used to independently rate the performance of a controller design. In this experiment the target is a CCD monochrome camera from Allied Vision Technologies. The processing of deriving intensity information from the camera image is discussed in Chapter 4

DM1 Iris
BS 1

500 mm (L2)

Pinhole

9 mm (L1)
0.4 mm

M1

HeNe Laser = 634 nm Spatial Filter 125 mm (L3)

125 mm (L4)
20 mm

400 mm (L5) 50 mm (L7) 75 mm (L8)


1.8 mm

M2 25 mm (L6)

SHWFS

BS 2

DM 2

Target Camera

Figure 2.1: Schematic of the optical layout.

10

Figure 2.2: Photo of the experiment in operation, with the red line illustrating the beam path. Note mirrors M1 and M2 are covered with the yellow and grey cloths. This setup possesses a linear polarizer, not depicted in Fig. 2.1, between lenses L3 and L4 to prevent the laser intensity from saturating the WFS or target camera image.

11

2.2

Optical Layout

As illustrated in Fig. 2.1, the initial 0.4 mm beam from the HeNe laser source immediately enters a spatial lter, where it passes through a 9 mm objective lens (L1). A 25 m pinhole aligned at the focal point of the lens lters out the secondary intensity rings around the central disk. The diverging beam then expands to approximately 30 mm in diameter before being collimated with L2, a 500 mm lens placed 500 mm from the pinhole. An iris pares this beam to the central 25 mm, the diameter of the DM reecting surfaces. Mirrors M1 and M2 are used for sensor calibration only, thus in normal operation the beam is directed by a non-polarizing cube beam splitter (BS1) to the control DM (DM1). Here, the phase of the beam is altered by the deformed reecting surface of the mirror. Light reected o the control DM travels through a relay telescope with unit magnication (lenses L3 and L4), imaging the beam through BS2 and onto the disturbance DM (DM2), which alters the phase once again. BS2 directs the light reected from the disturbance DM to a second reimaging telescope (lenses L5 and L6) with 0.0625 magnication, reducing the beam diameter from 25 mm to approximately 1.6 mm. Here a nal beam splitter (BS3) cleaves the beam into two branches. One branch contains a single 75 mm lens (L9), focusing the beam to a point at the target camera sensor plane. The second branch passes through a third relay telescope (bi-convex lenses L7 and L8) with unit magnication, imaging the beam onto the WFS lenslet array. Polarizing lters placed strategically throughout the optical path reduce the beam intensity, preventing WFS and target camera saturation without altering the wavefront.

12

Table 2.1: Imaging components of the experimental system and their e ective focal lengths. PCX lenses are plan-convex, and DCX lenses are bi-convex. Lens L1 L2 L3 L4 L5 L6 L7 L8 EFL (mm) 500 120 120 400 25 50 75 75 Type PCX PCX PCX PCX PCX DCX DCX DCX

2.3

Computer Control and Simulation

Matlab was the programming environment used to manage and run the experiment. A compilation of independent M-le scrips provided digital control of the deformable mirrors, wavefront sensor and target camera, and also performed the computation required to compute control commands. Controllers themselves were implemented in Simulink, which accesses the underlying m-les through extensive use of Embedded Matlab function blocks. The recursive least squares lattice lter described in Chapter 8 was written in C, and interfaced with Simulink through an S-function block. Figure 2.3 shows the Simulink block diagram used to run the experiment. All computational aspects of the experiment were performed during run-time on a Dell PC with a dual-core, 3.6 GHz Pentium 4 processor running 64-bit Windows 7. The wavefront sensor and target cameras were connected by a IEEE 1394 Firewire connection, while the deformable mirror drivers received digital commands from either a USB interface (DM1), or thorough an NI PCIe card (DM2). The experiment operated at a maximum closed-loop frequency of approximately

13

[T,d]

120

d y

C(z)

K*u Modes

K*u pinv(G)

MATLAB Function MATLAB Fcn Scope

DMs

SHWFS

n q getSR

E* u
Out1 In1

q_u TstartOL M K*u

Recon

K*u

SR

sr

Out1

In1

rhat

Tstart

Optimal Controller

1 z Unit Delay1

G(z)

Figure 2.3: Simulink diagram used to run the experiment. Numerous blocks are Embedded Matlab functions that accesss underlying m-les. 45 Hz, primarily limited by the WFS frame rate, and the computation time required to compute the WFS slope vector in the Matlab programming language. The experiment is is not meant to be a real-time implementation; the sampling rate is neither consistent, nor adequate for the high-performance adaptive optics engagements considered here. Instead, the experiment should be considered a hardware simulation, where results are analyzed using an assumed sampling rate determined by the rate at which the original disturbance wavefront data was collected. For example the aero-optics data discussed in Chapter 5 was collected at a sampling rate of 16 kHz. Each disturbance frame is placed sequentially on the disturbance DM, therefore control design and analysis for that disturbance sequence assumes that sampling rate. Simulink is also used to model the experiment with a reasonably accurate simulation. The Simulink diagram for the simulation is identical to that in Fig.

14

2.3, but with the DM and WFS blocks replaced with the deformable mirror poke matrices, discussed in Chapter 3. Because the experiment operation speed is limited, the simulation allows for quick tuning and investigation of control parameters.

15

CHAPTER 3 Phase Modication using Deformable Mirrors


Deformable mirrors (DM) are the primary sources of phase aberration and correction in the AO experiment. This chapter details their characteristics and how they alter the phase prole of an incident beam in response to control commands. A set of spatially orthogonal, frequency weighted basis modes are used to decompose the mirror surface. As the mirrors are the primary actuators in the experiment, the projections of the wavefront onto these basis functions serve as convenient control channels Section 3.1 of this chapter describes the operation of the deformable mirrors, and their integration into the adaptive optics experiment. The procedure for constructing the frequency weighted modes is presented in Sec. 3.2. The chapter concludes with a brief discussion about the unmodeled behavior of the deformable mirror and their sensitivity to environmental factors.

3.1

Membrane Deformable Mirrors

Deformable mirrors are active optical components that alter the phase of an incoming beam in response to a vector of control voltages. As shown in Fig. 3.1(a), a DM consists of a silvered membrane reecting surface suspended above a grid of electrostatic actuators [28]. These actuators pull on the DM surface in response to an applied voltage, altering the surface prole, and in turn, the

16

optical path length of an incident beam. The displacement of the mirror surface above an actuator, di , is proportional to the square of the actuator voltage [26, 57]. Thus
2 d i = ai v i

(3.1)

where vi is the voltage applied to actuator i and ai is a scalar constant that depends on the actuators proximity to the clamped mirror boundary. The drive electronics for the DMs restrict the actuator voltages to integers in the range vi [0, 255]. The control systems used in the experiment are designed assuming commands that are linear with respect to the surface displacement, hence the relationship between the ith actuator voltage and its corresponding control command is dened by q 2 vi = round ci + vb

(3.2)

where vb is a bias voltage independent of i. As the electrostatic actuators are only capable of pulling on the mirror surface, using an o set voltage facilitates positive and negative control commands. The bias is chosen to be at the midpoint of the displacement range, and therefore the square of the voltage range, so that p vb = 180 2552 /2. Because of the nite precision of vi , control commands

2 2 are limited to squares in the range ci [ vb , vb ]. Ignoring quantization, (3.2)

provides an a ne relationship between control commands and surface displacements:


2 di = ai (ci + vb )

(3.3)

The uniform bias voltage and clamped surface boundary results in a concave mirror prole when ci = 0, i, imparting a focus wavefront onto the beam. The bias shape can be reduced, though not completely eliminated, by adjusting the

17

spacing between lenses in subsequent beam expanders, particularly L4 in Fig. 2.1. This ensures that a beam incident on the DM retains close to its original phase if ci = 0 for all actuators. The surfaces of the deformable mirrors used in this experiment are continuous membranes, hence applying a command to a single actuator alters the entire surface prole beyond the actuators location. The change in DM shape in response to a unit control command at actuator i is called the inuence function for that actuator, i , an example of which is shown in Fig. 3.2 for i = 1. For the limited surface displacements allowed by voltages in the range vi [0, 255], the total change in the DM surface is a linear combination of inuence functions =
na X i=1 c

ci i

(3.4)

where na is the number of actuators and the bias shape. Consequently, beam by the control action.
c

is the deviation in the DM surface from

is the wavefront prole added to an incident

In practice, the actuator inuence functions in (3.4) are spatially sampled on a square grid with ns elements, and arranged as the columns of a matrix R2ns na called the poke matrix of the deformable mirror = [g1 g2 . . . gn ] . (3.5)

Under this representation, each actuator inuence function gk R2ns is expressed as a discrete slope vector measurement of the underlying continuous wavefront prole k . Slope vectors are the output of the wavefront sensor, and discussed in detail in Chapter 4. The matrix allows for a discretized version of (3.4)

that linearly maps control commands directly to the measured wavefront sensor

18

Spacer

Polymer membrame and frame Electrostatic pads

Electrical connector

(a)

(b)

(c)

Figure 3.1: (a) Exploded view of the typical membrane deformable mirror components, (b) actuator pattern for the 31 actuator DM used for control and (c) pattern for the 61 actuator DM used for disturbance generation.

19

Figure 3.2: Inuence function 1 computed from the rst column of the poke matrix .

output yc = c where yc is the slope vector measurement of


c

(3.6) and control vector c contains the

individual actuator commands. Typically, 2ns > na , and the localized nature of the actuator inuence functions ensures that has full column rank if the optical

system is properly aligned. If saturation and quantization are ignored, the range space of spans the space of wavefront aberrations that can be corrected by the

DM, as measured by the wavefront sensor. Figures 3.1(b) and 3.1(c) show the hexagonal layout for the 31 and 61 actuator deformable mirrors (DM1 and 2 respectively) used to correct and generate phase aberrations in the experiment. DM2 operates using an independent set of disturbance voltage commands. Both DMs have a maximum actuator throw of 10m.

20

3.2

Frequency Weighted Control Modes

The control systems considered here operate simultaneously on a set of parallel control channels. In this way the spatially distributed control problem of minimizing the total wavefront phase can be decomposed into a set of uncoupled control objectives for each channel. This is accomplished by projecting the phase measurements onto a set of modal basis vectors that parameterize the space of actuator commands. In light of (3.6), the actuator inuence functions are an obvious possibility, however they are localized and generally not orthogonal. Past adaptive optics e orts commonly use Zernike polynomials, or modes derived from the singular value decomposition of the poke matrix [7, 37], however these options are either not complete on a discrete domain, or lack the frequency organization described here. Karhunen-Loeve functions have been shown to be optimal from a wavefront representation perspective [19], but such modes are unique to the particular statistics of the wavefront disturbance sequence. By solving a certain eigenvalue problem one can create set of basis modes that are ordered by increasing spatial frequency, and orthogonal with respect to the actuator inuence functions [22, 31]. Ordering by spatial frequency is useful since the control actuators are subject to saturation constraints; compensating only lower-order modes, or only those that possess high disturbance power, is a direct method to implement spatial ltering [7]. For a DM with na actuators, we can construct na such modes, represented as a set of actuator commands, and store them as columns in a matrix V Rna na . While the actuator poke matrix in (3.6) maps actuator commands to WFS measurements, the modal poke matrix U = V maps a modal command

21

vector, q, to the slope vector measurement of the DM wavefront such that yc = V q = U q (3.7)

Figure 3.3 shows the 31 modes derived from the 31 actuator deformable mirror, and used as input and output channels for control design. The images in Fig. 3.3 represent the shape of the deformable mirror surface due to each mode, and were constructed by multiplying each column of V by theoretical, high-resolution actuator inuence functions.
1 2 3 4 5 6 7 8

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

Figure 3.3: Wavefront images of the frequency-weighted DM modes based on idealized actuator inuence functions

22

Computing Frequency-Weighted Modes Construction of the modes requires only knowledge of the DM geometry and actuator inuence functions. Following the method presented in [22] and [31], frequency weighting is accomplished by passing the modes through a two-dimensional low-pass lter and sequentially selecting those modes that have the maximum gain. For a continuous membrane DM such as those used in this experiment, the actuator inuence functions themselves are a form of low-pass lter that disperse the e ect of an impulse to a single actuator. Consider a matrix L whose na columns contain the vectorized inuence functions, and V = [1 2 . . . N ] where i is the ith control mode to be identied. Often L contains theoretical high-resolution inuence functions, and generally L is not the poke matrix containing slope vectors. The DM wavefront prole that results from applying the ith mode is Li , and since we require the modes be orthogonal in phase space, V must satisfy (LV )T LV = D (3.8)

where D is a diagonal matrix. The rst mode, 1 , is dened to be a constant command to all the actuators. This is convenient since the bias command is then equivalent to a constant in the rst modal channel only. Modes two through na are found by rst constructing a full-rank matrix B of size na na 1 whose independent columns are orthogonal to the rst mode in

the phase space, i.e. B satises (BL)T Li = 0 (3.9)

23

For any vector x, the command LBx is therefore orthogonal to L1 . The remaining modes can be found by considering the eigenvalue problem of nding X and such that B T LT LBX = B T BX (3.10)

where > 0 is a diagonal matrix of eigenvalues and X contains the corresponding eigenvectors. When X is sorted so that is in descending order, the remaining na 1 modes are given by BX . V = [1 . BX] . (3.11)

The columns of V are then scaled so their maximum absolute value is one.

3.3

Characterization of Long Term Drift

As shown by (3.2), an actuator command of zero corresponds to a nonzero bias voltage, and thus a non-planar bias wavefront prole on the DM. Even when the actuator commands are held constant, this bias prole is sensitive to environmental factors, especially temperature [25]. Changes to the DM1 bias signicantly modies the poke matrices and U , altering the orthogonality of the control

channels and appearing as plant modeling error. Long term drift of the bias DM shape can signicantly a ect the performance of the AO system due to the integral action of the classical controller. If the bias uctuates such that the poke matrix no longer reects the actual mirror behavior, it is possible that part of the wavefront projection onto the range space of the poke matrix is una ected by the actuators. This can quickly lead to command

24

saturation if the controller gain and pole are insu ciently small. Figure 3.4 is a lucid illustration of the drift behavior. The top plot in shows the time series for several modal coe cients over a ve hour period, during which both deformable mirrors were maintained at their bias voltages. Since even higher modes experienced signicant drift, the mirror bias change was not limited to low spatial frequencies. Modes 2 and 3 are analogous to tilt and tip, indicating that the average x and y slope vector measurements changed as well. The bottom plot in Fig. 3.4 shows a time series of the spatial RMS of the measured wavefront. The maximum RMS value approached 1 m, approximately 10% of the deformable mirrors maximum throw, after several hours. These results highlight the di culties inherent in active optical systems. Any realistic adaptive optics controller must possess strong robustness properties, and be able to operate with a moderate degree of dynamic plant uncertainty. The control loops in the experiment here were often driven to saturation over long run-times, possibly because of the drift behavior. Mitigating this e ect required frequent re-identication of the deformable mirror poke matrix, and strict control of the ambient room temperature.

25

x 10 1 0.5 0 0.5 1 0

Mode coefficient

Mode1 50

Mode2 100

Mode3 150 Time (min)

Mode5 200

Mode10 250

Mode30 300

1.2 1 RMS (m) 0.8 0.6 0.4 0.2 0

50

100

150 Time (min)

200

250

300

Figure 3.4: (top) Time series for several modal coe cients with zero command to either DM1 or DM2 over an extended period, (bottom) RMS wavefront error over the same period.

26

CHAPTER 4 Wavefront Sensing and Reconstruction


While the adaptive optics experiment contains a target camera representing a fareld intensity pattern, a dedicated wavefront sensor is the primary measurement device for identication and control. Phase sensing is a well developed eld, and there many wavefront measurement technologies with a wide spectrum of bandwidth and noise characteristics. The device used in the adaptive optics experiment here is a Shack-Hartmann wavefront sensor (SHWFS), which samples the beam cross section at distinct measurement locations instead of providing a continuous interference pattern. Reconstructing the wavefront from the output of a SHWFS involves a straightforward image centroid calculation, and if necessary, multiplication by a pre-dened pseudo-inverse. The multichannel control and identication methods discussed in this dissertation are largely independent of the particular sensing method employed, however. Other than specic noise issues, the results presented in subsequent chapters are compatible with any wavefront sensing method. Section 4.1 of this chapter presents an overview of the wavefront sensing calculations used in the experiment, and how optical alignment and WFS characteristics e ect the sensor output. A least squares method for reconstructing the wavefront prole from the WFS measurements is discussed in Sec. 4.2. Section 4.3 discusses various performance metrics used to score various control algorithms.

27

4.1

Shack-Hartmann Wavefront Sensing

A Shack-Hartmann wavefront sensor is composed of a planar grid of identical micro-lenses (lenslets), with an imaging device located at the focal plane. As depicted in Fig. 4.1(b), each the lenslet spatially samples the incident beam, forming focused spot within a particular subaperture area on the imaging sensor. If the average phase imaged by a particular lenslet is planar relative to the optical axis, the spot is located at the lenslets focal point. However if the local phase has an average tilt relative to the optical axis, the spot shifts by a proportional amount in the x and y-directions. Thus, the displacement of a lenslet spot from the focal point is directly proportional to the spatial derivative of the phase entering that subaperture. The WFS considered here uses a CMOS array capture the image of the lenslet array. The array can operate in a partial frame mode, where only a rectangular selection of the full 1280x1020 pixel frame is captured. The e ective frame rate of the camera scales with the area of the frame, thus high frame rates can be achieved by using a limited number of subapertures contained in a reduced frame. The frame rate can be further boosted by skipping every other CMOS pixel, providing an image that captures the same physical area, and number of lenslets, at half the resolution. These trade-o s between spatial and temporal sampling are acceptable given the nite actuator resolution of the DMs in the experiment. A typical Shack-Hartmann image is shown in Fig. 4.1(a), and is composed of a grid of spots, each formed by an individual lenslet. The WFS image is subdivided into a regular grid of subapertures or areas of interest (AOI), each representing approximately the area that circumscribes the corresponding lenslet. Since the lateral displacement of each spot from the optical axis indicates the local phase tilt entering that lenselet, reconstructing an estimate of the spatial derivative

28

(a)
Incident Beam f Lenslet CMOS Array

Subaperture optical axis

sx

Wavefront

Lenslet spot

(b)

Figure 4.1: (a) Image from a Shack-Hartmann array corresponding to zero disturbance and control command, and example subaperture image. The + indicates the reference centroid location. (b) Schematic of a Shack-Hartmann wavefront sensor. The grid of lenslets extends into the plane of the page.

29

of the entire wavefront entails tracking the location of each point within their respective subapertures. First, each subaperture is isolated based on a list of pre-dened AOI coordinates. The spot generally has a Gaussian-like intensity over each subaperture image, and the pixel corresponding to the centroid of image intensity is taken to be the spot location. Let I(i, j) represent the intensity of pixel (i, j) in a subaperture image. The centroid location (x, y) is then
max 1XX x = I(i, j)(x(i, j) I

jmax

xI ) yI )

(4.1)

i=0 j=0

1 y = I

imax jmax XX i=0 j=0

I(i, j)(y(i, j)

where x(i, j) and y(i, j) are x and y displacements, I(i, j) is the pixel value, P I(i, j) is the total intensity in the subaperture image. (I , yI ) is the and I = x wavefront, nominally located along the optical axis of the lenslet. The x and y

coordinate of a reference centroid for subaperture I corresponding to a planar

spatial derivatives of the wavefront over the subaperture is proportional to the centroid location [41] sx = 2 x f sy = 2 y f (4.2)

where is the length of each square pixel, and f is the lenslet focal length. Processing the entire WFS image involves applying (4.1), and (4.2) to each subaperture image. Note that (4.1) is linear, other than the division by the total intensity. If na lenslets are used, the result is a slope vector of length 2na R2na . containing the information for each subaperture y = [sx sx . . . sx a sy . . . sy a ]T 1 1 2 n n

30

Generating a Reference Wavefront

Figure 4.2: Reference image using mirrors M1 and M2 in lieu of the deformable mirrors, superimposed on the generated grid of AOIs.

Slope measurements are made relative to a set of reference slopes that approximate those of a planar wavefront. Theoretically, a plane wave should result in a subaperture spot at the intersection of the image plane and the optical axis of the lenslet, as illustrated in Fig. 4.1(b). In the experiment however, a nearly collimated beam is used to generate an actual reference slope vector, and provide a basis for constructing the AOI grid shown in in Fig. 4.2. As noted in Chapter 3, the deformable mirrors operate around a nonzero voltage, resulting in a curved wavefront bias when the control command vector is zero. Adjusting lens L4 in Fig. 2.1 can only partially compensate for this aberration, so both DMs are bypassed to generate the subaperture grid. In their place, at mirrors M1 and M2 are uncovered, and reect the beam through an otherwise identical optical path. The resulting WFS image is shown in Fig. 4.3. Software from AOS is used to generate a set of grid coordinates such that each focal point is approximately

31

centered in an AOI. To generate a reference wavefront M1 and M2 are blocked, both DMs are uncovered and set to their bias voltages, and the position of L4 is adjusted such that the lenslet spots are again centered in each subaperture, resulting in the image shown in Fig. 4.1(a). Finally, (4.1) is applied to each AOI with (I , yI ) = (0, 0) to determine the centroids for the reference image. x In the experiments that follow, the WFS utilized a 12 12 lenslet grid, and either 222 or 112 pixels per subaperture depending on the camera subsampling mode used. The dynamic range, of a WFS refers to the maximum tilt that

can be detected by a lenslet before the spot moves beyond the subaperture image. This depends on both the subaperture size and the spacing between the lenslet array and imaging sensor [41] = (r/2 f /2)2 (4.3) is diameter

where r is the length of the subaperture on the WFS CMOS array,

of the focused lenslet spot, and f is the focal length of the lenslet. The size of a spot at the focal plane of a lens is approximately = (2f )/r, and using the

values in Table B.1 we nd that the WFS has a dynamic range of = 0.44m per lenslet if pixel subsampling is not used.

4.2

Southwell Phase Reconstruction

The slope vector itself is used as a representation of the wavefront for control purposes, however it is sometimes necessary to analyze the actual phase prole, rather than slope measurements. Following the method presented in [50], this is accomplished using a least-squares reconstructor which estimates of the phase at the slope measurement locations. The reconstructor takes the form of

32

a pre-computed pseudo-inverse, such that reconstructing the wavefront involves multiplying the WFS measurements by the appropriate matrix. This method is similar to the fast identication method discussed in [29].

Figure 4.3: Slope measurement and phase reconstruction arrangement for a 4 4 subaperture grid. Dots symbolize phase reconstruction points, and vertical and horizontal lines represent x and y slope measurement locations. The actual lenslet grid used is 12 12. Consider the sensor geometry depicted in Fig. 4.3 where the phase reconstruction points are coincident with the slope measurements. Let the vertical and horizontal slopes at subaperture (i, j) be denoted (sx , sy ). The phase p in ij ij the interval between grid points is t to a quadratic polynomial p = a0 + a1 x + a2 x2 (4.4)

where x is a distance in the interval between grid points. Di erentiating creates a linear slope model sx = a1 + 2a2 x (4.5)

Consulting the geometry in Fig. 4.3, slope measurements from the wavefront sensor exist at each end of an interval, thus if the current subaperture is (i, j), then we can set a1 = sx . If each grid point is separated by distance h, then ij from (4.5) we have: a2 = (sx i+1,j sx )/(2h). With these constants known, we ij

33

can assign a0 = pij and manipulate (4.4) to yield a relationship for the phase at adjacent grid points 1 x (s 2 i+1,j 1 y (s 2 i,j+1 1 (pi+1,j h 1 sy ) = (pi,j+1 ij h sx ) = ij pij ) pij ) i = 1 . . . na i = 1 . . . na , 1, j = 1 . . . na 1 (4.6)

j = 1 . . . na

(the relationship in the y-direction is derived in the same manner). Equation (4.6) can be expressed as a single linear equation Dy = H , where y is a slope vector measurement of the underlying phase , and of is the estimate

reconstructed on the slope measurement grid. D is a sparse matrix that

performs the slope measurement averaging on the left-hand side of , and H is a sparse, rectangular matrix. Because a constant phase can be added to the wavefront without modifying the slope vector, H is rank decient. This can be addressed by forcing the phase to be zero mean by adding a row of ones to H, and zero to the corresponding element of y 2 H 1 3 5 2 D 0 0 1 3 5 2 y 0 3 5

H=4

D=4

y=4

(4.7)

With this modication, the phase estimate is given by the pseudoinverse [20] y = (H T H) 1 H T D = . y (4.8)

where = (H T H) 1 H T D is called the phase reconstructor. For simplicity, future references to the reconstructor use the notation = y to indicate (4.8).

The top plot in Fig. 4.4 shows a typical slope vector where the rst 144 elements represent wavefront tilts in the x-axis, and the following 144 represent

34

y-axis tilts. The bottom image in Fig. 4.4 is the corresponding phase surface reconstructed using (4.8).
6 4 2 slope 0 2 4 6 x slopes 8 0 50 100 150 200 y slopes 250

4 OPD (m) 2 0 2 4 12 10 8 6 4 y 2 2 4 x 6 8 10 12

Figure 4.4: (Top) A typical slope vector showing local x and y tilts of the wavefront prole, (bottom) resulting wavefront reconstructed using 4.8.

35

4.3

Performance Criteria

Numerous performance measures are used to gauge the temporal and spatial behavior of the adaptive optics system. The two sensors employed in the experiment, the wavefront sensor and target camera, both provide measures of the incident wavefront. However, as only wavefront sensor measurements are fed back for control, the target camera provides an independent measure of performance that mimics the primary AO objective of maximizing the far-eld intensity pattern.

Wavefront sensor performance measures For the WFS, a primary performance metric is the root mean square (RMS) value of the measured wavefront (the wavefront error). For a particular WFS measurement, the instantaneous spatial RMS may be calculated as the normalized 2 -norm of the reconstructed phase (t) = (t) na
2

(4.9)

where na is the number of subapertures (and number of points in the phase reconstruction grid), (t) = y(t) is the reconstructed phase dened by (4.8). denotes the RMS value.

With a subscript, the symbol

For a sequence of N reconstructed wavefronts combined as the columns of a matrix P = [ (t) (2) (N )], we can dene the RMS over time and space

as the normalized Frobenius norm " N X 1 P = N na k=1 #1/2 P F = N na

(k)

2 2

(4.10)

36

where

denotes the Frobenius norm. The use of (4.9) or (4.10) is clear from

context. Some results that follow use a moving RMS, which is equivalent to (4.10), where each value is centered in a window of size N . Note that under these denitions, minimizing the RMS value of a parameter is equivalent to minimizing the 2 or Frobenius norm.

Target camera performance measures The target camera in the experiment represents a far-eld intensity image, and maximizing the beam intensity on the target is the fundamental adaptive optics objective. However because it is not directly used for control, the target camera provides a convenient, independent measure of gross controller performance. Two performance metrics derived from a time series of target camera images rate the performance of the AO system. The J1 measure of performance is the maximum of the mean target image. For N target images, if I(t) is the target camera image recorded at time t, with image coordinates (i, j), then this is equivalent to
N 1 X J1 = max I(t) . i,j N t=1

(4.11)

The J2 performance measure is the mean instantaneous maximum N 1 X J2 = max I(t) . i,j N t=1

(4.12)

37

CHAPTER 5 Generating Disturbance Wavefronts from Aero-Optical Data


The adaptive optics experiment uses a 61 actuator deformable mirror to generate wavefront turbulence and impart phase disturbances onto the beam. The wavefront data originates from in-ight measurements of aero-optical aberrations. Several processing steps are required to translate the raw measurements to deformable mirror commands in a way that preserves the statistical characteristics. This chapter details those steps, and also the process of computing actuator commands to produce the desired wavefront on the DM surface. Section 5 of this chapter discusses aero-optical wavefront data that are ultimately used as disturbances in the adaptive optics experiment, and the image processing necessary to make them compatible with the experiments deformable mirror geometry. Mapping the processed wavefronts into deformable mirror commands is covered in Section 5.2. The wavefront data used in the adaptive optics experiment originated from the Airborne Aero-Optics Laboratory (AAOL), run by the University of Notre Dame [40]. The original data were gathered from measurements of in-ight aberrations over a at-windowed turret, during which a continuous wave laser was transmitted between two planes ying in a constant formation at an altitude of 4570 m. Internal track loops coupled to fast-steering mirrors minimized the optical jitter

38

in the wavefront measurements. Unlike long-range atmospheric turbulence, aero-optical aberrations result from density uctuations in the shear layer owing around an object. As a result, such aberrations possess statistics that di er signicantly from those experienced in astronomical imaging problems, which typically follow a Kolmogorov spectral distribution [46]. The airplanes in the AAOL experiment were separated by approximately 50 m to ensure aero-optical turbulence was the primary source of aberrations. Table 5.1: Turbulence parameters for the original AAOL aberration data. Turret Azimuthal angle 119 Turret Elevation angle 57 Freestream Mach 0.36 Altitude 4570 m Target Distance 50 m Aperture Size 10.1 cm Sampling rate 16 kHz

5.1

Least squares Estimation of Missing Wavefront Data

The wavefront images in the AAOL data contain a signicant obscuration in the center of the frame where phase data is absent, a trait common in many optical systems. As the adaptive optics experiment considered here does not utilize this geometry, we wish to somehow interpolate the missing wavefront data before using it as a disturbance source. Ideally, this interpolation should preserve the rst and second-order statistics of the disturbance, as well as mimic the overall ow velocity and turbulence characteristics. Since the aberrations originate from in-ight air turbulence over a windowed

39

Figure 5.1: The doughnut-shaped layout of the Notre Dame wavefront data, with regions used to ll the central, circular obscuration. The top and bottom halves of the obscuration are corrected individually. The approximate ow velocity, v , is estimated using an image correlation analysis similar to that described in Appendix A, and is equal to -0.96 pixels per time step in the horizontal direction, and 0.46 pixels per time step in the vertical direction.

turret, there is a strong underlying velocity eld driving the ow. A reasonable approach to lling the obscuration is to infer the missing phase information from neighboring aberrations in the direction of the ow velocity. This lends itself to a minimum-variance estimation formation, where the obscured phase is expressed as a linear combination of the past and future wavefronts in the adjacent regions. Consider the problem of approximating the set of nS obscured pixels in region S of Fig. 5.1 from a time series of nR pixels in regions R1 and R2 . Let y(t) RnS and u(t) RnR be vectors containing all the pixel values in S and R1 and R2 at time t, respectively. Assuming that S is primarily correlated with the past p and future q vectors in R1 and R2 , we can estimate y(t) from a linear combination

40

of past and future values of u(t)


p X k=1 q X =1

y(t) =

o u(0)

k u(t

k) +

u(t + )

(5.1)

where

k,

RnS nR are multichannel moving-average coe cients. Using past

and future values of u, and not just their instantaneous values, ensures that the estimated region preserves both temporal and spatial continuity. This formulation ignores the face that none of the available wavefront images contain samples of S that can be used to estimate
k

or . However, consider

regions R1 , R2 , and S in Fig, 5.1, which have the same dimensions as R1 , R2 , and S, but occupy a portion of the image that does not have an obscuration. If we assume the wavefront data in these regions have the same spatial and temporal statistics as those in R1 , R2 , and S, we can use aberration information from these regions to estimate the matrix coe cients in (5.1). Suppose y (t) RnS is a vector containing the pixel values in S, and u(t) RnR

is a vector containing the pixel values in and R1 and R2 at t. Wavefront data is present in region S, thus we can express y (t) using (5.1) with the addition of an error term (t)
p X k=1 q X =1

y (t) =

o u(0) +

k u(t

k) +

u(t + ) + (t)

(5.2)

where the matrix coe cients are assumed to be the same as those in (5.1). For a given sequence of wavefront images, the coe cient matrices to minimize the means square value of
k

and are chosen

in (5.2) over time and all pixels from

regions R1 and R2 .. Then, y(t) can be approximated for any t by inserting the estimated coe cients into (5.1), and applying them to the corresponding pixels from R1 and R2 .

41

In Matlab, the matrix coe cients are estimated from a batch calculation using all the available wavefront images. Dene the block matrices Y = [(p) y (p + 1) y y (f q 1)] (5.3) (5.4)

where f is the total number of available wavefront images. Then (5.2) can be extended to all the wavefront images as Y = H + . where is now a matrix of error vectors. The matrix that minimize (5.6) is then

=[ p ... 1 2 . . . q ] p 1 o 3 2 u(0) u(1) u(f q p 1) 7 6 7 6 6 u(1) u(2) u(f q p) 7 7 . 6 H=6 7 . . . .. . . . 7 6 . . . . 5 4 u(p + q) u(p + q + 1) u(f 1)

(5.5)

the least squares solution T = (HH T ) 1 HY T (5.7)

To ensure a smooth boundary at the edge of S, a regularization term can be added that forces pixels on the boundary of S to be close to the adjacent pixels in R1 and R2 , however this has proven to be unnecessary in this case. For the 27 27 wavefronts used here, little benet was witnessed beyond p = q = 5. Figure 5.2 depicts a sample wavefront from the original AAOL data, and the same wavefront with the obscuration lled using (5.1). The estimation preserves the continuity of the turbulence structures in the aberration, and the in-

42

x 10 1.5 1 0.5 0 0.5 1 1.5

x 10 1.5 1 0.5 0 0.5 1 1.5

Figure 5.2: (left) Original AAOL wavefront image containing the central obscuration (right) the same wavefront with the obscuration lled.

terface between S, R1 and R2 is not apparent. Viewing a time series of processed wavefront images reveals that overall ow across the aperture passes through S without any apparent discontinuities. Figure 5.3 shows the power spectral density for pixels from several regions in Fig. 5.1 compared to that of a single pixel in S. While there are clearly di erences in in the temporal statistics between various regions, the pixel from the estimated region shares many of the aggregate features. The overall bandwidth, for example, is largely the same, and the high frequency harmonics at 3300 Hz and 6600 Hz are well represented. No analysis has yet been performed to validate the assumption that the statistics in the regions in Fig. 5.1 are equal, but visual inspection of the results and PSDs such as Fig. 5.3 suggest it is a reasonable approximation for the azimuth and elevation angle here.

43

0 10 20 dB 30 40 50 60 dB p1 p2 S R1

0 10 20 30 40 50 60 p1 p3 S R1

2000

4000 6000 Frequency (Hz)

8000

2000

4000 6000 Frequency (Hz)

8000

0 10 20 dB 30 40 50 60 dB p1 p4 S S

0 10 20 30 40 50 60 p1 p5 S R2

2000

4000 6000 Frequency (Hz)

8000

2000

4000 6000 Frequency (Hz)

8000

Figure 5.3: Normalized power spectral densities for pixels in the AAOL data comparing numerous regions in Fig. 5.1 to a pixel p1 S.

Table 5.2: Parameters for the applied disturbance sequences derived from the AAOL data. ND1 Max peak-valley aberration (m) Scale factor Est. velocity vector 10.1 8 [ 0.96, 0.46] ND2 4.9 5 [0.96, 0.46] ND3 5.2 5 [ 0.46, 0.96]

44

5.2

Mapping Wavefront Images to DM Actuator Commands

Once the wavefront images are processed, we desire a set of disturbance DM commands such that the resulting WFS measurements posses statistics that mimic those of the original aberration data. This can be accomplished by considering the vectorized actuator inuence functions of the disturbance DM (represented in the same units as w, generally not slope vectors), organized into the columns of a matrix L. These may be either theoretical inuence functions, or derived from identication of the disturbance DM poke matrix. Suppose w is a column vector representing a vectorized wavefront image, resized to the same dimension at the inuence functions. The problem of computing the disturbance DM actuator commands cw that reproduce w in a least squares sense is formulated as min (w
cw

Lcw )T M (w

Lcw ) + cT cw w

(5.8)

where is a scale factor converting the native wavefront prole into the same units used in L, and M > 0, 0 are weighting and regularization matrices. Regularization is necessary because the actuators near the clamped DM boundary possess lower-amplitude inuence functions, and are thus prone to saturation in the un-regularized solution. The matrix L has full column rank for a properly aligned DM, and the disturbance command that minimizes (5.8) is given by cw = ( + LT M L) 1 LT M (w) Typically, = diag(r), where (5.9)

> 0 is a small number and r is a vector contain-

45

ing proportionally larger values for actuators likely to saturate. For simplicity, M is often chosen to be diagonal as well, with the diagonal terms chosen largely by trial and error. Figure 5.4 illustrates the results of this approach. The desired wavefront w is a 1212 wavefront image that has been processed using the methods in Section 5.1. The center image in Fig. 5.4 is the projection of w onto R(L), given by LL+ w. The actuator inuence functions serve as low pass spatial lters; combined with the reduced degrees of freedom, the estimated and applied disturbance wavefronts have much lower spatial resolution than w. The rightmost image in Fig. 5.4 is the measured wavefront after applying the disturbance command generated using (5.8). Despite this loss of spatial frequency content, the primary statistical features of the ow are well represented in the measured disturbance sequence. Figure 5.5 shows compares the power spectral density of a pixel in the original wavefront data with the corresponding pixel in wavefront sensor measurement. The spectral densities are quite similar, with both high and low-frequency features in the original ow appearing in the measured sequence. In an actual AO system, compensating for high spatial frequency aberrations can be accomplished using a deformable mirror with high actuator density. The experiment considered here is largely a hardware demonstration of wavefront prediction techniques, thus the temporal statistics of the wavefront disturbances are of primary concern. Table 5.2 contains details on the wavefront sequences used in the experiment. All three were derived from the AAOL experimental data using (5.9), and with varying degrees of rotation. For each case the parameters and M were chosen

heuristically to allow relatively accurate reproduction of the original wavefront, without excessive actuator saturation.

46

Figure 5.4: (Left) Resized wavefront image w to be mapped to the DM, (center) projection of w onto R(L), and (right) measured wavefront after applying the disturbance command cw calculated from (5.9).

10

Resized Measured

Normalized PSD

10

10

1000

2000

3000

4000 Frequency (Hz)

5000

6000

7000

8000

Figure 5.5: Normalized power spectral density for the time series of a single pixel in the original, resized wavefront images, and spectral density for the same pixel in the measured WFS sequence.

47

CHAPTER 6 Experimental Modeling


Although the details of the experimental setup are known, it is di cult to derive an exact analytical expression for the relationship between DM control commands and wavefront sensor measurements. In theory, Fourier optics could be employed to determine such a mapping, but the result would be sensitive to minute alterations to the optical path. Instead, we assume a linear model that is an extension of the relationship given by (3.3). Section 6.1 of this chapter explains the linear model and the signals used for control design. Section 6.2 details the design of a classical adaptive optics controller and an extended plant model that encapsulates the classical feedback loop. Finally, Section 6.3 covers an optimal minimum-variance controller that is used as a performance baseline in subsequent chapters.

6.1

Linear Model of the Adaptive Optics Problem


, is the wavefront error measured by the WFS. Under a

The residual phase,

linear model, this phase is what remains after the DM wavefront is subtracted from the disturbance phase and a constant bias wavefront = + .

48

The bias phase

contains the DM bias and other static disturbances due to


w c

optical misalignment. disturbance DM, and

is the turbulence phase added to the beam by the

represent the phase contribution from the control DM.

As in Chapter 3, the slope vector measurement of the DM wavefront is a linear transformation of the control commands; replacing
c

with (3.6) yields the

relationship between control commands and the output from the wavefront sensor y = yb + yw c (6.1)

where y is the slope vector measurement of the residual phase. The slope vector yw is the contribution of the disturbance phase to y, and is linearly related to the disturbance commands by the poke matrix for the 61 actuator DM, yw =
w cw ,

where cw is computed using (5.9). However,

is not used for control

design or implementation, and not observable in many realistic adaptive optics systems. The vector yb is the slope vector contribution from the bias wavefront
b.

Although signicant nonlinearities exist in the experiment, the system is tuned to remain largely in a linear regime, and the model given by (6.1) is usually su cient for tuning the AO control system [22]. Figure 6.1 is a block diagram of the adaptive optics experiment showing both the classical and optimal control loops where c is the command vector to the 31 actuator control DM. The vector sequence e = Ey represents the projection of the wavefront error onto the deformable mirror modes used for control, and u represents a supplemental control input from the optimal controller. The vector cw represents the sequence of commands to the disturbance DM generated from (5.9), but like
w,

these are not used for control.

The matrix V in Fig. 6.1 contains the orthogonal DM modes discussed in Section 3.2, and maps modal commands into equivalent actuator commands. The

49

Figure 6.1: Block diagram of the general closed loop experimental system. modal reconstructor matrix E projects slope vector measurements onto modal coordinates, and is chosen to be the pseudo-inverse of the modal poke matrix U = V such that E V = EU = I (6.2)

When this holds, an element of u only a ects the corresponding component of e, transforming the adaptive optics system into a parallel set of uncoupled control channels. Typically only a subset of the 31 DM modes are used for control, in
+ which case E is dened as E = Um where Um is a matrix containing the rst m

columns of U . The transfer function D(z) represents the deformable mirror and WFS dynamics. While some adaptive optics systems experience complex mirror or camera behavior [4, 23], the operating frequency of the experiment here is far below the DM resonance frequency [25, 27]. The long term drift discussed in Section 3.3 is interpreted as uncertainty in the poke matrix instead of a dynamic feature of

the mirror. As such, the temporal behavior of the DM is negligible, and D(z) is

50

an integer number of sampling delays D(z) = z


d

(6.3)

d = 1 for the experiment here, although d = 2 or greater in some real-time AO systems. These delays encapsulate the overall latency in the AO loop due to DM and WFS measurement behavior and read out times. The dynamics of the disturbance DM are not explicitly modeled; wavefront turbulence is placed on the 61 actuator DM using the procedure in Section 5, ignoring the temporal behavior of the disturbance DM. C(z) is the classical adaptive optics controller acting on the wavefront projection e, and is discussed in Section 6.2. U (z) is an optimal controller that augments the existing classical loop, and is either an adaptive nite impulse response or xed gain linear time-invariant lter. Subsequent chapters discusses the design and implementation of U (z) in detail. Note that D(z) and C(z) are scalar transfer functions operating on each modal control channel, however U (z) is a multichannel, non-diagonal transfer matrix. Under this linear model, the overall control objective of the AO system is to minimize the spatial and temporal RMS of the residual wavefront over the control interval [8] min X
t

(t)

(6.4)

With the parameterization of the actuator space given by V and the reconstructor dened by E, (6.4) translates into an equivalent control objective of minimizing the RMS of each modal control channel min X
t

e(t)

(6.5)

51

6.2

Classical AO Loop and Closed Loop Plant Model

When the optimal control loop is not closed, u = 0 and the system in Fig. 6.1 reduces to a classical feedback controller comprised of lter C(z). If (6.1) holds, the control command that minimizes the residual phase error is the projection of the disturbance phase onto the range space of the DM at every time t c(t) =
+

[yw (t) + yb ] .

(6.6)

In practice this control method is impractical due to actuator saturation, and the fact that an instantaneous measurement of the disturbance wavefront yw (t) is unavailable. Instead, the classical AO controller is a low pass lter of the form C(z) = K z z . (6.7)

K is the controller gain and typically equal for all channels. C is a pure integrator when = 1, however in practice < 1 to prevent actuator saturation.

Combined with the modal reconstructor E dened in (6.2), the classical AO loop incrementally approximates (6.6) for low-frequency disturbances. Choosing K involves a compromise between maximizing the control bandwidth of each channel at the expense of amplifying high frequency noise. Past authors quantied this balance in terms of the Kolmogorov turbulence spectrum of each modal channel [5]. However here the turbulence does not assume such a prole, and K is chosen by examining the sensitivity transfer function S(z) = e/Eyw for the closed loop system in Fig. 6.1. If the AO loop latency is d integer delays, S(z) = e zd zd 1 = d Eyw z zd 1 + K (6.8)

52

The Bode plot of S for a single delay and various values of K is shown in Fig. 6.2. The plot indicates that a gain of K = 0.3 to approximately K = 0.5 provides an adequate disturbance rejection bandwidth without signicantly amplifying high frequency inputs. Experience with the experiment here has shown that when the DM commands do not saturate, the steady-state performance of supplemental controllers is largely independent of K, as long as closed-loop stability is maintained.
Bode Diagram 10 5 0 5 Magnitude (dB) 10 15 20 25 30 35 40 90 K=0.08 K=0.30 K=0.50 K=1.00 K=1.20

45 Phase (deg) 0 45 10
0

10

Frequency2 (Hz) 10

10

10

Figure 6.2: Sensitivity transfer function S(z) of the classical AO loop for a variety of controller gains K. For these plots = 0.95 and d = 1. The sampling frequency is 16 kHz, the same as in the Notre Dame AAOL experiment.

The e ectiveness of controllers of the form (6.7) is limited because they do not address the latency in the adaptive optics system. Without prediction methods,

53

Figure 6.3: Block diagram used for optimal control design. G is the closed loop plant containing the classical AO loop. the only way to extend the error rejection bandwidth is to increase the sampling rate [53]. This thesis is concerned with constructing an optimal controller U (z) that augments the classical control loop. For the design of this system, C and the subsequent classical control loop are enveloped in plant model G which contains the closed classical AO loop. The transfer function G is thus called theclosed loop plant. While the placement of the summing junction for u and e in Fig. 6.1 indicates that the optimal control command u is immediately ltered by the low-pass integrator C(z), results in the following chapters reveal that the action of the integrator can in fact be inverted, and the exact placement of the summing junction is irrelevant. Consider the wavefront error e in Fig. 6.1 which is the sum of the control and turbulence phases e= DC D u+ Eyw . 1 DC 1 DC (6.9)

The closed loop plant transfer function is dened to be the mapping from u to e with the classical AO loop closed G(z) = DC . 1 DC (6.10)

Similarly, dene the signal w as the portion of the disturbance wavefront left

54

uncompensated by the classical AO loop, expressed in modal coordinates w= Then (6.9) reduces to e = Gu + w (6.12) D E(yw + yb ) . 1 DC (6.11)

resulting in the simplied block diagram in Fig. 6.3. In the adaptive optics system considered here, D(z) = z
1

and the extended plant becomes K z+K (6.13)

G(z) =

6.3

Optimal Minimum Variance Wavefront Correction

If the control strategy given by (6.6) could be implemented, the minimum residual phase would be the projection of the disturbance wavefront onto the orthogonal compliment of the image of the poke matrix R? ( ) smin (t) = (I
+

)[yw (t) + yb ]

(6.14)

This situation can approximated in the experiment by articially maintaining each disturbance wavefront for a period of steps, allowing the classical feedback loop to converge and mitigate any modeling error in . Taking the resulting

steady-state slope vector as the measured residual wavefront corresponding to that disturbance command is analogous to (6.6) if all the control modes are used, and saturation does not occur. Thus, while contrived, this method generates a useful performance benchmark for evaluating a controllers prediction and robustness properties.

55

6.4

Identication of the Deformable Mirror Poke Matrix

Since the wavefront error measured by the WFS is a linear function of the control commands in (6.1), the deformable mirror poke matrix can be determined by applying a sequence of independent actuator commands. Setting the disturbance command to zero such that yw = 0, sending n control commands results in a series of residual slope vectors, {y(1), y(2), . . . y(n)} such that = C where the slope vectors y (t) = y(t) yb are organized as columns of y (n)] (6.15)

= [(1) y (2) y and C = [c(1) c(2)

c(n)] contains the corresponding control commands.

When na subapertures are used in the WFS, R2na n , C R31n and R2na 31 . The estimate of the poke matrix minimizes the least squares objective T CT
T

(6.16)

The applied command sequence is a zero mean, uniformly distributed random sequence of command vectors. For identifying the poke matrix for the 31

actuator deformable mirror, n = 400. For the 61 actuator deformable mirror, the poke matrix
w

was identied using n = 1000. The commands to each actuator

are independent, and the poke matrix estimate that minimizes (6.16) follows from

56

the pseudo-inverse of C T T = (CC T ) 1 CT . (6.17)

Similarly, the modal poke matrix U can be determined by applying a sequence of modal commands, Q = [q(1) q2) ... q(n)], such that (6.18)

U T = (QQT ) 1 QT

Or, when the frequency weighted modes occupy the columns of V , U = V . A challenging aspect of this process is the nonlinear noise characteristics of the Shack-Hartman wavefront sensor. The sensor has a limited dynamic range where the centroid locations are linear functions of the wavefront slope. Carelessly applying random commands to the DM often results in a wavefront with high spatial frequency, causing sudden jumps in the centroid measurements as some lenslet spots exceed their dynamic range and move into adjacent subapertures. This often occurs at peripheral subapertures, thus a weighting term could be added to (6.16) and (6.18) to mitigate the importance of these measurements. An easier approach that has worked well in the experiment, is to limit the actuator commands to perturbations around the bias voltage. An example of the identied poke matrices for the 31 and 61 actuator deformable mirrors are shown in Fig. 6.4 through 6.6. While the columns of the actual poke matrices are expressed as slope vectors, the images here were generated by reconstructing the phase of each column using (4.8). The mode images in Fig. 6.4 compares well with the shapes computed from theoretical actuator inuence functions shown in Fig. 3.3. Also of note are the smaller amplitudes of the inuence functions around the DM boundaries.

57

Mode 1

Mode 2

Mode 3

Mode 4

Mode 5

Mode 6

Mode 7

Mode 8

Mode 9

Mode 10

Mode 11

Mode 12

Mode 13

Mode 14

Mode 15

Mode 16

Mode 17

Mode 18

Mode 19

Mode 20

Mode 21

Mode 22

Mode 23

Mode 24

Mode 25

Mode 26

Mode 27

Mode 28

Mode 29

Mode 30

Mode 31

Figure 6.4: Measured deformable mirror modes

Act. 1

Act. 2

Act. 3

Act. 4

Act. 5

Act. 6

Act. 7

Act. 8

Act. 9

Act. 10

Act. 11

Act. 12

Act. 13

Act. 14

Act. 15

Act. 16

Act. 17

Act. 18

Act. 19

Act. 20

Act. 21

Act. 22

Act. 23

Act. 24

Act. 25

Act. 26

Act. 27

Act. 28

Act. 29

Act. 30

Act. 31

Figure 6.5: Measured actuator inuence functions for DM1

58

Act. 1

Act. 2

Act. 3

Act. 4

Act. 5

Act. 6

Act. 7

Act. 8

Act. 9

Act. 10

Act. 11

Act. 12

Act. 13

Act. 14

Act. 15

Act. 16

Act. 17

Act. 18

Act. 19

Act. 20

Act. 21

Act. 22

Act. 23

Act. 24

Act. 25

Act. 26

Act. 27

Act. 28

Act. 29

Act. 30

Act. 32

Act. 33

Act. 34

Act. 35

Act. 36

Act. 37

Act. 38

Act. 39

Act. 40

Act. 41

Act. 42

Act. 43

Act. 44

Act. 45

Act. 46

Act. 47

Act. 48

Act. 49

Act. 50

Act. 51

Act. 52

Act. 53

Act. 54

Act. 55

Act. 56

Act. 57

Act. 58

Act. 59

Act. 60

Act. 61

Figure 6.6: Measured actuator inuence functions for DM2

59

6.5

Determining the Number of Control Channels


160 140 120 100

RMS

80 60 40 20 0 5 10 15 Mode 20 25 30

Figure 6.7: Modal power distribution for a typical disturbance sequence. Because the deformable mirror commands have a limited range, actuator saturation can severely impact the performance of an adaptive optics control algorithm. Optimal controllers with low wavefront prediction error could perform worse than classical integrator methods if they generate control commands that result in actuator nonlinearities. Past experience with the experiment and Simulink simulation reveals that using numerous high frequency modes tends to increase the likelihood of actuator saturation. Actuators near the clamped deformable mirror boundary possess smaller inuence functions than those near the center, thus mode shapes that have signicant features near the boundary are vulnerable to large amplitude voltage commands. Given the limited actuator range, it is reasonable to concentrate control e ort in spatial frequencies where the majority of the disturbance lies. As the deformable mirror modes are weighted by increasing spatial frequency, projecting

60

the turbulent wavefronts onto the modes provides a convenient basis for measuring power. Figure 6.7 shows the temporal RMS of each modal time series for an open loop wavefront sequence. Because the wavefront prole produced by each mode is not unit norm, the RMS of each time series is scaled by the norm of the corresponding modal wavefront. Figure 6.7 indicates relatively little power exists in modal channels beyond 15. Consequently, for the results in subsequent chapters, only the rst 15 modal channels were used for control.

61

CHAPTER 7 Optimal Wavefront Correction with LTI Control


This chapter details the design and implementation of a linear time-invariant (LTI) controller that augments the classical controller in Fig. 6.1. As discussed in Chapter 6, the bandwidth of the classical controller is limited because it does not compensate for the inherent latency in the AO loop. The main component of the optimal LTI controller is a minimum-variance prediction lter that explicitly predicts future disturbances from past wavefront sensor measurements. These minimum-variance techniques are largely based on H2 optimal control theory, and a related approach has been demonstrated in [15, 16] although several aspects of the controller design and system identication are di erent here. A considerable amount of past research in wavefront prediction has focused on identifying velocities in an assumed frozen ow model of turbulence [1, 6, 43, 44], where phase aberrations are assumed to originate from a linear combination of statistically stationary layers owing with static velocities. While often applicable to long-range atmospheric turbulence [42], the frozen ow model may not be su cient to model aero-optical data, especially at extreme turret viewing angles. The general nature of the LTI predictor considered here may reect any linear turbulence model, and will reproduce the second-order statistics of the ow provided that the measured covariance matrices are positive real sequences [18, 36]. Section 7.1 of this chapter discusses architecture of the optimal controller, and

62

reformulates the adaptive optics problem in a disturbance rejection framework. Section 7.2 illustrates how a mimum-variance wavefront predictor is generated from open-loop data using subspace system identication methods. Results from applying this control method in the adaptive optics experiment are presented in Section 7.3 .

Figure 7.1: Block diagram of the system used for optimal LTI control design. G is the augmented plant containing the classical AO controller.

7.1

Formulation as a Disturbance Rejection Problem

The classical control loop remains closed when the optimal controller operates, and the adaptive optics problem reduces to the block diagram in Fig. 6.3 where the extended plant G contains the classical controller given by (6.7), and maps the optimal control input u to the wavefront sensor output in modal coordinates e. Figure 7.1 presents the structure of the optimal controller, where the control command has the form u = F r. The wavefront error sensor measurement e is the sum of the optimal control command, ltered by G, and the remaining

63

disturbance wavefront e = w + GF r (7.1)

where both e and w are wavefronts represented in modal coordinates. The con troller contains a linear internal model of the plant, G, that is valid when actuator saturation and quantization are ignored, and the identied poke matrix is accurate. Past research in [54] indicates that the structure of the controller in Fig. 7.1 is the optimal linear controller when the internal model is accurate and G = G. When this condition holds, the vector sequence r is an estimate of w, r=e Gu = e Gu = w . (7.3) (7.2)

Using (7.1), the wavefront error sequence becomes e = w + GF w (7.4)

Employing the internal model in this fashion allows the AO objective to be reformulated as a feed-forward disturbance rejection problem, where the optimal control objective is to nd the lter F that minimizes the RMS of the wavefront error measurement given by e = w + GF w . (7.5)

Although the condition E V = I in (6.2) ensures that the control channels in G are uncoupled, F is a multichannel, non-diagonal transfer matrix that uses

64

spatial and temporal information from all channels to generate a control command. Past research has shown some utility in including the plant nonlinearities in G [35], both here and in the adaptive control scheme in Chapter 8, however in this case the parameters of the classical controller C can be adjusted such that the plant avoids actuator saturation and remains approximately linear. The transfer matrix F that minimizes (7.5) is a linear time-invariant lter with innite impulse response. When G has a causal inverse, the optimal lter is clearly F = G 1 , yet this is rarely the case. For an arbitrary G, the optimal control

problem of choosing F to minimize e follows from standard LQR methods [32, 33], and the solution over lters with nite impulse response follows from a standard Wiener-Hopf formulation [12, 34, 48]. However in the adaptive optics problem here, the lack of signicant deformable mirror behavior allows for a more direct approach. Because the mirror dynamics are captured by an integer number of delays, G may be factored as G = z
d

Go , where Go is minimum-phase. This

allows the denition of a lter F such that F = Go 1 F (7.6)

With this, the optimal control problem is recast as a d-step prediction problem for the sequence w e = w+z
d

Fw .

(7.7)

7.2

Subspace Identication of the Minimum-Variance Predictor

The lter F that minimizes (7.7) is a d-step Kalman predictor which minimizes the variance of the prediction error. In the standard approach to Kalman lter-

65

ing, the minimum-variance predictor originates from a chosen plant model and explicit assumptions about the disturbance and noise statistics [51, 56]. However, the method here follows a subspace identication approach, where a stochastic realization of the disturbance process results directly from an estimate of the Kalman lter states, and subsequently the Kalman lter state space equations [18, 36]. The particular identication algorithm used to generate the multichannel disturbance model here is detailed in [3]. Subspace system identication methods typically require the QR factorization of a large Hankel matrix, often containing several thousand samples of noisy data. Rather than perform a batch QR factorization, the method in [3] uses a multichannel, least-squares lattice lter to execute adaptive QR factorization of the underlying data matrices. Computational e ciency and accuracy are important qualities for any identication algorithm used in adaptive optics due to the large number of input and output channels. Identication of a linear model for the disturbance is performed in the experiment by applying a sequence of disturbance wavefronts with only the classical AO loop closed, i.e. with u = 0, resulting in a sample time series of the signal w. This sequence is processed by the identication method in [3], resulting in a model of the disturbance that has state-space form x(t + 1) = Ax(t) + K (t) w(t) = Cx(t) + (t)

W :

(7.8)

where w(t) is the estimate of w(t) based in information up to time t, (t) = w(t) w(t) is the prediction error or innovations sequence, and K is the Kalman

gain. The matrices A, K and C, along with the covariance of , implicitly cap-

66

ture the rst and second-order statistics of the disturbance, hence the wavefront turbulence must be wide-sense stationary over the time period encompassing identication and control. The components of the state x(t) are mathematical constructs, and generally have no relationship to physical quantities, although an estimate of the overall ow velocity may be obtained from the solution to a certain Lyapunov equation involving A, K and C, without generating a realization of the wavefront sequence (see Appendix A). The estimate of w at time t + 1 based on information up to time t results from rearranging (7.8) w(t + 1) = Cx(t + 1) = CAx(t) + CK (t) = CAx(t) + CK[w(t) = C(A Cx(t)] (7.9)

KC)x(t) + CK w(t)

The state update equation can be expressed in closed-loop form by expressing the innovations as (t) = w(t) Cx(t), KC)x(t) + K w(t) (7.10)

x(t + 1) = (A

If (7.2) holds, then r(t) = w(t). Since d = 1 in the AO experiment, the one-step predictor minimizes (7.7), and combines (7.9) and (7.10) to form x(t + t) = (A w(t + 1) = C(A KC)x(t) + Kr(t) KC)x(t) + CKr(t)

F :

(7.11)

67

7.3

Experimental Results

Statistically stationary disturbance wavefronts, generated using the methods in Section 5, were applied in the adaptive optics experiment to compare three control scenarios: Open-loop: c = 0, u = 0. No control command to DM1 Classical AO: c = 0, u = 0. Classical loop only closed. Optimal LTI Control: c = 0, u = 0 for t > 100. Both the classical and optimal control loops are closed with the optimal LTI controller as U (z), after F is identied using 5000 samples of a separate but statistically identical disturbance sequence. The optimal control loop is closed after 100 time steps to allow the classical controller to reach steady-state. Figures 7.2 through 7.8 depict the results from applying the ND1 disturbance sequence to the experiment using 15 modal control channels. During the classical AO and optimal control trials, the classical controller gain and pole were set to K = 0.5 and = 0.95. The prediction lter F was constructed from (7.11)

before the control experiments by applying the rst 5000 wavefronts from ND1, and identifying the system matrices (A, K, C) using the subspace method outlined in Sec. 7.2. Then, ND1 wavefronts 5001 through 7000 were applied for each of the three control scenarios above. Figure 7.2 shows the spatial RMS of the residual wavefront for the three control cases. (A moving RMS of the original time series helps clarify the relative controller performance.) While the classical AO loop addressed some of the openloop error, the optimal LTI predictor further reduced the RMS at the majority of time steps. There are instances where the the classical AO loop performed

68

as well as the optimal controller, however the latter clearly avoided the large spikes in the wavefront error that plagued both the open-loop and classical cases. Table 7.1 summarizes RMS wavefront error over space and time given by (4.10); while the classical AO loop reduced the open-loop RMS error by 22.5%, using the optimal controller resulted in a 46.5% reduction. Because of the maximum far-eld intensity is closely linked to the Strehl ratio, mitigating the RMS wavefront error has a direct e ect on mean intensity pattern from the target camera, shown averaged over the last 1500 time steps in Fig. 7.4. The mean target image for the open loop case is relatively di use; some form of wavefront control is required to generate a peaked intensity prole. While the classical AO loop increased the maximum intensity by 82% compared to the open loop case, the optimal LTI controller produced a nearly 140% enhancement. Some of this improvement can be attributed to reduced jitter in the location of the maximum point, as shown buy the reduction in the standard deviation of the peaks location in Table 7.1. However at many instances the maximum value itself increased with the optimal controller, as seen in the time series of the maximum intensity in Fig. 7.3. The J2 performance measure in Table 7.1 is the average maximum of the instantaneous target image, and indicates a 21.6% increase with the classical controller compared to open loop, compared to 29.3% improvement with the optimal LTI controller. Examining the wavefront measurements from a modal perspective allows for more detailed analysis of the underlying statistics. Since only the projections
+ onto the rst 15 DM modes were used for control, E = U15 , and the measurement

sequence e only contains information for those control channels. To investigate the coupling between the controlled and uncontrolled modes, the sequence of slope vectors from the experiment {y(1), y(2) } is projected onto the entire

69

basis of 31 modes q(t) = U + y(t) for t = 1, 2, . . . (7.12)

where q(t) R31 is the projection of y(t) onto all the columns of U . While the DM modes are theoretically orthogonal, the modal reconstructor uses experimentally identied modes that may not be, and in general the rst 15 elements of q(t) do not equal e(t). Figure 7.6 displays the temporal RMS value of the time series of each element in q. The e ectiveness of the controllers is evident from the top plot, which shows the modal power distribution for each of the three control cases. The optimal LTI controller clearly outperformed the classical AO loop in every controlled modal channel, especially those which have high open loop power such as mode 5. The e ect is more striking in the bottom plot in Fig. 7.6, which shows the modal power for the controllers normalized the the corresponding open loop value. Over the rst 15 modes, the classical loop coe cients had a temporal RMS, on average, was 77.5% of the open loop value, compared to a mean of 52.73% for those with the LTI controller. Improvement versus open loop is also evident in the uncontrolled modes greater than 15, indicating that some coupling of the higher frequency control channels existed. Figure 7.7 shows the power spectral densities of a selection of modal coe cients from the sequence q. Consistent with the sensitivity transfer function in Fig. 6.2, the classical AO controller attenuated disturbances up to approximately 1500 to 2000 Hz in the controlled channels, while the optimal controller achieved signicant reduction up to 4000 Hz. The classical controller had no e ect on the spike at 3300 Hz, for example, while the optimal LTI controller signicantly minimized disturbances at that frequency. Note that due to the waterbed e ect, the whitening performance of the optimal controller came at the cost of amplied

70

high frequency disturbances, typically occurring in frequencies above 4000 Hz. These features of the power spectral densities are reected in the time series of the modal coe cients in q, shown in Fig. 7.8 for the same modes as the PSDs. The classical loop distinctly reduced the amplitude of the measurements relative to the open loop case. The optimal LTI controller mitigated the coe cients even further in the controlled channels, but also increased the high frequency content of the time series.
1.8 1.6 1.4 1.2 RMS (m) 1 0.8 0.6 0.4 Open Loop 0.2 0 200 400 600 800 1000 Time step 1200 Classical AO 1400 1600 Optimal LTI 1800 2000

Figure 7.2: 20 step moving RMS of the spatial RMS of the residual wavefront.

71

130 Open loop 120 110 100 Max Pixel Value 90 80 70 60 50 40 Classical AO Optimal LTI

200

400

600

800

1000 1200 Time step

1400

1600

1800

2000

Figure 7.3: 50 step moving average of the maximum intensity of the target camera image.

72

Open Loop Max = 9.0657

Classical AO Max = 16.5268

Optimal LTI Max = 21.6625

20

20

20

15

15

15

10

10

10

Figure 7.4: Mean target camera intensity image averaged over 1500 frames.

73

Open loop

Classical AO

10 20 30 40 50 60

10 20 30 40 50 60 10 20 30 40 50 60 Optimal LTI 10 20 30 40 50 60

10 20 30 40 50 60

10

20

30

40

50

60

10

20

30

Figure 7.5: Scatter plot showing the number of time steps the maximum instantaneous intensity was located in each pixel

74

160 140 120 100 RMS 80 60 40 20 0 5 10 15 20 25 30 Open Loop Classical AO Optimal LTI

1.2

1 RMSLTI/RMSOL

0.8

0.6

0.4

0.2

10

15 Mode number

20

25

30

Figure 7.6: (Top) temporal RMS values of individual modal time series. (Bottom) temporal RMS values of the modal time series for the optimal LTI and classical controller results, normalized by corresponding open-loop values.

75

Mode 1 180 160 140 dB 120 100 80 0 2000 4000 6000 Frequency (Hz) Mode 4 180 160 140 dB 120 100 80 0 2000 4000 6000 Frequency (Hz) Mode 10 180 160 140 dB 120 100 80 0 2000 4000 6000 Frequency (Hz) Open Loop 8000 dB 180 160 140 120 100 80 0 2000 8000 dB 180 160 140 120 100 80 0 2000 8000 dB 180 160 140 120 100 80 0 2000

Mode 2

4000 6000 Frequency (Hz) Mode 5

8000

4000 6000 Frequency (Hz) Mode 25

8000

4000 6000 Frequency (Hz)

8000

Classical AO

Optimal LTI

Figure 7.7: Power spectral density for a selection of modal coe cients.

76

Mode 1 6000 4000 2000 0 2000 4000 6000 500


4

Mode 2 6000 4000 2000 0 2000 4000 6000

550

600 Mode 4

650

700

500
4

550

600 Mode 5

650

700

1.5 1

x 10

2 1

x 10

0.5 0 0.5 1 1 1.5 500


4

550

600 Mode 10

650

700

2 500
4

550

600 Mode 25

650

700

2 1 0 1

x 10

1 0.5 0 0.5

x 10

2 500

550

600

650 Open Loop

700

1 500 Classical AO

550

600

650

700

Optimal LTI

Figure 7.8: Time series for a selection of modal coe cients.

77

2 2 Table 7.1: Target camera performance metrics. i and j are the variances of the instantaneous maximum location in the horizontal and vertical directions. J1 and J2 are the far eld performance metrics dened by (4.11) and (4.12).

J1 Open loop 9.07 Classical AO 16.53 Optimal LTI 21.66

J2 66.31 80.32 85.79

2 i

2 j

13.82 9.04 9.97 5.31 7.84 3.72

Table 7.2: RMS of the wavefront error over time and space for time steps 200 to 2000. RMS is the mean reduction in the temporal RMS versus open loop over the rst 15 modal coe cients. RMS (m) Open loop Classical AO Optimal LTI 1.32 1.03 0.705 RMS 20.5 42.4

78

7.4

Conclusions

Compared to the classical AO loop, the identied LTI produced markedly superior performance in both wavefront sensor and target camera metrics. Spatially, the optimal LTI controller signicantly reduced the RMS of the measured wavefront error. Because of the link between the Strehl ratio and the RMS wavefront error, this translated into higher maximum intensities on the target camera, and an overall reduction in beam jitter. These results conrm those in [22] that the optimal LTI loop produces Strehl ratios that are not achievable by the classical AO loop, even when the integrator gain is chosen to maximize the disturbance rejection bandwidth. The power spectral densities of the modal coe cients reveals the e ectiveness of the optimal LTI controller from a temporal perspective. While the classical AO bandwidth is limited by the sensitivity transfer function to less than 2000 Hz, the optimal LTI controller attenuates disturbances up to approximately 4000 Hz. The spectral densities with the LTI controller are nearly white, and its reasonable to infer that the optimal controller would be e ective with increasingly broadband disturbance.

79

CHAPTER 8 Optimal Wavefront Correction with Adaptive Control


The controller in Chapter 7 relies on the a priori identication of a state-space predictor to compensate for the AO loop latency. This chapter considers an adaptive implementation that recursively identies a minimum-variance prediction lter during closed-loop operation. The adaptive lter has the form of a recursive least squares (RLS) lattice lter [2, 17], which is e cient and numerical stable in the presence of multiple channels. A high and variable order version of the lter, with few control channels, has been used for control of laser beam jitter [32, 33]; the version presented here is xed at a relatively low order, but with many control channels, and is similar to the implementations in [22] and [31]. Adaptive optics applications frequently encounter statistically non-stationary wavefront turbulence due to changing ow topology and alterations to the optical path. In the AAOL experiment, for example, aberrations arise from the spatially varying density eld formed around the laser turret; results in [11] and [40] reveal a strong relationship between the spatial statistics of the wavefront turbulence and the turret viewing angle. Such a system would encounter dynamic disturbance statistics if the turret orientation were altered during AO operation, for instance amid a slewing maneuver to track a moving target. An identied minimum-variance disturbance model would quickly become sub-optimal in that

80

scenario, and the resulting LTI controller could potentially perform as poorly or worse than the classical AO loop. In contrast, the adaptive control scheme identies the optimal prediction lter on-line using current wavefront data, and thus has the ability to track changes to the aberration statistics as they occur. Section 8.1 of this chapter presents the architecture of the adaptive controller, and discusses key details of the RLS lattice lter implementation. Experimental results are presented in Section 8.2 for both statistically stationary and nonstationary wavefront turbulence.

8.1

Formulation as an Adaptive Control Problem

The architecture of the adaptive controller is shown in Fig. 8.1. As in the LTI case, the adaptive optics objective is re-formulated as a feed-forward disturbance rejection problem by employing a linear model of the extended plant, G, given by (6.13). The primary component in the adaptive controller is the prediction lter F . Unlike in Chapter 7, the lter is not constructed from an explicit statespace model of the wavefront turbulence. Instead, F is an RLS lattice lter that recursively estimates the statistics of the disturbance sequence to minimize the residual wavefront. For the purpose of generating a control command, F is a multichannel N th order nite impulse response lter, having the the general form
N 1 X k=0 k

F (z) =

Fk z

(8.1)

where Fk is a matrix gain. Each element of the command vector u is generated using information from all control channels, and the matrix coe cients are generally non-diagonal. As with the LTI lter, this ensures that both spatial and

81

temporal correlations are used to determine the optimal control output. In the actual lter realization, the RLS lattice equations represented by F are more complex than would be the case with the general FIR model in (8.1), updated using a classical RLS algorithm [48].

Figure 8.1: Block diagram of the extended plant with the adaptive controller. The controller contains two copies of F . One implements the fully adaptive lter realization shown in Fig. 8.2 using the ltered signal Ge as the input. The diagonal arrow in Fig. 8.1 indicates that the gains of this lter are adjusted during operation to minimize the RMS value of the e. The second instance of F has the same impulse response as the adaptive copy. Occupying the corresponding position as the prediction lter in Fig. 7.1, this block generates the optimal control command from the signal r. The lattice lter algorithm attempts to recursively minimize the RMS value

82

of the tuning signal e, which is an estimate of the wavefront error e = r + F Gr (8.2)

When G = G, the signal r becomes an estimate of w. Making use of the fact that G is a scalar transfer function and commutes with F (i.e. an example of the ltered-x method [58]), e is equivalent to r + F Gr = r + GF r = w + GF r = w + Gu = e (8.3)

As with the LTI controller, when the extended plant has a causal inverse, the optimal lter is F = G 1 . The RLS lter here can be viewed as an example

of adaptive inverse control [58], where the controller attempts to invert the plant dynamics with an FIR approximation of the true inverse. Because loop latency fundamentally limits AO performance, the numerical e ciency of an optimal control implementation is a concern. For a static LTI controller, the computational cost of computing a control command is negligible once the optimal predictor has been identied, a step that occurs o -line before the start of the control sequence. In comparison, the exibility of an adaptive ltering approach comes at the cost of a considerable computational burden, and the prediction ability of an adaptive control algorithm must be substantial enough to overcome any subsequent increase in the loop latency. Although in theory any RLS algorithm could be used to update the gains of F , the adaptive lter here uses the RLS lattice structure shown in Fig. 8.2, and is based on the multichannel, unwindowed implementation described in [17]. Of the various RLS estimation algorithms, lattice lters are among the most computationally e cient. The version in [17] is particularly suited to adaptive

83

optics applications because of a unique channel-orthogonalization process that results in superior numerical stability with many input and output channels. The lattice lter structure in Fig. 8.2 is composed of a cascade of N identical stages. The equations for each forward and backward block are complex, and given in [17] for the MIMO case. A unique benet of the lattice lter, not explicitly used in the results here, is that each stage produces a control command un , which is the optimal control command if F were an FIR lter with order n < N.

-z

A A A

- Forward
Block

1;1

-z

- Forward A A A A U A
Block n=2

2;1

- p p p -z

N ;0

A A

- Forward
Block

A A U A

n=1

- Backward e1;k Block

u1

- Backward e2;k - p p p - Block - ppp u2

A A A U A

n=N

eN ;0 uN
1

- Backward - Block
uN

Figure 8.2: The lattice lter implementation for an FIR lter of order N .

8.2

Experimental Results

Multiple experiments were run with the adaptive controller to assess its performance characteristics versus the optimal LTI and classical AO control loops. Each experiment compared the e cacy of four control scenarios in response to the same sequence of disturbance commands: Open-loop: c = 0, u = 0. No control command to DM1 Classical AO: c = 0, u = 0. Only the classical AO loop is closed.

84

Optimal LTI Control: c = 0, u = 0 for t > 100. Both the classical and optimal control loops are closed with the optimal LTI controller as U (z), after F is identied using 4000 samples of a separate but statistically identical disturbance sequence. The optimal control loop is closed after 100 time steps to allow the classical controller to reach steady-state. Adaptive Control: c = 0, u = 0 for t > 200. Both the classical and optimal control loops are closed with the adaptive lter as U (z). The adaptive controller begins a period of 100 learning steps at t = 100, after which generates control commands. For the experiments in this section, the F had the form of a 4th order FIR lter with RLS forgetting factor of 0.99999.

8.2.1

Adaptive Control with Statistically Stationary Wavefront Turbulence

The rst adaptive control experiment used disturbance sequence ND2 to compare the performance of the optimal LTI and adaptive controllers. 15 modal channels were used for control, and the classical AO parameters were set to K = 0.3 and = 0.95. Figure 8.3 shows a moving RMS of the spatial RMS of the wavefront error calculated using (4.9) at each time step. The classical AO loop succeeded in reducing some of the larger open loop disturbances, however the optimal LTI loop showed immediate improvement over classical AO curve once the supplemental control loop is closed at time step 100. The adaptive loop took several hundred additional steps to converge, but closely matched the RMS of the LTI controller from approximately the 500th time step onwards. The degree of similarity between the LTI and adaptive curves is striking considering they represent entirely di erent lter structures, and are constructed from vastly di erent procedures.

85

Also shown is the spatial RMS resulting from the ideal integrator case described in Sec. 6.3. The performance of the LTI and adaptive controller was essentially bounded by that of the classical AO and ideal cases. Table 8.2 summarizes the spatial and temporal RMS values for the nal 1000 time steps. While the classical AO loop reduced the open loop RMS by 20.4%, the optimal LTI and adaptive controllers mitigated the open loop value by 45.9% and 46.5%, respectively. For the ideal case, the maximum reduction is 73.1%, most likely a limit resulting from the nite resolution of the deformable mirror. Figure 8.4 shows the mean far-eld intensity pattern from the target camera for each of the four control scenarios, and closely matches the results in Fig. 8.3. Like the wavefront error RMS values, the LTI and adaptive controllers possess nearly identical performance results, and show a marked improvement in the maximum intensity and sharpness of the mean image, compared to the classical AO case. The di erence between the LTI and adaptive maximums is in fact within the noise level commonly witnessed on the target camera from multiple exposures of the same intensity pattern, suggesting that the controllers o ered practically identical performance enhancements with regards to the far-eld intensity. As with the LTI results in Sec. 7.3, some of the increase in the maximum intensity is due to a reduction in beam jitter. Figure 8.5 reveals that the LTI and adaptive controllers increased the instantaneous maximum of the target intensity. As shown in Table 8.1, the mean instantaneous maximum intensity (the J2 performance measure) increased by 14.3% with the classical AO controller compared to the open loop value, while the LTI and adaptive controllers provided an increase of 33.2% and 33.6%, respectively. Modal analysis, as described in Sec. 7.3 for all of the deformable mirror modes, provides further detail on the similarities between the LTI and adaptive

86

controllers. Figure 8.7 shows the temporal RMS of each modal sequence, and the RMS for each controlled scenario normalized by the corresponding open loop value. The results show that the LTI and adaptive controllers produced analogous improvements in every controlled modal channel, and not just in gross measures of performance such as the spatial RMS. For this disturbance sequence, the temporal RMS values were close to the open loop values for the uncontrolled modes 16 through 31, indicating a high degree of orthogonality between channels. Table 8.2 shows the mean of the normalized temporal RMS, averaged over modes 1 thorough 15. The classical AO values are on average 79% of the open loop RMS, while the LTI and adaptive controllers reduce the RMS to 56% and 52% the open loop value. Shown again are the values for the ideal integrator case, which reduces the mean temporal RMS over the controlled modes to 27% their open loop values. Figure 8.8 shows the power spectral density for a selection of the modal coe cients. The classical AO controller mitigated open loop disturbances up to approximately 1000 to 1500 Hz in the controlled channels, as predicted by the sensitivity transfer function in Fig. 6.2. The optimal LTI and adaptive controllers provided much higher bandwidth, signicantly attenuating disturbances up to approximately 4000 Hz. For instance, both signicantly reduced the peak at 3300 Hz, which was unabated by the classical controller. As expected from Fig. 8.7, none of the control methods had a signicant e ect in the uncontrolled channels. A segment of the time series for the same selection of modes is shown in Fig. 8.9. As implied by the power spectral densities, the classical AO controller lessened the e ect of some of the large disturbances, but had little e ect on clearly periodic turbulence. In contrast, the LTI and adaptive controllers severely damp-

87

ened the large disturbances, especially those with a discernible periodicity. Due to the waterbed e ect, the coe cients experienced some high frequency amplication with the LTI and adaptive controllers, but the amplitude of these additional disturbances is small in comparison to the overall disturbance reduction.
1 Open Loop 0.9 0.8 0.7 RMS (m) 0.6 0.5 0.4 0.3 0.2 0.1 0 0 500 1000 1500 2000 Time step 2500 3000 3500 4000 Classical AO Optimal LTI Adaptive Ideal

Figure 8.3: 50 step moving RMS of the RMS wavefront error.

88

Open loop Max = 15.124

Classical AO Max = 18.6242

Optimal LTI Max = 26.7798

Adaptive Max = 26.1713

25

25

25

25

20

20

20

20

15

15

15

15

10

10

10

10

Figure 8.4: Mean target camera image over 4000 samples in response to a statistically stationary disturbance sequence.
110 100 90 Max Pixel Value 80 70 60 50 40 30 20 500 1000 Open loop 1500 Classical AO 2000 Time step 2500 Optimal LTI 3000 3500 Adaptive 4000

Figure 8.5: 25 step moving average of the maximum target camera intensity.

89

Open loop 15 20 25 30 35 40 45 50 55 20 30 40 50 15 20 25 30 35 40 45 50 55 20

Classical AO

30

40

50

Optimal LTI 15 20 25 30 35 40 45 50 55 20 30 40 50 15 20 25 30 35 40 45 50 55 20

Adaptive

30

40

50

50

100

150

Figure 8.6: Scatter plot showing the number of time steps the maximum instantaneous intensity was located in each pixel

90

150 Open Loop Classical AO Optimal LTI Adaptive Ideal

100 RMS 50 0

10

15

20

25

30

1 RMS/RMSOL 0.8 0.6 0.4 0.2 0

10

15 Mode

20

25

30

Figure 8.7: Modal power. (Top) Temporal RMS values of each modal time series. (Bottom) Temporal RMS of each modal time series normalized by the corresponding open loop value.

91

Mode 1 160 140 120 dB 100 80 60 dB 160 140 120 100 80 60

Mode 2

2000

4000 6000 Frequency (Hz) Mode 4

8000

2000

4000 6000 Frequency (Hz) Mode 5

8000

160 140 120 dB 100 80 60 dB

160 140 120 100 80 60

2000

4000 6000 Frequency (Hz) Mode 10

8000

2000

4000 6000 Frequency (Hz) Mode 25

8000

160 140 120 dB 100 80 60 dB

160 140 120 100 80 60

2000

4000 6000 Frequency (Hz) Open Loop

8000

2000

4000 6000 Frequency (Hz) Adaptive Ideal

8000

Classical

Optimal LTI

Figure 8.8: Power spectral densities of select modal sequences

92

Mode 1 4000 Modal Coefficient 2000 0 2000 4000 1500 Modal Coefficient 4000 2000 0 2000 4000 1500
4

Mode 2

1550

1600 Time step Mode 4

1650

1700

1550

1600 Time step Mode 5

1650

1700

5000 Modal Coefficient Modal Coefficient

1 0.5 0 0.5

x 10

5000

10000 1500
4

1550

1600 Time step Mode 10

1650

1700

1 1500

1550

1600 Time step Mode 25

1650

1700

1 Modal Coefficient 0.5 0 0.5

x 10

6000 Modal Coefficient 4000 2000 0 2000 4000 1500

1 1500

1550

1600 Time step

1650

1700

1550

1600 Time step Adaptive

1650

1700

Open Loop

Classical

Optimal LTI

Figure 8.9: Coe cients for a selection of modal control channels

93

2 2 Table 8.1: Target camera performance metrics. i and j are the variances of the instantaneous maximum location in the horizontal and vertical directions. J1 and J2 are the far eld performance metrics dened by (4.11) and (4.12).

J1

J2

2 i

2 j

Open loop 15.12 55.82 28.17 46.39 Classical AO 18.62 63.80 15.14 31.25 Optimal LTI 26.78 74.39 10.36 13.59 Adaptive 26.17 74.57 11.10 13.54

Table 8.2: RMS of the wavefront error over time and space. RMS is the mean reduction in the temporal RMS versus open loop over the rst 15 modal coe cients. RMS (m) Open loop Classical AO Optimal LTI Adaptive Ideal benchmark 0.48 0.38 0.26 0.25 0.13 RMS (%) 20.7 43.7 46.0 72.8

94

8.2.2

Adaptive Control with Non-Stationary Wavefront Turbulence

To investigate the tracking ability of the adaptive controller, a disturbance sequence was applied where the ow characteristics change midway through the control process. Two sequences of 8000 disturbance commands were generated: ND2, which was used for identication and control, and ND3, which was used for control only. The AAOL wavefront images used to generated ND3 were identical to those responsible for ND2, except for a 90 degree clockwise rotation. Other parameters related to ND2 and 3 are listed in Table 5.1. For the experiment here, the classical AO parameters were K = 0.3 and = 0.95. The rst 15 deformable

modes were used for control, however analysis was performed by projecting the WFS measurements onto the entire modal basis. First, an optimal LTI controller was identied by applying frames 1 through 4000 of ND2 with the classical loop closed, and following the procedure outlined in Sec. 7.2. Then the experiment was run using the four control scenarios in Sec. 8.2.1 for 8000 time steps. The disturbance commands consisted of wavefronts 4001 through 8000 from ND2, followed immediately by wavefronts 4001 through 8000 from ND3. Using this combination ensured that any performance results were due solely to changes in the ow orientation, and not variations in the turbulence strength between dissimilar segments of ND2 and 3. For the results that follow, T1 denotes time steps 200 through 4000 during which the ND2 sequence was used. Time steps 4001 through 8000 are referred to as T2 , during which the ND3 sequence was used. Figure 8.10(a) shows the RMS of the residual wavefront measured by the WFS. As in the results from Sec. 8.2.1, the adaptive and LTI controllers initially show similar performance improvements over the classical AO controller once the RLS algorithm converges. Table 8.4 shows that the RMS over space and time

95

over T1 with the LTI controller was 48.3% of the open loop value, while with the adaptive controller it was 45.9%. This compares with a value of 73.5% with the classical AO loop. The level of performance with LTI controller is expected since the disturbance statistics for this period were the same as those used to identify the prediction lter in the LTI controller. Also shown as a performance benchmark is the ideal integrator case detailed in Sec. 6.3. Both the LTI and adaptive controllers approached this ideal case at numerous time instances, and were otherwise bounded by that benchmark and the RMS from the classical AO loop. The situation changed after the 4000th time step, when the disturbance sequence switched to commands from the ND3 sequence. The optimal LTI controller was instantaneously rendered sub-optimal, and its performance su ered considerably. The RMS error with the adaptive controller also experienced a jump at this point, however it diverged from the LTI case and settled to an average value similar to during the ND2 sequence. During T2 , the spatial and temporal RMS with the LTI controller was 65.2% of the open loop value, similar to the 71.9% value with the classical AO controller. In contrast the adaptive loop still achieved an RMS of 47.0% the open loop value for the second 4000 time steps. Note that the curve for the ideal case during T2 is quite similar to that during T1 , which is to be expected since the underlying AAOL wavefronts were identical except for a rotation. As indicated in Table 8.4, the adaptive controller outperformed the LTI loop by a wide margin during this period, but did not achieve the absolute level of performance seen in the rst half of the experiment. This could be due to slight misalignment between the disturbance and command DMs, causing turbulence owing in one direction to be more di cult to predict than another. These performance results are also evident in Fig. 8.10(b), which shows the di erence between the residual RMS for the adaptive and LTI

96

controllers and the ideal benchmark. The time-averaged target camera images shown in Fig. 8.11 reect the behavior of the wavefront error RMS. For T1 , the LTI and adaptive controllers showed a similar increase in the maximum intensity; the LTI controller increased the maximum value of the mean intensity by 68.7% compared to the classical AO value, while the adaptive controller provided a 64.25% increase. For the images averaged over the subsequent 4000 time steps the LTI controller was closer to the classical AO values. Here the maximum with the LTI controller was onlt 19.7% above the classical AO maximum, while the adaptive controller produced a 72.6% increase over the same period. Figure 8.2.2 indicates that the increases in the mean target images with the adaptive and LTI controllers resulted partly from an increase in the instantaneous maximum, and not solely from a reduction in beam jitter. The similarity between the LTI and classical loops during the period T2 suggests that the optimal control commands for this period were not signicantly impacting the overall wavefront, and essentially entered the classical controller as noise. Figures 8.2.2 and 8.2.2 compare the power spectral densities for several modes for T1 and T2 . For the initial period, both LTI and adaptive controllers behaved similarly here as well, attenuating correlated noise to near the benchmark ideal value up to approximately 2000 Hz in controlled modal channels. Note that the performance of these controllers is e ectively bounded by waterbed e ect, and can only atten the PSD to the diagonal values in the covariance matrix of the white-noise input in (7.8). The ideal benchmark scenario, however, is not

bound by this restriction, and can mitigate disturbances to levels unachievable by linear feedback controllers. After the 4000th time step, Fig. 8.2.2 shows that the e ectiveness of the LTI controller dropped to closely match the limited bandwidth

97

of the classical loop; generally less than 1000 Hz, and roughly corresponding to the gain crossover frequency in the sensitivity transfer function in Fig. 6.2. Further, while the bandwidth of the LTI controller constricted, it still amplied high-frequency noise in a fashion similar to the adaptive controller. Thus, the sub-optimal LTI lter performed with the least desirable aspects of the classical and adaptive loops: low-bandwidth disturbance rejection, combined with high frequency amplication.

98

0.9 Open Loop 0.8 0.7 0.6 RMS (m) 0.5 0.4 0.3 0.2 0.1 Classical AO Optimal LTI Adaptive Ideal

1000

2000

3000

4000 Time step

5000

6000

7000

8000

(a)
0.35 Optimal LTI Adaptive 0.3

0.25 RMS RMSideal(m)

0.2

0.15

0.1

0.05

1000

2000

3000

4000 Time step

5000

6000

7000

8000

(b)

Figure 8.10: (a) 20 step moving RMS of the RMS wavefront error, (b) Di erence between the RMS wavefront error for the LTI and adaptive controllers, and the ideal benchmark case.

99

Open Loop Max = 23.54

Classical AO Max = 32.59

Optimal LTI Max = 51.32

Adaptive Max = 49.25

45 40 35 30 25 20 15 10 5 0

45 40 35 30 25 20 15 10 5 0

45 40 35 30 25 20 15 10 5 0

45 40 35 30 25 20 15 10 5 0

(a)
Open Loop Max = 28.72 Classical AO Max = 36.95 Optimal LTI Max = 38.75 Adaptive Max = 49.37

45 40 35 30 25 20 15 10 5 0

45 40 35 30 25 20 15 10 5 0

45 40 35 30 25 20 15 10 5 0

45 40 35 30 25 20 15 10 5 0

(b)

Figure 8.11: (a) Mean target camera image for the time period T1 , (b) Mean target camera image for the period T2 .

100

180 Open loop 160 Classical AO Optimal LTI Adaptive

140 Max Pixel Value

120

100

80

60

40

500

1000

1500

2000 Time step

2500

3000

3500

4000

(a)
180 Open loop 160 Classical AO Optimal LTI Adaptive

140 Max Pixel Value

120

100

80

60

40

4500

5000

5500

6000 Time step

6500

7000

7500

8000

(b)

Figure 8.12: (a) 25 step moving RMS of the maximum target camera intensity over time period T1 , (b) 25 step moving RMS of the maximum target camera intensity over T2 .

101

Open loop 15 20 25 30 35 40 45 50 55 20 30 40 50 15 20 25 30 35 40 45 50 55 20

Classical AO

30

40

50

Optimal LTI 15 20 25 30 35 40 45 50 55 20 30 40 50 15 20 25 30 35 40 45 50 55 20

Adaptive

30

40

50

50

100

Figure 8.13: Scatter plot showing the number of time steps the maximum instantaneous intensity was located in each pixel for during period T1 .

102

Open loop 15 20 25 30 35 40 45 50 55 20 30 40 50 15 20 25 30 35 40 45 50 55 20

Classical AO

30

40

50

Optimal LTI 15 20 25 30 35 40 45 50 55 20 30 40 50 15 20 25 30 35 40 45 50 55 20

Adaptive

30

40

50

50

100

Figure 8.14: Scatter plot showing the number of time steps the maximum instantaneous intensity was located in each pixel for during period T2 .

103

Mode 1 160 140 dB 120 100 80 dB 160 140 120 100 80

Mode 4

2000

4000 6000 Frequency (Hz) Mode 7

8000

2000

4000 6000 Frequency (Hz) Mode 9

8000

160 140 dB 120 100 80 dB 0 2000 4000 6000 Frequency (Hz) Mode 13 160 140 dB 120 100 80 dB 8000

160 140 120 100 80

2000

4000 6000 Frequency (Hz) Mode 25

8000

160 140 120 100 80

2000

4000 6000 Frequency (Hz) Open Loop

8000

2000

4000 6000 Frequency (Hz) Ideal

8000

Classical

Optimal LTI

Adaptive

Figure 8.15: Power spectral densities of select modal sequences over period T1 .

104

Mode 1 160 140 dB 120 100 80 dB 160 140 120 100 80

Mode 4

2000

4000 6000 Frequency (Hz) Mode 7

8000

2000

4000 6000 Frequency (Hz) Mode 9

8000

160 140 dB 120 100 80 dB 0 2000 4000 6000 Frequency (Hz) Mode 13 160 140 dB 120 100 80 dB 8000

160 140 120 100 80

2000

4000 6000 Frequency (Hz) Mode 25

8000

160 140 120 100 80

2000

4000 6000 Frequency (Hz) Classical

8000

2000

4000 6000 Frequency (Hz) Ideal

8000

Open Loop

Optimal LTI

Adaptive

Figure 8.16: Power spectral densities of select modal sequences over T2 .

105

Table 8.3: Target camera performance metrics with the non-stationary distur2 2 bance sequence. i and j are the variances of the instantaneous maximum location in the horizontal and vertical directions. J1 and J2 are the far eld performance metrics dened by (4.11) and (4.12). J1 T1 T2 T1 J2 T2 T1
2 i 2 j

T2

T1

T2

Open loop 23.5 28.7 85.8 94.9 34.8 Classical AO 32.6 36.9 101.5 109.1 21.7 Optimal LTI 51.3 38.7 128.1 111.3 7.2 Adaptive 49.2 49.3 121.2 122.4 7.6

13.4 22.1 29.7 7.8 12.3 16.3 8.4 6.1 9.9 6.1 5.8 6.1

Table 8.4: RMS of the wavefront error over time and space. RMS is the mean reduction in the temporal RMS versus open loop over the rst 15 modal coe cients. RMS (m) T1 Open loop Classical AO Optimal LTI Adaptive Ideal benchmark 0.56 0.41 0.27 0.26 0.18 T2 0.59 0.42 0.36 0.28 0.19 RMS (%) T1 T2 26.1 27.1 51.2 37.2 55.9 52.7 66.4 68.6

106

8.3

Conclusions

Despite the FIR structure of the prediction lter, the adaptive lter performance was comparable to the IIR optimal LTI lter in response to statistically stationary turbulence. On a modal level, the controllers whitened the output power spectral densities over a comparable bandwidth, far outperforming the classical AO controller, while introducing noticeable, but negligible high frequency noise. In gross measures of AO performance, such as the spatial RMS and mean far-eld intensity, the adaptive and LTI results are nearly indistinguishable; di erences between the results are close enough to be attributable to measurement noise. While the identied LTI lter is theoretically optimal for the disturbance sequence being used, in practice an adaptive FIR lter of su cient length will match or exceed the performance of a pre-identied IIR lter. Since it attempts to minimize the residual wavefront directly, the adaptive lter has some ability to compensate for errors in the plant model G, or changes in the poke matrix estimate U . This is particularly useful in the application here due to the longterm DM drift discussed in Sec. 3.3. However the correspondence between the controllers in the experiments here suggest that there was minimal plant modeling error, and the internal model of the plant satised (7.2). Of particular interest is that an FIR lter of order 4 is adequate was match the performance of the IIR prediction lter at the heart of the LTI controller. This suggests that spatial correlations, represented by the multichannel nature of the RLS lattice implementation, are su cient to predict a large portion of the wavefront; the turbulence wavefronts are not signicantly correlated beyond 4 time steps in this disturbance sequence. The ow velocity may play a role in this phenomenon, in that it takes 4 time steps or less for a particular turbulence structure to move across the WFS aperture.

107

Changing the disturbance statistics during the control process exposed the weakness of using a pre-identied LTI lter. While the controller successfully rejected disturbances for which it was optimal, changing the turbulence statistics reduced its e ectiveness that of the classical AO controller. The power spectral density for this case conrms that the disturbance rejection bandwidth of the LTI controller contracted to no more than the classical case, yet it still amplied high frequency noise. In comparison, after a transitory learning period the adaptive controller maintained its performance in both sets of disturbance statistics. A rough comparison of the adaptive results to the ideal integrator case reveals that approximately 500 to 1000 learning steps were necessary before the RLS lattice lter converged to a steady-state value. This adaptability comes a the cost of signicant computational complexity, an aspect not accounted for in the pseudo-real time nature of the experiment.

108

APPENDIX A Estimating Flow Velocity from Disturbance Models


Many approaches exist in adaptive optics literature that provide an estimate of the overall ow velocity based in sample turbulence wavefronts. The majority of these techniques rely on analysis of the 2-dimensional correlation image between wavefronts separated by a known time interval. Often, these methods attempt to identify a maximum point either directly from the image covariance [1, 49], or from a frequency-based modal decomposition of the wavefront [42]. Velocity estimation from image correlations is grounded in the assumption that the wavefronts result from a static phase prole shifting across the WFS aperture with a constant velocity. Under this model, the wavefront measurement at time t is a shifted version of the measurement at time t + , for su ciently small, with the addition of some amount of uncorrelated turbulence at the boundaries. The direction and magnitude of the shift, divided by yields the ow velocity. This section considers an alternative method to directly computing the correlation image from a sequence of wavefronts. Instead, the correlation image is constructed directly from the output covariance matrix of an identied state-space model of the wavefront turbulence. This allows for the covariance image to be constructed for any number of time delays, without generating sample estimates directly from data.

109

The Two-Dimensional Cross-Correlation Image Consider two square images of equal dimension, A RN N and B RN N , represented as matrices where ai,j denotes element (i, j) of A.1 The 2-dimensional cross-correlation image is a matrix of size (2N is given by cn,m = X
i,j

1)(2N

1), where each element

ai,j bi

p,j q

(A.1)

where the indices p and q are for notational convenience and given by p= q= (N (N 1) + n 1) + m 1 1. (A.2)

Note that in practice, evaluating the product and summation in (A.1) may require padding the matrix A with additional zeros. Despite the complex indexing, C is conceptually similar to a 1-dimensional cross-correlation, where the location of the maximum point in C represents the shift B would undergo to maximize the cross-correlation with A. In particular, to calculate cn,m , the matrix B is translated p elements vertically and q elements horizontally over A, where p and q are given by (A.2). The element cn,m is then equal to the sum of the product of the overlapping elements. For example, the center of the cross-correlation image, cN,N , corresponds to zero translation with (p, q) = (0, 0), and is given by the sum of ai,j bi,j , 1 i, j N . The relationship (A.1) can be extended to non-square matrices with di erent dimensions at the expense of additional notational complexity, but for the method that follows, the matrices A and B are always the same size.
1 This equation and what follows uses standard Matlab (row, column) notation, where the row indices increase from top to bottom, and the column indices increase from left to right.

110

Figure A.1: Graphical example of the calculation of the cross-correlation element c3,5 for images A, B R44 . Figure A.1 graphically illustrates the calculation of an element of the crosscorrelation image for matrices A, B R44 . For the example shown, B is shifted one pixel horizontally and one pixel vertically relative to A, thus (p, q) = ( 1, 1) under the sign conventions used here. Since N = 4, this shift corresponds to (n, m) = (3, 5), and c3,5 is equal to the sum of the product of the pixels in the shaded, overlapping region. Equation (A.1) can be extended in a stochastic sense to an auto-correlation for a sequence of images. Consider a sequence of images {W (1), W (2), W (3) }, where each image W (t) is N N . The 2-dimensional auto-correlation of the sequence, CW (), is dened by replacing A and B in (A.1) with instances of W separated by a given time delay , and taking the expectation ( X
i,j

cn,m () = E

wi,j (t)wi

p,j q (t

+ )

(A.3)

where E{} represents the sample expectation, and p and q are still given by

111

(A.2). If the images W are related by uniform translations at each time step, with uncorrelated data replacing vacated pixels, the overall ow velocity can be estimated by searching for a peak in Cw (). If (nmax , mmax ) is the coordinate of the maximum pixel in CW (), for a particular [0, ], then the velocity in the horizontal and vertical directions can be estimated as vi = nmax mmax , vj = (A.4)

The maximum time shift that yields an adequate estimate depends on the size of the images relative to the actual translation velocity. Equation (A.4) clearly provides a poor approximation if the velocity is such that images are entirely uncorrelated in time. As an example, suppose each W (t + 1) is equal to shifting W (t) horizontally one column to the right, and inserting uncorrelated data into the new column. Under the sign convention here, this translates into a velocity of v = [0, 1] pixels per time step. Clearly, W (t+1) is most correlated with W (t) by shifting W (t+1) it one column to the left, thus we would expect to nd a peak in CW (1) at a location corresponding to (p, q) = (0, 1).

Auto-Correlation Image from the State-Space Covariance Matrix For adaptive optics systems, (A.3) and (A.4) can be used to estimate the aberration ow velocity if the images are a sequence of wavefront proles, and the ow velocity is low relative to the spatial resolution and size of the wavefront sensor [49]. In practice, an accurate velocity estimate requires many sample frames of wavefront data over which the ow velocity must be reasonably con-

112

stant. The optimal LTI methods in Sec. 7 rely on a linear model of the disturbance generated from wavefront samples. Suppose the wavefront sequence {W (1), W (2), W (3) }, where each W is a N N image, is modeled using the methods in Sec. 7.2 as the output of a stable LTI system driven by a while noise sequence (t) with covariance matrix Q = E[ (t) (t)T ], x(t + 1) = Ax(t) + K (t) w(t) = Cx(t) + D (t) where the output w(t) RN (with no subscripts) is a vectorized wavefront image with the same statistics as W (t). Note that in Sec. 7, the output of the identied disturbance model is the wavefront expressed in modal coordinates, thus the state-space matrices must be multiplied by the appropriate reconstructor matrix to construct F . As discussed in Sec. 7, the states in (A.5) generally have no physical meaning. However, the matrices A, K, and C encapsulate the ow statistics, and the overall ow velocity can be estimated directly from without generating sample wavefronts. Consider the output covariance matrix for the system F w () = E{w(t)w(t + )T } (A.6)
2

F :

(A.5)

Since w(t) and w(t + ) are vectorized wavefront images with the same rst and second order statistics as the sequence W , every element in the auto-correlation image CW () can be constructed as a linear combination of elements in w (). The benet of this approach is that w () may be computed easily for any time lag directly from F as w () = CA C T + DQDT (A.7)

113

where is the state covariance matrix satisfying the discrete-time Lyapunov equation = AAT + KQK T (A.8)

The exact process of extracting the correct elements from w () depends on the indexing system used to vectorize 2-dimensional matrices. In Matlab, the function sub2ind provides the linear index, , corresponding to the position (i, j) in a matrix of size N N = f (i, j, N ) = N (j 1) + i (A.9)

where f () represents the sub2ind function. In the following, double subscripts as in wi,j denotes the (i, j) element of the 2-dimensional matrix W , while single indices, as in w refers to the th element in the vector w. With this notation, the (n, m) element of CW () may computed from elements of vectorized images ( X
i,j

cn,m () = E

wi,j (t)wi

p,j q (t

+ )

X
i,j

E {w (t)w (t + )}

(A.10)

Hence, using the denition of the output covariance matrix cn,m () = X


i,j

, ()

(A.11)

where , () is the (, ) element of w (), and the indices and are given by = f (i, j, N ) = f (i p, j q, N ) (A.12)

114

and p and q are given by (A.2). Equation (A.11) can therefore be used to calculate every element in CW () for an arbitrary time delay directly from the state-space realization of the wavefront sequence given by (A.5).

A Wavefront Flow Example To illustrate the velocity estimation method, a 20 state state-space model was identied from wavefronts generated by WaveTrain, a wave optics simulation program by MZA Associates Corporation. The turbulence was the result of a single phase screen model, with the ow entirely in the j direction. The velocity

vj was generated by calculating the output covariance matrix w () and applying (A.11). For each time delay, the peak location of CW () resulted from a linear spline interpolation on a ner grid of points, however this did not signicantly impact the results for the example here. Figure A.2 shows the correlation image constructed from the output covariance matrix for a time delay of = 7. Note that while there are numerous local extrema, the global maximum is located along the ow direction on the j axis. Figure A.3 illustrates the location of the maximum of the correlation image versus the number of time delays. While not completely linear, the slope of this curve provides an adequate estimate of the ow velocity to be vj = 4.54 pixelsx/time step . (A.13)

This value matches well with visual inspection of the ow.

115

x 10 6 4 2 0 2 4 35 30

12

25 20 15 10 5 0 5 0 15 10 25 20 30 35

Figure A.2: Surface plot of CW () constructed from (A.11) with the WaveTrain realization for time delay = 7 .

0 1 Location of max(CW) 2 3 4 5 6 7 8 1 2 3 4 5 6 Time delay 7 8 9 10

Figure A.3: Location of the maximum of CW () along the j axis versus the time delay.

116

APPENDIX B Primary Component Specications


Table B.1: Shack-Hartmann Wavefront Sensor Specications Camera model AVT Marlin F131B Image Resolution 280 280 pixels Subaperture Resolution 12 12 E ective Frame Rate 400 fps Pixel pitch 6.7 m Lenslet focal length 6.7 mm Subaperture grid size used 12 12 Dynamic Range 0.44 m/lenslet Sensitivity 0.16 m/pixel

Table B.2: Target Camera Specications Camera model AVT Marlin F131B-NIR Image Resolution 300 300 pixels E ective Frame Rate 400 fps Pixel pitch 6.7 m

Table B.3: Membrane Deformable Mirror Specications Diameter 25 mm Max focus throw 10 m Surface reectivity 80% Approx. resonance >500 Hz Damage threshold 3 J/cm2

117

References
[1] A. Beghi, A Cenedese, and A Masiero, On the estimation of atmospheric turbulence layers, Proc. of the 2008 IFAC World Congress (2008). [2] C. C. Chang and J. S. Gibson, Parallel control loops based on spatial subband processing for adaptive optics, American Control Conference (Chicago), IEEE, June 2000. [3] N. Y. Chen and J. S. Gibson, Subspace system identication using a multichannel lattice lter, American Control Conference (Boston, MA), vol. 1, 30 2004-july 2 2004, pp. 855 860 vol.1. [4] C. Correia, H Raynaud, C Kulcsr, and J. Conan, On the optimal recona struction and control of adaptive optical systems with mirror dynamics, J. Opt. Soc. Am. A 27 (2010), no. 2, 333349. [5] B. L. Ellerbroek, C. Van Loan, N. P. Pitsianis, and R. J. Plemmons, Optimizing closed-loop adaptive-optics performance with use of multiple control bandwidths, J. Opt. Soc. Am. A 11 (1994), no. 11, 28712886. [6] R. Fraanje, J. Rice, M. Verhaegen, and N. Doelman, Fast reconstruction and prediction of frozen ow turbulence based on structured kalman ltering, J. Opt. Soc. Am. A 27 (2010), no. 11, A235A245. [7] J. S. Gibson, C. C. Chang, and N. Chen, Adaptive optics with a new modal decomposition of actuator and sensor spaces, American Control Conference (Arlington, VA), vol. 6, June 2001, pp. 46194625 vol.6. [8] J. S. Gibson, C. C. Chang, and B. L. Ellerbroek, Adaptive optics: Wavefront correction by use of adaptive ltering and control, Appl. Opt. 39 (2000), no. 16, 25252538. [9] J. H. Gladstone and T. P. Dale, Researches on the refraction, dispersion, and sensitiveness of liquids., Proceedings of the Royal Society of London 12 (1862), pp. 448453. [10] J. W. Goodman, Introduction to fourier optics, 3rd ed., Roberts and Company, Englewood, Colorado, 2005. [11] S. Gordeyev, J. A. Cress, E. Jumper, and A. B. Cain, Aero-optical environment around a cylindrical turret with a at window, AIAA Journal 49 (2011), 308315.

118

[12] S. Haykin, Modern lters, MacMillian Publishing Company, New York, NY, 1989. [13] E. Hecht, Optics, Addison Wesley, San Fransisco, 2002. [14] K. Hinnen, N. Doelman, and M. Verhaegen, H2-optimal control of an adaptive optics system: part ii, closed-loop controller design, vol. 5903, SPIE, 2005, p. 59030A. [15] K. Hinnen, M. Verhaegen, and N. Doelman, Exploiting the spatiotemporal correlation in adaptive optics using data-driven h2-optimal control, J. Opt. Soc. Am. A 24 (2007), no. 6, 17141725. [16] , A data-driven h2 -optimal control approach for adaptive optics, Control Systems Technology, IEEE Transactions on 16 (2008), no. 3, 381 395.

[17] S. B. Jiang and J. S. Gibson, An unwindowed multichannel lattice lter with orthogonal channels, IEEE Transactions on Signal Processing 43 (1995), no. 12, 28312842. [18] T. Katayama, Subspace methods for system identication, Springer-Verlag, London, 2005. [19] R. G. Lane and M. Tallon, Wave-front reconstruction using a shack hartmann sensor, Appl. Opt. 31 (1992), no. 32, 69026908. [20] A. Laub, Matrix analysis for scientists and engineers, SIAM, Philadelphia, PA, 2005. [21] Y. T. Liu and J. S. Gibson, Adaptive optics with adaptive ltering and control, American Control Conference (Boston, MA), vol. 4, 30 2004-july 2 2004, pp. 3176 3179 vol.4. [22] , Adaptive control in adaptive optics for directed-energy systems, Optical Engineering 46 (2007), no. 4, 046601.

[23] D. P. Looze, Realization of systems with ccd-based measurements, Automatica 41 (2005), no. 11, 2005 2009. [24] J. D. Mansell, Getting started with aos hardware, Application note, Active Optical Systems, 2008. [25] , Membrane deformable mirror frequency response, Application note, Active Optical Systems, 2010.

119

[26] J. D. Mansell and R. L. Byer, Micromachined silicon deformable mirror, vol. 3353, SPIE, 1998, pp. 896901. [27] J. D. Mansell and B. Henderson, Temporal and spatial characterization of polymer membrane deformable mirrors, vol. 7466, SPIE, 2009, p. 74660D. [28] J. D. Mansell, B. Henderson, B. Wiesner, R. Praus, and S. Coy, A low-cost compact metric adaptive optics system, vol. 6711, SPIE, 2007, p. 67110K. [29] C. L. Meng and J. S. Gibson, Fast identication of parameters in optical systems, Journal of Optimization Theory and Applications 99 (1998), 73 101, 10.1023/A:1021748126888. [30] S. Monirabbasi and J. S. Gibson, Adaptive control in an adaptive optics experiment with simulated turbulence-induced optical wavefronts, Conference on Advanced Wavefront Control (San Diego, CA), SPIE, SPIE, August 2009. [31] , Adaptive control in an adaptive optics experiment, J. Opt. Soc. Am. A 27 (2010), no. 11, A84A96.

[32] P. K. Orzechowski, High-performance adaptive control of optical jitter in laser beam systems, Ph.D. thesis, UCLA, 2007. [33] P. K. Orzechowski, N. Chen, J. S. Gibson, and T. C. Tsao, Optimal suppression of laser beam jitter by high-order rls adaptive control, IEEE Transactions on Control Systems Technology 16 (2008), no. 2, 255267. [34] P. K. Orzechowski, J. S. Gibson, and T. C. Tsao, Characterization of optimal r gains and minimum-variance performance for adaptive disturbance rejection, American Control Conference (New York, NY), July 2007, pp. 1908 1913. [35] P. K. Orzechowski, S. Gibson, T.C. Tsao, D. Herrik, and V. Beazel, Adaptive control in the presence of quantization and saturation: Application to laser beam steering by a liquid crystal device, ACC Procedings (2009). [36] P. Van Overschee and B. DeMoor, Subspace identication for linear systems: Theory-implementation-applications, Kluwer Academic Publishers, Boston, MA, 1996. [37] R. N. Paschall and D. J. Anderson, Linear quadratic gaussian control of a deformable mirror adaptive optics system with time-delayed measurements, Appl. Opt. 32 (1993), no. 31, 63476358.

120

[38] C. Petit, J.-M. Conan, C. Kulcsr, and H.-F. Raynaud, Linear quadratic a gaussian control for adaptive optics and multiconjugate adaptive optics: experimental and numerical analysis, J. Opt. Soc. Am. A 26 (2009), no. 6, 13071325. [39] C. Petit, J.-M. Conan, C. Kulcsar, H.-F. Raynaud, T. Fusco, J. Montri, F. Chemla, and D. Rabaud, O -axis adaptive optics with optimal control: experimental and numerical validation, vol. 5903, SPIE, 2005, p. 59030P. [40] C. Porter, S. Gordeyev, M. Zenk, and E. Jumper, Flight measurements of aero-optical distortions from a at-windowed turret on the airborne aerooptics laboratory (aaol), 41st AIAA Plasmadynamics and Lasers Conference (Honolulu, HI), AIAA, 2011. [41] J. Porter, H.M Queener, J.E. Lin, K. Thorn, and A. Awwal, Adaptive optics for vision science, John Wiley and Sons, Hoboken, NJ, 2006. [42] L. Poyneer, Marcos van Dam, and Jean-Pierre Vran, Experimental verie cation of the frozen ow atmospheric turbulence assumption with use of astronomical adaptive optics telemetry, J. Opt. Soc. Am. A 26 (2009), no. 4, 833846. [43] L. A Poyneer, Bruce A. Macintosh, and Jean-Pierre Vran, Fourier transe form wavefront control with adaptive prediction of the atmosphere, J. Opt. Soc. Am. A 24 (2007), no. 9, 26452660. [44] L. A. Poyneer and Jean-Pierre Vran, Toward feasible and e ective predictive e wavefront control for adaptive optics, vol. 7015, SPIE, 2008, p. 70151E. [45] T. A. Rhoadarmer, L. M. Klein, J. S. Gibson, N. Chen, and Y. T. Liu, Adaptive control and ltering for closed-loop adaptive-optical wavefront reconstruction, Conference on Advanced Wavefront Control (San Diego, CA), vol. 6306, SPIE, August 2006. [46] M. C. Roggemann and B. Welsh, Imaging through turbulence, CRC, New York, NY, 1996. [47] G. Rousset and J. C. Fontanella et al., First di raction limited astronomical images with adaptive optics, Astron. and Astrophys. 230 (1990), L29L32. [48] A. H. Sayed, Fundamentals of adaptive ltering, John Wiley and Sons, Hoboken, NJ, 2003. [49] M. Schck and E. J. Spillar, Measuring wind speeds and turbulence with a o wave-front sensor, Opt. Lett. 23 (1998), no. 3, 150152.

121

[50] W. H. Southwell, Wave-front estimation from wave-front slope measurements, J. Opt. Soc. Am. 70 (1980), no. 8, 9981006. [51] J. L. Speyer and W. H. Chung, Stochastic processes, estimation, and control, SIAM, Philadelphia, PA, 2008. [52] J. Tesch and J. S. Gibson, Optimal and adaptive correction of aero-optical wavefronts in an adaptive optics experiment, Unconventional Imaging and Wavefront Sensing VII (San Diego, CA), SPIE, 2011. [53] J. Tesch, J. S. Gibson, S. Gordeyev, and E. Jumper, Identication, prediction and control of aero optical wavefronts in laser beam propagation, 41st AIAA Plasmadynamics and Lasers Conference (Honolulu, HI), AIAA, June 2011. [54] T.C. Tsao, Optimal feed-forward digital tracking controller design, Journal of Dynamic Systems Measurement and Control-transactions of The Asme 116 (1994). [55] R. K. Tyson, Principles of adaptive optics, Academic Press, New York, 2000. [56] M. Verhaegen and V. Verdult, Filtering and system identication, a least squares approach, Cambridge University Press, New York, NY, 2007. [57] P.K.C. Wang and F.Y. Hadaegh, Computation of static shapes and voltages for micromachined deformable mirrors with nonlinear electrostatic actuators, Microelectromechanical Systems, Journal of 5 (1996), no. 3, 205220. [58] B. Widrow and E. Walsh, Adaptive inverse control, Prentice Hall, Eaglewood Cli s, NJ, 1996.

122

Potrebbero piacerti anche