Identification, Prediction and Control of Aero Optical Wavefronts in Laser Beam Propagation

Similar documents
Adaptive Optics with Adaptive Filtering and Control

Adaptive Control of a MEMS Steering Mirror for Suppression of Laser Beam Jitter

AFRL-SR-AR-TR

Development of a Low-order Adaptive Optics System at Udaipur Solar Observatory

Sensors & Transducers Published by IFSA Publishing, S. L.,

Adaptive Optics for LIGO

Variable-order adaptive control of a microelectromechanical steering mirror for suppression of laser beam jitter

Calibration of AO Systems

Adaptive Control of a Tilt Mirror for Laser Beam Steering*

Measurement of Beacon Anisoplanatism Through a Two-Dimensional, Weakly-Compressible Shear Layer

Proposed Adaptive Optics system for Vainu Bappu Telescope

MODULAR ADAPTIVE OPTICS TESTBED FOR THE NPOI

The Extreme Adaptive Optics test bench at CRAL

Optimization of coupling between Adaptive Optics and Single Mode Fibers ---

Wavefront control for highcontrast

Non-adaptive Wavefront Control

Measurement of Beacon Anisoplanatism Through a Two- Dimensional Weakly-Compressible Shear Layer

Horizontal propagation deep turbulence test bed

High contrast imaging lab

ADVANCED CONTROL FILTERING AND PREDICTION FOR PHASED ARRAYS IN DIRECTED ENERGY SYSTEMS

The Wavefront Control System for the Keck Telescope

Performance of Keck Adaptive Optics with Sodium Laser Guide Stars

Modeling the multi-conjugate adaptive optics system of the E-ELT. Laura Schreiber Carmelo Arcidiacono Giovanni Bregoli

Status of the DKIST Solar Adaptive Optics System

Open-loop performance of a high dynamic range reflective wavefront sensor

Focal Plane and non-linear Curvature Wavefront Sensing for High Contrast Coronagraphic Adaptive Optics Imaging

AgilOptics mirrors increase coupling efficiency into a 4 µm diameter fiber by 750%.

POCKET DEFORMABLE MIRROR FOR ADAPTIVE OPTICS APPLICATIONS

Breadboard adaptive optical system based on 109-channel PDM: technical passport

Large-aperture approximation for notso-large

Feedforward Adaptive-Optic Mitigation of Aero- Optic Disturbances

Ultra-Flat Tip-Tilt-Piston MEMS Deformable Mirror

Implementation of a waveform recovery algorithm on FPGAs using a zonal method (Hudgin)

Adaptive optic correction using microelectromechanical deformable mirrors

White-light interferometry, Hilbert transform, and noise

Deep Horizontal Atmospheric Turbulence Modeling and Simulation with a Liquid Crystal Spatial Light Modulator. *Corresponding author:

Optimization of Existing Centroiding Algorithms for Shack Hartmann Sensor

Understanding the performance of atmospheric free-space laser communications systems using coherent detection

Wavefront Sensing In Other Disciplines. 15 February 2003 Jerry Nelson, UCSC Wavefront Congress

MALA MATEEN. 1. Abstract

MAORY ADAPTIVE OPTICS

Comparative Performance of a 3-Sided and 4-Sided Pyramid Wavefront Sensor. HartSCI LLC, 2555 N. Coyote Dr. #114, Tucson, AZ

AY122A - Adaptive Optics Lab

Keck Telescope Wavefront Errors: Implications for NGAO

Implementation of decentralized active control of power transformer noise

Hartmann Sensor Manual

GROUND LAYER ADAPTIVE OPTICS AND ADVANCEMENTS IN LASER TOMOGRAPHY AT THE 6.5M MMT TELESCOPE

IAC-08-C1.8.5 OPTICAL BEAM CONTROL FOR IMAGING SPACECRAFT WITH LARGE APERTURES

Integrated Micro Machines Inc.

WIND SPEED ESTIMATION AND WIND-INDUCED NOISE REDUCTION USING A 2-CHANNEL SMALL MICROPHONE ARRAY

High-speed wavefront control using MEMS micromirrors T. G. Bifano and J. B. Stewart, Boston University [ ] Introduction

Phase Retrieval Techniques for Adaptive Optics

Spatially Varying Color Correction Matrices for Reduced Noise

MAORY E-ELT MCAO module project overview

Robust Wave-front Correction in a Small-Scale Adaptive Optics System Using a Membrane Deformable Mirror

Robo-AO: Robotic Laser Guide Star Adaptive Optics on the Palomar 60 in Christoph Baranec (PI) & Nick Law (PS)

Wavefront sensing by an aperiodic diffractive microlens array

Payload Configuration, Integration and Testing of the Deformable Mirror Demonstration Mission (DeMi) CubeSat

Experimental results of a MEMS-based adaptive optics system

Adaptive optics performance over long horizontal paths: aperture effects in multiconjugate adaptive optical systems

Digital Photographic Imaging Using MOEMS

Analysis of Hartmann testing techniques for large-sized optics

UNIVERSITÉ DE SHERBROOKE

A Prototype Wire Position Monitoring System

Ocular Shack-Hartmann sensor resolution. Dan Neal Dan Topa James Copland

Long-Range Adaptive Passive Imaging Through Turbulence

Identifying Scatter Targets in 2D Space using In Situ Phased Arrays for Guided Wave Structural Health Monitoring

Deep Horizontal Atmospheric Turbulence Modeling and Simulation with a Liquid Crystal Spatial Light Modulator. *Corresponding author:

Improving techniques for Shack-Hartmann wavefront sensing: dynamic-range and frame rate

UCLA Adaptive Optics for Extremely Large Telescopes 4 Conference Proceedings

CHARA AO Calibration Process

MULTIPLE SENSORS LENSLETS FOR SECURE DOCUMENT SCANNERS

PRELIMINARY STUDIES INTO THE REDUCTION OF DOME SEEING USING AIR CURTAINS

Adaptive-Optic Correction of a Regularized Compressible Shear Layer

Aberrations and adaptive optics for biomedical microscopes

Spatial Correlation Effects on Channel Estimation of UCA-MIMO Receivers

Chapter 2 The Test Benches

A Kalman-Filtering Approach to High Dynamic Range Imaging for Measurement Applications

Characterization of wavefront aberration in laser beam propagating over saline water and sands

PYRAMID WAVEFRONT SENSOR PERFORMANCE WITH LASER GUIDE STARS

Adaptive Notch Filter Using Real-Time Parameter Estimation

First results from the adaptive optics system from LCRD's Optical Ground Station One

Study of self-interference incoherent digital holography for the application of retinal imaging

Optimizing closed ioop adaptive optics performance using multiple control bandwidths. Brent L. Ellerbroek

Optical phase-coherent link between an optical atomic clock. and 1550 nm mode-locked lasers

Optical Correlator for Image Motion Compensation in the Focal Plane of a Satellite Camera

Adaptive optics for laser-based manufacturing processes

TIMA Lab. Research Reports

Adaptive Optics lectures

ELEC E7210: Communication Theory. Lecture 11: MIMO Systems and Space-time Communications

Receiver Signal to Noise Ratios for IPDA Lidars Using Sine-wave and Pulsed Laser Modulation and Direct Detections

Bias errors in PIV: the pixel locking effect revisited.

Tip-Tilt Correction for Astronomical Telescopes using Adaptive Control. Jim Watson

Coherent Laser Measurement and Control Beam Diagnostics

MIMO-LTI Feedback Controller Design -Status report-

Be aware that there is no universal notation for the various quantities.

Lab Report 3: Speckle Interferometry LIN PEI-YING, BAIG JOVERIA

arxiv: v1 [cs.sd] 4 Dec 2018

Effect of segmented telescope phasing errors on adaptive optics performance

In-line digital holographic interferometry

Transcription:

42nd AIAA Plasmadynamics and Lasers Conference<br>in conjunction with the<br>18th Internati 27-30 June 2011, Honolulu, Hawaii AIAA 2011-3276 Identification, Prediction and Control of Aero Optical Wavefronts in Laser Beam Propagation Jonathan Tesch and Steve Gibson Mechanical and Aerospace Engineering, University of California, Los Angeles 90095-1597 Stanislav Gordeyev and Eric Jumper Aerospace and Mechanical Engineering, University of Notre Dame, Notre Dame, IN 46556-5637 This paper describes linear time-invariant optimal prediction and control of aero-optical wavefronts derived from recent flight-test data. The optimal controller is based on an optimal prediction filter that has the form of a multi-input-multi-output state-space model that captures the statistics of the aero-optical wavefront sequence. Experimental result in the paper show the improvement in wavefront correction achieved by an optimal controller based on the identified prediction filter, as opposed to a classical adaptive optics loop. I. Introduction In recent research on adaptive optics, improved wavefront correction has been achieved through wavefront prediction by both adaptive filtering an control 1 6 and by optimal linear time-invariant (LTI) filtering and control. 9 11 For stationary or quasi-stationary turbulence, optimal LTI controllers are potentially as effective as fully adaptive controller, while being less complex and less black-box than fully adaptive controllers. As with adaptive controllers that have been used for adaptive optics (AO), optimal LTI controllers predict wavefront sequences to mitigate loop latency and the limited bandwidth of classical AO loops. Without a prediction capability, the only way to improve the bandwidth of wavefront correction is to increase the frame rate of the AO loop. This paper describes an optimal (i.e., minimum-variance) LTI controller based on a wavefront-prediction filter identified from measured wavefront data. The control and system identification methods are demonstrated by results from an adaptive optics experiment in which the disturbance wavefront sequence is derived from wavefront data obtained in recent flight tests that measured aero-optical effects on a laser beam propagating between two airplanes. The optimal prediction filters have the form of multi-input-multi-output state space models that capture the statistics of the aero-optical wavefront sequences. The prediction filter in this paper has the form of Kalman predictors, but it is not obtained by the standard methods of Kalman filter/predictor design, where known or assumed plant models along with disturbance and noise statistics are used to design a Kalman filter. Rather, the prediction filters are identified directly from measured wavefront data sequences by a subspace system identification method, with no required a priori knowledge of the turbulence statistics. A related approach to adaptive optics has been demonstrated in [10], although several aspects of the controller design and system identification are different 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-filter based subspace system identification algorithm. II. Description of the Experiment The main components of the experiment are a laser source, two deformable mirrors, a Shack-Hartmann wavefront sensor and a target camera (i.e., scoring camera). Figure 1 shows a photograph of the experiment, and Figure 2 shows a corresponding diagram. 1 of 11 Copyright 2011 by the, Inc. All rights reserved.

The control actuator is the deformable mirror denoted by DM1 in Figure 2, which is a 31-actuator oneinch membrane mirror from Active Optical Systems (AOS). The deformable mirror denoted by DM2 adds the aero-optical wavefronts to the beam; DM2 is a 61-actuator one-inch membrane mirror from AOS. The Shack-Hartmann sensor measures the wavefront on a 10 10 subaperture array. A PC running MATLAB reads the Shack-Hartmann and target-camera images, converts the Shack-Hartmann images to wavefront slope vectors, computes the control commands and sends them to DM1. The experiment runs at a frame rate of approximately 30 Hz due to hardware limitations. However, the original aero-optical flight-test data was sampled at a 16 KHz frame rate, which determines the temporal statistics of the wavefront sequence used in the experiment reported here. III. Commands to the DMs A. Control Commands The deformable mirror DM1 is identical to the DM used in [6] for adaptive control in adaptive optics. The actuator command voltages v i generate electric fields that produce DM displacements (i.e., phase shifts) d i at the individual actuators, with displacements and commands related by d i = av 2 i (1) where a is a constant. The electronic driver for the deformable mirror provides eight-bit resolution for the DM commands taking integer values between 0 and 255. For the control loops, the command voltages are parameterized as ( ) v i = round c i + vb 2, v b = 1 255 2 /2 (2) where c i is a control command and v b is an applied bias voltage. The bias voltage v b, which is the same for all actuators, produces a focus shape on the DM. The commands c i can take values in the interval vb 2 c i vb 2. If the quantization error due to the rounding is neglected, then (1) and (2) yield B. Disturbance Commands d i = a(c i + v 2 b). (3) The deformable mirror DM2 is similar to DM1 but with 61 actuators instead of the 31 actuators for DM1. The relationship among the command voltages, disturbance commands and displacements of DM2 are parameterized as in (1) (3), with the vector of disturbance commands for DM2 denoted by c w. To determine the disturbance commands for DM2, a sequence of wavefronts measured in the Notre Dame airborne aero-optics laboratory 7 were mapped to the geometry of DM2. Of course, the original wavefronts had higher resolution than the 61 degrees of freedom provided by DM2, but the dominant wavefront statistics were preserved. The measured aero-optical wavefronts also had a central hole due to a common obscuration in the optical system. Before mapping the wavefronts to the DM2 geometry, the hole was filled by a multichannel prediction filter that used both spatial and temporal information from the measured pixels to estimate the values of missing pixels. The filter gains were determined by least-squares estimation of a prediction model for the measured pixels. IV. Adaptive Optics and Optimal Control A. The Adaptive Optics Problem The block diagram in Figure 3 shows the signals and control loops for the adaptive optics problem. For control design, the following model represents the part of the block diagram between c and y: y = y b + w Γ c. (4) Here Γ is the poke matrix, which was by least-squares estimation with input/output data from a preliminary experiment in which DM1 was driven open-loop with a broadband command sequence. The classical adaptive optics (AO) loop shown in Figure 3 consists of a low-pass digital filter with gain K = 0.5 and pole α = 0.98, the modal transformation matrix U and the least-squares reconstructor E. The 2 of 11

Figure 1. Photograph of the experiment. BS 1 DM1 Aperture 500 mm (L2) Pinhole 9 mm (L1) HeNe Laser λ = 634 nm Spatial Filter 125 mm (L3) 125 mm (L4) 400 mm (L5) 50 mm (L7) 75 mm (L8) SHWFS BS 2 25 mm (L6) DM 2 Target Camera Figure 2. Diagram of the experiment. 3 of 11

low-pass filter, which is an integrator when α = 1, is a diagonal multi-input-multi-output (MIMO) filter with each diagonal element equal to the same scalar transfer function. The values of K and α were chosen through experimental trial and error with the Turbulence 1 wavefront disturbance to maximize the error-rejection bandwidth of the classical AO loop without amplifying high-frequency disturbance and noise. The z 1 block in Figure 3 indicates that the AO system has a one-frame loop latency. The columns of U represent a set of frequency-weighted deformable mirror (DM) modes that are orthogonal with respect to the actuator influence functions. These modes are discussed in detail in [6]. The mode images (i.e.; DM surface shapes) are shown in Figure 4. Both the classical AO loop and the optimal control loop use the DM modes as control channels. The reconstructor matrix E is the pseudo inverse of Γ U. Since the poke matrix Γ has linearly independent columns, E Γ U = I, (5) which implies that the modal control channels are uncoupled in the linear model of the system when only the classical AO loop is closed. U y b DM2 c w DM1 y WFS z 1 K z z α Classical AO Loop u E Ĝ(z) F (z) e Optimal Control Loop Figure 3. Block diagram of AO experiment with control loops. B. Optimal Control The optimal control loop is shown inside the dashed box in Figure 3. The optimal control loop augments the classical AO loop, which remains closed when the optimal control loop is closed. As shown in these block diagrams, the optimal controller is a feedback controller but its structure is based on a feedforward disturbance-rejection design with the optimal prediction filter F (z), discussed in Section V. For the optimal controller, the input is the wavefront error signal e and the output is the optimal control command u, both of which are in DM modal coordinates. The transfer function Ĝ(z) in Figure 3 is a model of the true transfer function G(z) from u to e with only the classical AO loop closed. The optimal controller assumes that (4) and (5) hold, and this implies that G(z) is a diagonal linear transfer function with each diagonal term given by K Ĝ(z) = z + K α. (6) In hardware applications, G(z) and Ĝ(z) probably never are exactly equal, and often G(z) is nonlinear. 4 of 11

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 4. Deformable mirror modes. V. Minimum-Variance Wavefront Prediction and System Identification The one-step prediction filter has the state-space form x(t + 1) = Ax(t) + K[y(t) ŷ(t)] ŷ(t) = Cx(t) (7) where y(t) is a vectorized wavefront sequence, as represented by the modal coefficients, and t is the frame or sample number. The one-step prediction of y(t) is ŷ(t). The the relationship of the mathematical state vector x(t) to the physics of the wavefront sequence is convoluted, so that the components of x(t) (usually numbering in the hundreds) should be considered as just internal variables of the prediction model. The constant matrices A, C and K are identified by a subspace system identification method from a representative wavefront sequence. The multi-channel lattice-filter based subspace identification algorithm in 8 was used. The methods described here can be generalized to multi-step prediction. Most AO applications require either one or two-step prediction. The statistics of the wavefront sequence are represented implicitly in the matrices A, C and K and the covariance matrix for the prediction-error or innovations sequence The one-step prediction model can be written in the innovations form e = y(t) ŷ(t) (8) x(t + 1) = Ax(t) + Ke(t) y(t) = Cx(t) + e(t) (9) 5 of 11

Theoretically, with sufficiently long, stationary data sequences, the prediction error (i.e., the innovations sequence) is white, and the prediction filter becomes a steady-state Kalman filter. But the filter is not constructed in the classical way from a known plant model and assumed disturbance statistics, neither of which is known or assumed here. For the experimental results in this paper, a 7000 frames of the experimentally measured aero-optical data were mapped to the DM2 geometry. The first 5000 frames were used to identify the A, C and K in (7). The remaining 2000 frames were used for the disturbance sequence in the experiment. VI. Experimental Results A. Performance Enhancements with Optimal Control Figures 5 through 8 and Table 1 depict experimental results from applying turbulent wavefronts to the setup in Figure 2. A sequence of 7000 disturbance commands for DM2 were generated from the Notre Dame data. The first 5000 frames were used for identification of an optimal LTI controller, while the remaining 2000 were used in the experiment to compare controller performance. Three control schemes were implemented. No correction was applied to the wavefront during the openloop experiment, while only the classical AO loop shown in Figure 3 was closed for the integrator experiment. For the third experiment, an optimal LTI controller was first identified using the 5000 frame disturbance sequence with the classical AO loop closed. Then, the 2000 frame sequence was applied with the optimal LTI loop augmenting the classical loop as in Figure 3. Figure 6 shows a portion of the RMS wavefront error time series for each of these experiments. While the classical AO loop addresses some of the open-loop error, the optimal LTI predictor further reduced the error at the majority of time steps. Table 1 summarizes the RMS wavefront error over time and space; while the classical AO loop reduced the open-loop RMS error by 22.5%, using the optimal controller resulted in a 46.5% reduction. Mitigating the wavefront error resulted in an improvement in the maximum target intensity. Figure 5 shows the mean target intensity over 2000 frames for each experiment, where the wavefront error reduction with the optimal controller produced a significant increase in the peak average intensity compared to the classical controller. Part of this improvement can be attributed to the reduced movement in the peak location; shown in Table 1, the standard deviation of location of the maximum intensity dropped when the optimal controller was used. However at many samples the maximum value itself increased, as seen in the time series of the maximum intensity in Figure 6. B. Modal Analysis Further performance results can be seen by projecting the residual wavefronts from the experiments onto the DM modes in Figure 4, resulting in a time series of modal coefficients for each experiment. Figure 7 shows the power of the measured wavefronts in each mode (i.e. the temporal RMS of each modal sequence), and also the modal power for the classical and optimal cases normalized by the open-loop power. 15 of the 31 modes were used for control in both the classical and optimal loops, evidenced by the increase in the modal power in modes 16 through 31, relative to the open-loop case. Figure 8 shows the power spectral densities for several representative modal sequences. The classical AO controller succeeds in reducing disturbances until approximately 1.5-2 khz. However the optimal controller is effective for a considerably wider range of frequencies, showing significant improvement until 4 khz in some modes. Acknowledgments Jon Tesch and Steve Gibson were supported by the High Energy Laser Joint Technology Office (HEL JTO) and the U.S. Office of Naval Research, ONR Grant number N00014 07-1-1063. Stanislav Gordeyev and Eric Jumperwere supported by the High Energy Laser Joint Technology Office (HEL JTO) and the U.S. Air Force Office of Scientific Research, AFOSR Grant number FA9550-07-1-0574. 6 of 11

Performance measures for correction of wavefront error and improvement of on-target energy distri- Table 1. bution. Wavefront error Peak intensity Standard deviation Control Type RMS (µm) over of average of peak intensity position time and space target image horizontal vertical Open Loop 0.66 9.07 13.82 9.04 Classical AO Loop 0.51 16.53 9.97 5.31 Optimal Control 0.35 21.66 7.84 3.72 References 1 J. S. Gibson, C.-C. Chang, and B. L. Ellerbroek, Adaptive optics: wavefront correction by use of adaptive filtering and control, Applied Optics, Optical Technology and Biomedical Optics 39, 2525 2538 (2000). 2 Y.-T. Liu and J. S. Gibson, Adaptive optics with adaptive filtering and control, in American Control Conference, (Boston, MA, 2004), pp. 3176 3179. 3 T. A. Rhoadarmer, L. M. Klein, J. S. Gibson, N. Chen, and Y.-T. Liu, Adaptive control and filtering for closed-loop adaptive-optical wavefront reconstruction, in SPIE Conference on Advanced Wavefront Control, vol. 6306, (SPIE, San Diego, CA, 2006), pp. 63060E 1 63060E 12. 4 J. S. Gibson and Yu-tai Liu, Adaptive control in adaptive optics for directed energy systems, Optical Engineering 46, 046601 1 046601 13 (2007). 5 Salman Monirabbasi and J. S. Gibson, Adaptive control in an adaptive optics experiment with simulated turbulenceinduced optical wavefronts, (SPIE, San Diego, CA, 2009), vol. 7466 of SPIE Conference on Advanced Wavefront Control, pp. 746608 1 746608 11. 6 Salman Monirabbasi and Steve Gibson, Adaptive control in an adaptive optics experiment, J. Opt. Soc. Am. A 27, 84 96 (2010). 7 C. Porter, S. Gordeyev, M. Zenk and E. Jumper, Flight Measurements of Aero-Optical Distortions from a Flat-Windowed Turret on the Airborne Aero-Optics Laboratory (AAOL), in 42nd AIAA Plasmadynamics and Lasers Conference, Honolulu, HI, June 2011. 8 N. Chen and J. S. Gibson, Subspace system identifcation using a multichannel lattice filter, in American Control Conference, (Boston, MA, 2004), pp. 855 860. 9 C. Petit, J.-M. Conan, C. Kulcsár, H.-F. Raynaud, T. Fusco, J. Montri, F. Chemla, and D. Rabaud, Off-axis adaptive optics with optimal control: Experimental and numerical validation, (2005), vol. 5903 of Proc. SPIE, pp. 227 235. 10 K. Hinnen, M. Verhaegen, and N. Doelman, Exploiting the spatio-temporal correlation in adaptive optics using datadriven H 2 -optimal control, J. Opt. Soc. Am. A 24, 1714 1725 (2007). 11 C. Petit, J.-M. Conan, and C. Kulcsár, LQG control for AO and MCAO: experimental and numerical analysis, J. Opt. Soc. Am. A 26, 1307 1325 (2009). 7 of 11

Figure 5. Average target camera images. 8 of 11

1 Open Loop Classical AO Optimal Control 0.8 RMS (µm) 0.6 0.4 0.2 0 300 320 340 360 3 400 420 440 460 4 500 t 250 200 Open Loop Classical AO Optimal Control Maximum Intensity 150 50 0 300 320 340 360 3 400 420 440 460 4 500 t Figure 6. RMS and peak. 9 of 11

RMS 200 150 50 Temporal RMS Open Loop Classical LTI RMS LTI /RMS OL 0 0 5 10 15 20 25 30 35 Mode Temporal RMS Reduction 1.5 Classical LTI 1 0.5 0 0 5 10 15 20 25 30 35 Mode Figure 7. Modal power. 10 of 11

1 Mode 1 Open Loop Classical LTI 1 Mode 2 0 2000 4000 6000 00 1 Mode 4 0 2000 4000 6000 00 1 Mode 10 0 2000 4000 6000 00 0 2000 4000 6000 00 1 Mode 5 0 2000 4000 6000 00 1 Mode 25 0 2000 4000 6000 00 Figure 8. Power spectral densities of representative modal sequences. 11 of 11