5:: 5-: LNMWJq.rH6 . .nrvvqsfl.... a... .. «flung. , RM? ”when“: .0 t . NM... «5.: .. at... i .l. 3.5.5:! nm... wfi .‘hfit . l 3 {JV .1. :19. i )1 . .V S . a. P9 4 L‘ , u! 2. ‘11. I... ’1 A. t I. 5.. mm“. .. .I a: .I ‘. _. .931 $2.. 5 T; x :3 {Manama {fir . 31‘ :5.H 1‘ ‘2 , n33? THESIS at: @9125)” This is to certify that the dissertation entitled INVESTIGATION OF THE WALL-SHEAR-STRESS SIGNATURE IN A BACKWARD-FACING—STEP FLOW USING OSCILLATING HOT-WIRE SENSORS presented by YONGXIANG Ll has been accepted towards fulfillment of the requirements for the PhD. degree in Mechanical Emeerirfl WWW Major Professor’s Signature B/Za/o 4f- Date MSU is an Affirmative Action/Equal Opportunity Institution LIBRARY Michigan State University PLACE IN RETURN Box to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 6/01 c'JCIRCJDateDuesz-sz INVESTIGATION OF THE WALL-SHEAR-STRESS SIGNATURE IN A BACKWARD-FACING—STEP FLOW USING OSCILLATING HOT-WIRE SENSORS By Yongxiang Li A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Mechanical Engineering 2004 ABSTRACT INVESTIGATION OF THE WALL-SHEAR-STRESS SIGNATURE IN A BACKWARD-FACING—STEP FLOW USING OSCILLATING HOT-WIRE SENSORS By Yongxiang Li A new high—fiequency oscillating-hot-wire (OHW) sensor for measurement of the wall-shear—stress magnitude and direction was developed for the purpose of studying the wall-shear-stress field behind an axi-symmetric backward-facing step (BFS). The stress direction is determined from the phase angle between the imposed and measured oscillation velocity, and the corresponding magnitude is obtained from the low-pass filtered signal of the sensor, after removal of the modulating influence of the oscillation. Two OHW sensors with oscillation frequency of 2.8 kHz were utilized to capture and investigate the space-time characteristics of the wall-shear signature behind the BFS flow at three Reynolds numbers of 4300, 8700 and 13000 (based on step height). One of the sensors was fixed at approximately 5H (where, H is the step height) downstream of the step, while the other sensor was moved to acquire data at 14 different streamwise (x) locations ranging from 0.3H to 10H. The results agreed qualitatively with existing one- point measurements, such as the mean/rms skin-fiiction distribution, forward flow probability, and power spectra, in planar BF S flows. However, some fundamental quantitative differences were found including a shorter reattachment length and a larger x location for the peak rms wall-shear fluctuations. These differences were attributed to the axi-symmetric nature of the present geometry, transverse curvature of the step, or differences in the measurement methods. Consistent with the wall-pressure literature in similar type of flows, the power spectra of the fluctuating wall-shear stress revealed the existence of two characteristic frequencies: f = 0.1 and 0.65. The former is associated with low-frequency shear-layer flapping and the latter corresponds to the passage of the separated shear-layer vortex structures. The downstream convection velocity of the vortices was inferred from the two-point cross-correlation of the wall-shear fluctuations. The resulting value was approximately 0.45 of the fi'eestream velocity, which falls below the range reported in the wall-pressure literature of 0.5 — 0.6 of the freestream velocity, in investigations of various separating/reattaching flows in planar configurations. The lower value is apparently a result of the deceleration of the vortices as they approach the mean reattachment point by the stronger adverse streamwise pressure gradient associated with mean-pressure recovery over the shorter reattachment distance in the axi-symmelric geometry. ACKNOWLEDGEMENTS I would like to express my deepest gratitude to my advisor Dr. Ahmed Naguib for giving me an opportunity to work on this project and for his support and encouragement throughout my graduate studies, for his help in my professional and personal growth, and for his thorough review of the manuscripts of this thesis. I would also like to thank other committee members Dr. Dean Aslam, Dr. Brage Golding, Dr. John Foss, and Dr. Manoochehr Koochesfahani for many helpful comments and discussions of this thesis. I appreciate the help from my fellow graduate students. Thanks are due to Laura Michele Hudy for the great help and contribution in designing, aligning the experimental model and acquiring the boundary layer data. Antonius Aditj andra has been a great help during the data acquisition. I have also had many valuable discussions with Mohamed Daoud. I am indebted to my father, my mother, my brother and sister for their support through all these years. Finally, my ultimate thanks go to my wife Yuehui (Christine ) Qin for her understanding and love. Her support and encouragement was in the end what made this dissertation possible. This study was made possible through fimding from the National Science Foundation under grant number CTS-0116907, monitored by Dr. Michael W. Plesniak, director of the Fluid Dynamics and Hydraulics program. iv TABLE OF CONTENTS LIST OF TABLES .................................................................................................. vii LIST OF FIGURES ............................................................................................... viii LIST OF SYMBOLS ............................................................................................ xiii CHAPTER 1 ................................................................................................................ 1 INTRODUCTION .................................................................................................... 1 1.1 Background ......................................................................................................... 1 1.2 Literature Review ................................................................................................ 4 1 .3 Motivation ......................................................................................................... 1 5 1.4 Objectives ......................................................................................................... 16 CHAPTER 2 .............................................................................................................. 18 EXPERIMENTAL SETUP ................................................................................ 18 2.1 Wind Tunnel ..................................................................................................... 18 2.2 Axi-symmetric Backward-facing—step Model ................................................... 20 2.3 Boundary Layer Profiles ................................................................................... 32 2.4 Couette Flow Calibration Facility ..................................................................... 34 CHAPTER 3 .............................................................................................................. 42 OSCILLATING-HOT—WIRE (OHW) TECHNIQUE .......................... 42 3.1 Review of Techniques for Measurement of Direction-reversing Wall-shear Stress 42 3.2 OHW Principle .................................................................................................. 46 3.3 OHW Response ................................................................................................. 53 3.4 OHW Measurements (Phase Method and Phase Calibration) .......................... 61 3.5 Effect of Wire Height above the Wall .............................................................. 65 3.6 Effect of Wire Oscillation on the Flow ............................................................. 69 3.7 Specification of OHW Sensors Use in the Present Experiments ...................... 74 CHAPTER 4 .............................................................................................................. 78 ONE-POINT MEASUREMENT RESULTS ............................................ 78 4.1 Boundary-layer Velocity Profiles at Separation ............................................... 78 4.2 One-point Wall-shear Measurements in the Separated/reattaching Flow ......... 85 4.2.1 Mean Reattachment Length ...................................................................... 85 4.2.2 Mean/rms Skin-fiiction Coefficient and FFP ........................................... 90 4.2.3 Probability Density Function (pdf) ......................................................... 100 4.2.4 Power Spectra ......................................................................................... 103 4.2.5 Temporal Autocorrelation Function ....................................................... 110 CHAPTER 5 ............................................................................................................ 116 TWO-POINT MEASUREMENT RESULTS ........................................ 116 5.1 Two-point Cross-correlation ........................................................................... 1 16 5.1.1 Two-point Cross-correlation with the Reference Sensor Located near x, 1 16 5.1.2 Filtered Two-point Cross-correlation with the Reference Sensor Located near x, 127 5.1.3 Two-point Cross-correlation with the Reference Sensor Located near 0.5x, 132 5.2 Frequency Dependence of the Convection Velocities .................................... 137 5.3 Additional Discussion ..................................................................................... 146 CHAPTER 6 ............................................................................................................ 153 CONCLUSIONS AND RECOMMENDATIONS ............................... 153 6.1 Conclusions ..................................................................................................... 153 6.2 Recommendations ........................................................................................... 1 5 8 BIBLIOGRAPHY ................................................................................................. 161 vi LIST OF TABLES Table 3.1 Combinations of the sensors and spacers corresponding to different measurement locations of the movable probe ........................................................... 76 Table 3.2 Measurement positions of the movable sensor ............................................... 77 Table 4.1 Parameters for the boundary layer at separation ............................................. 79 Table 4.2 Reattachment length values ............................................................................ 87 vii LIST OF FIGURES Figure 1.1 Backward-facing-step flow structure .............................................................. 2 Figure 2.1 Schematic diagram of the wind tunnel (dimensions in mm) ......................... 19 Figure 2.2 Picture of the axi-symmetric backward-facing—step model ........................... 21 Figure 2.3 Schematic of the BF S model with dimensions (/in mm) ............................... 22 Figure 2.4 Piano-wires and boundary layer trip downstream of the model nose ........... 23 Figure 2.5 Magnified view of the measurement zone showing wall-sensor covers: (a) 3D view, (b) Cross-section view ..................................................................................... 25 Figure 2.6 Picture of the sensor cover containing two OHW sensors ............................ 26 Figure 2.7 Static pressure distribution downstream of the step at four azimuthal locations and U00 = 15 m/s ......................................................................................... 27 Figure 2.8 Picture of the setup for checking azimuthal symmetry of the flow ............... 28 Figure 2.9 Azimuthal symmetry of the flow at separation for U... = 15 m/s ................... 29 Figure 2.10 Freestrearn static pressure distribution upstream of the step ....................... 31 Figure 2.11 Typical image of the probe tip and step area used for hot-wire positioning near the wall: (a) Wire at the initial position, (b) Wire at 0.5 mm above initial position ................................................................................................................ , ...... 33 Figure 2.12 Picture of the Couette Flow Facility ............................................................ 36 Figure 2.13 Cross Section of the Couette Flow Facility (dimensions in mm) ................ 36 Figure 2.14 Image of the gap between the rotating and stationary discs ....................... 38 Figure 2.15 Comparison between DNS results and Couette flow assumption ............... 41 Figure 3.1 Schematic diagram illustrating the OHW sensor concept ............................. 46 Figure 3.2 Schematic diagram illustrating OHW sensor response for different flow directions: (a) Flow direction is towards the right, (b) Flow direction is towards the lefi ............................................................................................................................. 48 viii Figure 3.3 Schematic of a two-layer bending piezo-element ......................................... 48 Figure 3.4 Picture of OHW sensor plug .......................................................................... 49 Figure 3.5 Cross-section view of OHW sensor plug (dimensions in mm) ..................... 50 Figure 3.6 Sketch showing wire connections to the OHW sensor .................................. 51 Figure 3.7 Video image sequences of the hot wire during oscillation: (a) Wire is straight during oscillation, (b) Wire bends during oscillation ............................................... 55 Figure 3.8 Picture of calibration-facility adaptor plug .................................................... 56 Figure 3.9 Dimensions of adapter plug for calibration of OHW sensors (dimensions in mm) ........................................................................................................................... 56 Figure 3.10 Magnitude calibration of the hot wire with/without oscillation ................... 57 Figure 3.11 Phase averaged hot-wire output at the oscillation frequency ....................... 58 Figure 3.12 Demonstration of the effect of the flow direction on the phase of the OHW modulation signal: (a) Modulated velocity signals, (b) OHW driving signals ......... 60 Figure 3.13 A typical direction calibration of the OHW sensor ...................................... 64 Figure 3.14 OHW Operation schematic diagram ............................................................ 65 Figure 3.15 Power spectra of three different runs for wire height y = 48 um ................. 67 Figure 3.16 Histogram of three different runs for wire height y = 48 um ....................... 68 Figure 3.17 Power spectra for wire heights y = 48, 96 and 124 um ................................ 68 Figure 3.18 Histogram for wire heights y = 48, 96 and 124 um ..................................... 69 Figure 3.19 Comparison of power spectra of OHW & conventional hot wire (HW) at different x positions: (a) x/x, = 0.06, (b) x/x, = 0.92, (c) x/x, = 1.54 .......................... 72 Figure 3.20 Comparison of histogram of OHW & conventional hot wire (HW) at different x positions: (a) x/x, = 0.06, (b) x/x, = 0.92, (c) x/x, = 1.54 .......................... 74 Figure 3.21 One of the combinations of the sensor cover (in mm) ................................. 75 Figure 3.22 Picture of the sensor cover during insertion in the test mode] ..................... 76 ix Figure 4.1 Mean velocity profiles of the boundary layer at separation .......................... 80 Figure 4.2 Comparison of a mean velocity profile from the current study to that of turbulent and laminar boundary layers. .................................................................... 81 Figure 4.3 Inner—scaled Mean velocity profiles of the boundary layer at separation ..... 84 Figure 4.4 Rms velocity profile of the boundary layer at separation .............................. 84 Figure 4.5 Streamwise distribution of the mean skin-friction coefficient .............. 85 Figure 4.6 Streamwise distribution of forward flow probability .................................... 87 Figure 4.7 Comparison of the mean skin friction coefficient results with other studies: (a) ReH< 10000, (b) Rey> 10000 ............................................................................. 92 F igurc 4.8 Minimum skin-fiiction coefficient versus Re” .............................................. 94 Figure 4.9 Comparison of forward flow probability results with other studies .............. 95 Figure 4.10 Comparison of ms skin-friction coefficient results with other studies: (a) ReH< 10000, (b) ReH> 10000 ................................................................................... 97 Figure 4.11 Wall-shear pdf at different x positions: (a) ReH= 4300, (b) ReH= 8700, (c) Re” = 13000 ............................................................................................................ 102 Figure 4.12 Wall-shear power spectra at different x locations: (a) ReH= 4300, (b) Rey= 8700, (c) Re” = 13000 ............................................................................................. 106 Figure 4.13 Power spectra of the fluctuating wall pressure at different x locations (Hudy (2003), private communication) .............................................................................. 106 Figure 4.14 Wall-shear power spectra plotted on semil-log scale for different x locations: (a) Rey= 4300, (b) Rey= 8700, (c) Re” = 13000 ................................................... 108 Figure 4.15 Wall-shear power spectra plotted on semi-log scale at x/xr z 0.07 ............ 109 Figure 4.16 Autocorrelation coefficient of the fluctuating wall-shear stress at different x locations: (a) ReH= 4300, (b) ReH= 8700, (c) Re” = 13000 ................................... 1 12 Figure 4.17 Autocorrelation coefficient of the fluctuating wall-shear stress at x/xr z0.07 ................................................................................................................................. 112 Figure 4.18 Contour maps of the auto-correlation coefficient of the fluctuating wall- shear stress: (a) ReH= 4300, (b) ReH= 8700, (c) Re” = 13000 ............................... 1 15 Figure 5.1 Cross-correlation coefficient of the fluctuating wall-shear stress with the reference sensor located near x,: (a) ReH= 4300, (b) ReH= 8700, (c) Re” = 13000120 Figure 5.2 Contour maps of the cross-correlation coefficient of the fluctuating wall- shear stress with the reference sensor located near x,: (a) ReH= 4300, (b) Re”: 8700, (c) Re” = 13000 ............................................................................................. 122 Figure 5.3 Correlation-peak time delay at different x location and associated line fits 125 Figure 5.4 “Local” convection velocity at different x positions ................................... 126 Figure 5.5 Local convection velocity and pressure coefficient versus strearnwise locations .................................................................................................................. 127 Figure 5.6 Cross-correlation coefficient of the low-pass filtered fluctuating wall-shear stress with reference sensor located near x,: (a) Re” = 4300, (b) ReH= 8700, (c) Re” = 13000 ................................................................................................................... 130 Figure 5.7 Contour maps of the cross-correlation coefficient of the low-pass filtered fluctuating wall—shear stress with reference sensor located near x,: (a) Rey= 4300, (b) ReH= 8700, (0) Re” = 13000 ............................................................................. 132 Figure 5.8 Cross-correlation coefficient of the fluctuating wall-shear stress with the reference sensor located near 0.5x,: (a) Re”: 4300, (b) Re”: 8700, (c) Re” = 13000 ................................................................................................................................. 135 Figure 5.9 Contour maps of the cross-correlation coefficient of the fluctuating wall- shear stress with reference sensor located near 0.5x,: (a) ReH= 4300, (b) ReH= 8700, (0) Re” = 13000 ....................................................................................................... 136 Figure 5.10 Coherence of the fluctuating wall-shear stress with the reference sensor located near x,: (a) ReH= 4300, (b) ReH= 8700, (c) Rey = 13000 .......................... 141 Figure 5.11 Streamwise dependence of the phase angle at selected frequencies: (a) Re”= 4300, (b) Re”: 8700, (c) Re” = 13000 ................................................................... 144 Figure 5.12 Convection velocities obtained from phase-angle analysis ........................ 145 Figure 5.13 Comparison of near-wall turbulence-intensity profiles between measurement results and one-dimension model prediction (Devenport and Sutton (1991)) ........ 148 Figure 5.14 Comparison between wall shear and pressure power spectra .................... 150 Figure 5.15 Comparsion of ms skin-fiiction coefficient from the unfiltered and high- pass-filtered signals ................................................................................................. 151 xi Figure 5.16 Percent 1’“, energy associated with vortex passage .................................... 152 IMAGES IN THIS DISSERTATION ARE PRESENTED IN COLOR. xii 3.9.. \\ LIST OF SYMBOLS deflection of the piezo-element amplitude of wall-pressure—gradient fluctuations amplitude of traveling wave amplitude of piezo-element driving voltage capacitance of the piezo-element mean skin-friction coefficient minimum mean skin-friction coefficient rms skin-friction coefficient high-pass-filtered rms skin-friction coefficient meaerressure coefficient diameter of hot-wire sensor diameter of the separated-shear-layer vortices frequency normalized frequency (f' = fx,./ U 00 ) gap between the two discs of the Couette-flow facility step height half the height of the fence in fence/splitter-plate flow boundary-layer shape factor current through the piezo-element wavenumber sensing length of the hot-wire sensor xiii ’ns m¢ Poo Il‘l f). . I1172 I’. i 7272 rw Rz’r’ I? 1. 7172 lie» 1Q611 length of the piezo-element slope of the correlation-peak locus slope of phase-angle plot instantaneous pressure reference mean pressure mean surface pressure power consumption of the piezo-element power spectra normalized power spectrum, P; = Pxx /(l/2 p U 020 ) power spectrum of 1w from the reference sensor 0 cross-spectrum of rw between the reference and movable sensors power spectrum of 1;, from the movable sensor radius of the rotating disc of the Couette-fow facility radial position of the hot-wire sensor in the Couette-flow facility model radius autocorrelation coefficient cross-correlation coefficient 9.122 Reynolds number based on gap size between two discs, Re h = v Reynolds number based on step height, Re” = U ,0 H/v xiv Reg Reynolds number based on momentum thickness of the boundary layer at separation, Ree = U do 6/v t time u instantaneous streamwise velocity u, flow velocity um measured oscillation velocity u, friction velocity (u, = ./ rw / p ) u' fluctuating streamwise velocity u'_v' shear Reynolds stress U mean streamwise velocity Uc convection velocity U0 oscillation velocity U 00 freestream velocity U+ inner-scaled mean velocity (U+ =U/u ,) fluctuating normal velocity V voltage across the piezo-element VM forward measurement velocity of an OHW sensor VIM backward measurement velocity of an OHW sensor V, radial velocity component in the Couette-flow facility Va azimuthal velocity component in the Couette-flow facility Vy wall-normal velocity component in the Couette-flow facility W width of the piezo-element XV Twl TwZ streamwise coordinate (measured from the step) mean reattachment length Fourier transform of the reference sensor signal conjugate of the Fourier transform of the movable sensor signal distance normal to the wall y in wall units (y+ = yu, /v) spanwise coordinate proportionality constant between Cm," and Rey“, CM", = orRef"5 boundary-layer thickness boundary-layer displacement thickness oscillating boundary-layer thickness, 50 = 2—V a) _ 1/2 , 2 normalized oscillating boundary-layer thickness, 6+ = (TI—0H1“, /p] momentum thickness of the boundary layer at separation time delay dynamic viscosity kinematic viscosity fluid density mean wall-shear stress wall-shear stress obtained from the reference sensor wall-shear stress obtained from the movable sensor xvi ¢ohw fluctuating wall-shear stress I phase difference of rwbetween the movable and reference sensors phase lag between the oscillating-hot-wire output and the piezo-driving voltage angular frequency spanwise vorticity wall-normal vorticity coherence VI I -r'tfi streamwise separation distance between the two sesnors initial phase angle ‘ r rotational angular velocity xvii CHAPTER 1 INTRODUCTION 1.1 Background The importance of separating and reattaching flows in practical engineering applications and research of fundamental fluid mechanics is well known. These flows have been extensively studied using various experimental and numerical techniques during the past decades. Some recent investigations on the topic include: wall-shear measurements by Tihon et al. (2001) and Spazzini et al. (2001), wall-pressure measurements by Hudy et al. (2003) and Lee and Sung (2002), laser Doppler velocimetry (LDV) by Rinoie et al. (2002), particle image velocirnetry (PIV) by Kostas et al. (2002) and Scarano et al. (1999), direct numerical simulation (DNS) by Le et al. (1997), and large eddy simulation (LES) by Fureby (1999). A backward-facing step (BFS) is one of the simplest geometries in this class of flows. However, the flow structure downstream of the step is rather complicated. As Shown in Figure 1.1, there are four main regions in the flow field behind the step: a separated free shear layer, primary and secondary recirculation regions, and a redeveloping boundary layer. The boundary layer upstream of the step separates at the step edge due to the discontinuity in geometry. The separated flow has low-speed fluid on one side and high- speed fluid on the other side. According to the Kelvin-Hehnoltz instability theory, such velocity variation across the thin shear layer result in the formation of spanwise vortices downstream of the step edge. The vortices convect downstream and grow in size via pairing and entrainment and, eventually, impinge on the wall before continuing to travel downstream. The formation and growth of the vortices has been visualized by many researchers and, recently, detected in the PIV measurements by Scarano et a1. (1999) and Kostas et al. (2002). Separated free _ _ _ shear layer 13“”de streamline I Redeveloping l/l/l/I/l/fl/ boundarVIaVCI' / Hfi / i i \ 1 —-> x Secondary Primary Mean reattachment recirculation recirculation point region region Figure 1.1 Backward-facing-step flow structure The reattachment of the separated shear layer leads to a primary recirculation region, where the flow is highly turbulent and unsteady. The reattachment mechanism is driven by the need to balance the entrainment flow into the separated shear layer. In particular, the curvature of the streamlines associated with the reattachment process produces an adverse pressure gradient that drives the backflow into the recirculation region to provide the necessary flow for this balance. The curvature of the streamlines was eliminated in the experiments of Morris (2002) to establish a single stream shear layer by supplying the entrainment flow externally to create a zero pressure gradient in the free stream for the separated shear layer to grow. In the back-step flow, the recirculation region is distinguished from the separating Shear layer by a dividing streamline, which is a contour of the time-mean stream fimction originating from the limiting streamline at the step and terminating at the wall as shown in Figure 1.1. The reattachment point is defined as the position where the dividing streamline ends on the wall. It is generally understood that instantaneously the reattachment point moves upstream and downstream due to the ‘flapping’ motion of the separated shear layer as discussed in Eaton and Johnston (1981), Driver et al. (1987), Spazzini et al. (2001), Lee and Sung (2002) and Hudy et al. (2003), amongst others. Just downstream of the step, there is a secondary recirculation region or corner eddy, which was observed, for example, from mean streamline plots in the particle- fi'acking-velocimetry measurements in a BF S flow by Kasagi and Matsunaga (1995) and PIV measurements by Kostas et al. (2002). The plots also demonstrated that the velocities in this region of the flow were very small. Finally, a new sub-boundary-layer begins to develop downstream of the reattachment point. J ovic (1998) investigated the turbulence structure in the redeveloping boundary layer region in a BFS flow fi'om the measurements of mean static pressure, Skin-friction and velocity. He found that the near-wall boundary layer reached an equilibrium turbulent boundary layer by 15 step heights away from the step, and the eddy viscosity in the outer layer recovered to the equilibrium value more than 100 step heights downstream of the step. Although the primary interest of the current study is the near-wall flow region, the following literature review will first summarize the flow structure behind a BFS obtained fi'om flow field measurements with emphasis on more recent studies employing techniques that yield instantaneous spatial information such as PIV and sensor arrays. This is done in order to provide the reader with the background necessary to draw links between the wall-shear measurements done here and the dominant flow features in separating/reattaching flows. Next, the discussion will focus on the flow information extracted from the most recent studies of the near-wall region, including those based wall-pressure as well as wall-shear. 1.2 Literature Review Eaton and Johnston (1981) provided a comprehensive review of the mean reattachment length (x,), structure of the separated shear layer and turbulence data above the wall for the BFS flow from the literature available at that time. The mean reattachment length is one of the most important parameters and has been used as a length scale to normalize the streamwise position in correlations of the mean-pressure coefficient, skin-fiiction coefficient and probability of flow direction. The values of x, reported in almost all experiments fall in a range which extends fiom 4.9 through 8.2 step heights. According to Eaton and Johnston (1981), x, could be affected by any of the following parameters: initial boundary-layer state (laminar or turbulent), initial boundary- layer thickness, freestream turbulence, pressure gradient and aspect ratio (step width over step height). Eaton and Johnston (1981) concluded that, for a fully developed turbulent boundary layer, x, is generally 5.5-7.0 step heights, independent of the Reynolds number. This was confirmed later in the study by Armaly et al. (1983). According to the studies by de Brederode and Bradshaw (1972), the effect of the aspect ratio could be neglected if it is larger than 10. The other two parameters (boundary-layer thickness and freestream turbulence) have an adverse effect on x,, That is, x, decreases with the increasing boundary-layer thickness and freestream turbulence. In addition to the above influences, the blockage ratio (the area ratio of the model over the test section of a wind tunnel) should also affect x, for test models suspended in a unifrom stream such as the axi-symmetric BFS used in the current study (Figure 2.2 and Figure 2.3 display the model details with associated discussion provided in section 2.2). To predict x, and other aspects of the back-step flow, researchers have dedicated special attention to studying the structures. Eaton and Johnston (1981) found that the structure of the separated shear layer upstream of the reattachment zone was very similar to a plane mixing-layer, where large-eddy motion resulted fiom the roll-up of the shear layer into spanwise vortices and multiple pairings of these vortices. This picture of the flow structure could be drawn more accurately with recent advances in PIV, which yields multi-dimensional, quantitative data of the spatial flow structure. Scarano and Riethmuller (1999) investigated the instantaneous two-dimensional velocity distribution using digital PIV measurements of the flow over a BFS at Reynolds number (based on step height H) of 5000. The vector plot of the instantaneous velocity field suggested that the vortex formation started at x/H =1. Scarano et al. (1999) further analyzed their data using a pattern recognition algorithm to characterize the spatial occurrence and size of the spanwise vortices in and around the shear layer. From the statistical characteristics of the Spatial occurrences, they detected clockwise vortices with vortex diameter (d.) as small as 0.12H and attributed the origin of these vortices to the Kelvin-Helmoltz instability of the plane shear layer. The vortices approached the bottom wall at about x/x,=0.58, then continued to travel in the streamwise direction. Surprisingly, Scarano et al. (1999) also found significant number of counter- clockwise rotating vortices with size of d/H = 0.12 in the free shear layer downstream of x/x, = 0.67. They attributed the generation of these vortices to the three-dimensional instability and breakdown of the primary clockwise rotating vortices. The vortex characteristics were further analyzed through application of conditional averaging techniques to the results of the pattern recognition process. The conditionally averaged flow field exhibited elliptic and quasi-circular streamline patterns at x/H of 1.4 and 7, respectively, which proved the existence of the expected swirling motion associated with the spanwise vortices. The peak vorticity at the vortex core was also found to be 3 - 5 times larger than the mean vorticity level in the region 5 < x/H < 8. Kostas et al. (2002) performed another PIV investigation of a two-dimensional BFS flow at Reynolds ntunber, based on step height and fiee stream velocity, of 4660. Their work not only Showed the measurements of velocity and vorticity in the x-y and x-z planes, but also examined the contribution of the various length scales of the flow structure to the in-plane Reynolds stress (u'v') and turbulent kinetic energy dirstribution within and around the separation bubble. The instantaneous vorticity field depicted vortical structures in both staggered (i.e. the vortices are discretely distributed) and in- line (i.e. the vortices are next to each other like a train of vortices) configurations, illustrating the dominance of vortices in the flow and their mutual interaction. Measurements of the Reynolds stress and turbulent kinetic energy production showed that the vortex interactions occurred at locations of the peak values of turbulent kinetic energy production and Reynolds stresses. Upstream of reattachment, small scales were found to be responsible for the Reynolds stress and turbulent kinetic energy production. On the contrary, the Reynolds stress and turbulent kinetic energy production were predominantly affected by the large-scale structures downstream of reattachment. Kostas et al. (2002) also used the mean vorticity distribution to identify the separated shear layer by locating the region of concentrated negative spanwise vorticiy (0);). In this manner, the shear layer growth could be found from the width of the contour plot of the mean vorticity distribution in the x-y plane. As anticipated, this showed the shear layer to be very thin immediately downstream of the step, but it continued to grow along the streamwise direction until the reattachment position. Downstream of the mean reattachment point, the vorticity decreased gradually with increasing x. On the other hand, a non-negligible level of wall-normal vorticity (0),) was detected in the x-z plane located at y = H and near the step. Similar levels of (1), could not be found at higher and lower y locations (0.5H and 1.5H to be specific). The authors suggested that the observed a), at y = H reflects the reminiscence of the near-wall streamwise vortices associated with the boundary-layer low-speed streaks. As these vortices are swept into the separated shear layer, they get kinked and stretched exhibiting re—orientation of (1),, (streamwise vorticity) into (1),. Evidence in support of this scenario was provided from the two-point spanwise correlation function of coy in the separated shear layer at y = H and x = 2H. The correlation results indicated the existence of a quasi-periodic structure in the spanwise direction with an average wavelength equal to the typical spacing of the boundary layer streaks. In addition to PIV experiments in a BF S, Furuichi et al. (2003) investigated the spatio-temporal velocity field of the flow through an axisyrnmetric sudden expansion in a pipe using an ultrasonic velocity profiler. They plotted the instantaneous reattachment location and showed its obvious fluctuation. They also attributed this fluctuation to the flapping of the separated shear layer. A particularly interesting finding was that the amplitude of the fluctuation in the reattachment position was smaller for turbulent- boundary—layer separation than for laminar or transitional boundary layer separation. The main frequency of the flapping motion was found from the power spectrum of the streamwise fluctuating velocity around the reattachment region. The peak of the power spectrum was located at f H/Uc z 0.01 (or f x/Uw z 0.1), where Uc is the downstream convection velocity of the vortical structures (extracted from velocity measurements) and U ,0 is the freestream velocity upstream of the step. As have most other researchers, the authors obtained a downstream convection velocity of 0.5U..o in the turbulent regime. They reasoned that the flapping motion resulted from the interaction between the downstream-traveling shear-layer vortices and part of the recirculation flow moving upstream. In their wavenumber-frequency plots, up and downstream convection velocities were observed in the separation of the transitional boundary layer. However, only downstream convection velocity was detected for the separation of laminar or turbulent boundary layers. The unsteady behavior of the reattaching shear layer of the BFS flow was studied systematically by Driver et al. (1987). From the streamline plots conditioned on short, intermediate and long reattachment cases (where the reattachment length was detected using concurrent pulsed wire measurements that was operated as a thermal tuft), the authors identified the amplitude of the flapping motion to be less than 20% of the thickness of the shear layer at reattachment. The fi'equency of the flapping motion was investigated using wall-pressure and near-wall streamwise velocity measurements. In the power spectra plots of the pressure and velocity, there was a peak at a low frequency of fir/Ugo = 0.18, which was believed to be the frequency of the flapping motion. However, the energy level at this fiequency was very low, compared to that of the dominant frequency in the spectrum (fx,/U.,o = 0.6), which was the same as the characteristics fiequency of the vortical structures in free shear layers investigated earlier by Winant and Browand (1974). Thus, the low frequency flapping motion actually did not contribute too much to the total energy of the flow. As Driver et a1. (1987) did, many researchers focused on the investigation of the near-wall flow structure. Two common methods for studying wall/near-wall effects of the flow structure are wall-pressure and wall-shear measurements. Knowledge of the surface pressure field is necessary to understand and control vibration and noise generated due to pressure fluctuations. Measurements of the wall shear, on the other hand, directly define the reattachment point, where the mean wall-shear stress is zero. There is a wealth of information available on one- and two-point measurements of the wall pressure in separating/reattaching flows. More recently, Hudy et al. (2003) conducted more comprehensive measurements using wall-pressure-array measurements in a splitter plate with a fence configuration. Lee and Sung (2001, 2002) did multiple-arrayed fluctuating pressure measurements in a BFS flow model. Compared with the one- and two-point measurements, the array measurements enabled attainment of more details of the spatio as well as temporal information of the flow structure. Therefore, only these references of wall-pressure measurements are reviewed here. Lee and Sung (2001, I) used an array of 32 microphones to measure the pressure fluctuations in the separated and reattaching flow over a BFS at Reynolds number (based on step height) of 33000. A variety of signal processing techniques were employed in the analysis of the array data: time-domain and spectral methods in Lee and Sung (2001, I), unsteady wavelet analysis in Lee and Sung (2001, II) and spatial box filtering and conditional averaging in Lee and Sung (2002). Consistent with most one- and two-point measurements, Lee and Sung (2001, I) identified two peaks in the pressure power spectra at: fH/U... = 0.015 (or fir,./U..o = 0.11) and fH/U... = 0.068 (or fx/U.o = 0.5). These frequencies corresponded to the flapping motion and passage of the large-scale vortical structures, respectively. From their coherence and wavenumber spectra plots, they suggested that the pressure fluctuations were primarily associated with the large-scale vortical structures and boundary-layer—like decaying modes at high frequencies. Their later analysis in Lee and Sung (2001, H) showed that the pressure fluctuations were predominantly generated by the vortices, which were associated with two types of signatures: global oscillation, where all pressure fluctuations were in phase across the entire separation bubble, and propagating disturbance, with a convection velocity of 0.6Uw. The two modes are synchronized with the flapping motion that induced expansion and contraction of the separation bubble. The authors provided further support for these observations using conditional measurements of the velocity field that are based on wall- pressure events in Lee and Sung (2002). Hudy et al. (2003) investigated the space-time character of the fluctuating wall pressure on a splitter-plate that was attached to and placed in the wake of a fence. The Reynolds number based on the fence height above the splitter plate was 7900. The majority of the data analysis was based on measurements from 28 microphones on the centerline of the plate. The autocorrelation of the pressure fluctuations implied the 10 existence of two time scales. A large time scale dominated the pressure signature from the fence to 0.25x,. Downstream of 0.25x,, a shorter time scale prevailed. The former was attributed to the flapping motion and the latter corresponded to the Shear-layer structures. These observations were consistent with the power spectra results. More specifically, the peak of the power spectra obtained close to the fence was found at f(2Hj/Uw= 0.02-0.03 (i.e. fits/U“,= 0.12-0.18); Farther downstream, the spectral peak shifted to a frequency of f(2Hfl/U¢,= 0.1-0.15 (i.e. fir/U“,= 0.6-0.9). The convective nature of the wall-pressure signature was clearly illustrated in their cross-correlation and phase-angle results. For x > 0.25x,, an average convection velocity of 0.57 U... in the downstream direction was found. More interestingly, Hudy et al. (2003) also found that an upstream convection velocity of 0.21 U... was associated with low-frequency pressure fluctuations in the region extending from the fence to 0.25x,. This, combined with other results, led them to propose that the flapping motion was produced by the existence of an absolute instability of the recirculation bubble. Compared to wall-pressure measurements, there is little information available on the turbulent wall-shear-stress field in the literature. Since the near-wall flow behind a backward-facing step is highly turbulent and flow reversal occurs frequently, it is difficult to conduct measurements of the wall-shear stress in this region. Although modern techniques such as PIV and LDV are capable of capturing the unsteady direction- reversing flow information downstream of the BFS, their capability to conduct near-wall measurements is limited or they may require provisions that are too elaborate and/or costly to resolve the near-wall flow. F emholz et al. (1996) did a comprehensive review of techniques for measurement of the wall-shear stress in wall-bounded and separated flows. 11 They concluded that only wall hot wires and pulsed wires were capable of measuring the unsteady wall shear. But wall hot wires can’t be used in cases involving flow reversal. Wall pulsed wires, on the other hand, measure the wall-shear magnitude and direction, but their frequency response is limited to tens of hertz. As a result of these difficulties, only recently, information other than the rms (root mean square) and instantaneous direction of the fluctuating wall-shear in backward-facing—stcp flows have become available in Spazzini et al. (2001) and Tihon et al. (2001). A review of the pertinent literature on mean and fluctuating wall-shear measurements in BPS flows follows. The wall shear is generally presented in the following normalized form in the literature: t C = W (1.1) f 1/2pU020 where Cf is the skin-friction coefficient, rw is the wall-shear stress, p is the fluid density. Westphal et al. (1981) measured mean and ms skin-fiiction coefficients in a BFS flow using a wall pulsed-wire. Devenport and Sutton (1991) investigated an axi- symmetric sudden expansion also using a wall pulsed-wire. Later, Jovic and Driver (1995) studied the effect of Reynolds number on the mean skin fiiction coefficient in a BF S flow using a Laser-oil Flow Interferometry (LOI). Most recently, Tihon et al. (2001) investigated wall-shear vectors behind a BFS using a three-segment electrodiffusion probe in a water channel, while Spazzini et al. (2001) measured the wall shear to investigate the flapping motion in a BF S flow using a special wall-mounted double hot- wire probe. In all of the above studies, the distribution of the mean skin-friction coefficient 12 'n I 5.11 “ was similar. The mean Cf has a small positive value from the step to a certain xo/xn the value of which was 0.34 in Tihon et al. (2001), 0.2503 in Spazzini et al. (2001), 0.23 in Jovic and Drivier (1995), 0.10 in Devenport and Sutton (1991), and 0.13 in Westphal et al. (1981). Downstream of this position, the mean Cf crosses zero and becomes negative beneath the recirculation zone from xo/x, to x/x, = 1. The peak value of negative skin fiiction was located at x/x, of 0.69 in Tihon et al. (2001), 0.60 in Spazzini et al. (2001), Jovic and Drivier (1995), and Westphal et a1. (1981) and 0.55 in Devenport and Sutton (1991). Downstream of reattachment, the mean Cf increases fi'om zero to a certain positive value at x/x, around 1.5 in most studies. Finally, farther downstream of x/x, =1.5, the mean Cf decays very slowly. In the mean skin-fiiction distribution, there are two positions where the mean Cf value is equal to zero. The first one indicates the time-mean location of the separation point at the boundary between the primary and secondary recimlation regions. The second one corresponds to the mean reattachment location. The magnitude of the mean Cf was found to depend on the Reynolds number. This dependence was investigated by Jovic and Driver (1995). They found that the skin- fiiction coefficient magnitude decreased with increasing Reynolds number. The minimum skin-friction coefficient (Cflm), i.e., that corresponding to the largest negative shear stress, was found to have a power law relationship with Reynold number with an exponent of —1/2. This suggested that the flow near the wall within the recirculation region Was viscous-dominated or laminar-like. Tihon et al. (2001) also found a similar relationship and supported the same conclusion. The rms skin-friction coefficient results are also similar in the literature. CM,” 13 exhibits very low values beneath the secondary recirculation region, but it increases sharply to a maximum value at x/x, of about 0.78 in Tihon et al. (2001), 0.70 in Spazzini et al. (2001), 0.70 - 0.90 in Devenport and Sutton (1991), 0.60 in Westphal et a1. (1981), and 1.0 in Adams and Johnston (1988). Downstream of this position, the rms skin-friction begins to decay slowly, maintaining a steady value after reattachment. Overall, it may be stated that the peak rms skin-fiiction coefficient is located between the location where the mean skin-friction coefficient reaches its negative peak and the mean reattachment point. The percentage downstream/upstream flow, also referred to as forward/reverse flow probability, is another quantity used to determine the reattachment position in a BFS flow. The method was used by Eaton (1980), who employed a thermal tufi to detect the wall-shear direction, Westphal et al. (1981), Tihon et al. (2001) and Spazzini et al. (2001). The location of zero mean skin-fiiction coefficient was found to be consistent with that of a forward or reverse flow probability of 50%, as demonstrated in the literature. Furthermore, the dependence of the probability on the location within the separation bubble is consistent with that of the mean skin-fiiction coefficient. In particular, most of the time, the flow is towards the downstream direction beneath the secondary recirculation region and in the upstream direction below the primary one. Downstream of reattachment, in the redeveloping boundary-layer region, the downstream flow probability is larger than 50%. Tihon et al. (2001) and Spazzini et al. (2001) obtained time-resolved information of skin-fiiction measurements extending fi'om the step to beyond x,. From the power spectra of fluctuating wall-shear stress, Tihon et al. (2001) found that the frequency (fir/Ugo) of the peak value in the spectra varied fiom 0.15 to 0.65 between x/x, =0.3 and 14 0.8. These frequency values were consistent with those obtained from wall-pressure studies of Lee and Sung (2001, I) and Hudy et al. (2003). Spazzini et al. (2001) focused their investigation on the low-frequency motion: i.e., the flapping motion. As found in Hudy et al. (2003), the wall-shear fluctuation was also dominated by low-frequency unsteadiness from the step to x/x, z 0.25-0.30, where the secondary recirculation zone ended. Downstream of x/x, 5:: 0.25, the spectra were dominated by a higher frequency component. From wavelet analysis of the measurements and flow visualization, Spazzini et al. (2001) proposed that the cause of the flapping motion was a cyclic motion of growing and breakdown of the secondary recirculation bubble. 1.3 Motivation Although there is a wealth of information on the flow structure above the wall, and associated wall-pressure in BFS flows, only a few studies of the turbulent wall-shear stress are available. The latter are also limited to one-point measurements as seen from the above review. Insight into the spatio-temporal character of the wall-shear-stress field can’t be examined fiom these measurements and questions pertaining to this character remain open. For example, is the spatial and temporal behavior of the wall-shear-stress field consistent with the time and spatial scales of the separated Shear layer identified in the above literature review? Do the low-frequency wall-shear-stress fluctuations contain both up and downstream propagating disturbances as found in wall-pressure measurements of Hudy et al. (2003)? Addressing such issues is important to answer some of the open questions concerning this classical, yet important, flow. The database to be accumulated here to answer these questions will also be 15 beneficial in supporting modern development in computational fluid dynamics (CF D). CFD has developed rapidly over the past decade. Direct Numerical Simulation (DNS) and Large Eddy Simulation (LES) are able to provide the detailed flow structure information. However, computational results need to be verified and validated with experimental data. Spatio-temporal wall-shear measurements to verify wall influences of computed turbulent structures are still rare, if existent at all. Therefore, the current study contributes information that is useful for benchmarking of DNS and LES codes. The axi- symmetric geometry of the back-step model used here was selected in part due to suitability for comparison with computations based on periodic boundary conditions. It is noted here also that there are very few experimental studies of BFS flow in the literature that are based on such an axi-symmetric geometry. 1.4 Objectives The corresponding objectives for the current work are summarized as follows: 1. To develop a new hot-wire-based sensor for near-wall measurements in unsteady direction-reversing flows with frequency content up to hundreds of hertz or a few kilohertz. 2. To develop a low-speed calibration facility for calibrating near-wall hot- wire-based sensors. 3. To measure the wall-shear stress in an axisymmetric backward-facing-step flow at different Reynolds numbers. Information such as the mean/rms of the wall-shear stress, flow-direction probability, power spectra, etc, will be collected and compared with the literature. 16 To conduct two-point measurements in the backward-facing-step flow to obtain spatio-temporal information of the wall-shear stress such as the two-point space-time cross-correlation function. The spatio-temporal information from the wall—shear will be compared with those obtained from wall-pressure and velocity-field measurements in the literature. 17 CHAPTER 2 EXPERIMENTAL SETUP This chapter discusses the experimental setup and measurement techniques. The setup consists of the wind ttmnel, axi-symmetric backward-facing—step (BF S) model and wall-shear-stress calibration facility. Each of these components is presented in details in sections 2.1 through 2.4. 2.] Wind Tunnel As shown in Figure 2.1, the wind tunnel is composed of five sections: contraction, test section, pre-diffuser, diffirser and axivane fan driven by a GE 15 HP DC wound shunt motor and controlled using a GE gp-100 adjustable speed controller. Each section is supported on wheels and could be easily moved around. The flow is drawn into the contraction section and discharges to the atmosphere at the exit of the fan section. The total length of the tunnel is approximately 9 m and its centerline height is 1.346 m. The duct of the contraction section is made of a composite lamination of fiberglass-reinforced plastic and a rigid wood core material molded over precision tooling. The contraction has a 6.25:1 area ratio and an inlet section that is 1.549 x 1.549 m in height and width. There is a precision aluminum honeycomb and three graduated, high porosity screens mounted in front of the contraction to reduce the turbulence intensity of the intake air. The freestream turbulence intensity in the test section was found to be less than 0.45% from measurements in the velocity range of 5 -— 15 m/s. l8 Fan Diffuser Pre-Diffuser Adj-11.591316 Test Contraction \ Ceiling Section ”‘7‘ ’/—*-¢‘/= I‘_ 1 486—>~——1 829———><—1005—fiH——1 829—><———2096——> Figure 2.1 Schematic diagram of the wind tunnel (dimensions in mm) The test section is 1.829 m long, and its upstream end is bolted together with the contraction section with a sponge filling material in-between for sealing purposes. An adjustable ceiling that is hinged at the flow entrance is used to establish zero pressure gradient in the freestream, leaving a cross sectional flow area of 609.6 x 609.6 mm at the upstream end. Two 12.7 mm wide slots in the middle of the ceiling are used for insertion of hot-wire and pitot-tube probes into the tunnel. The slots could also be used to install thin plates fitted with static pressure taps to monitor the pressure distribution along the freestream flow inside the tunnel. The pre-diffuser is 990.6 mm long. The bottom wall has an angle of 6° with respect to the horizontal direction. The upstream end of this section is attached to the test section, while the downstream end leads to the diffuser. An air gap, covered with rubber filling material, is left between the latter and the pre-diffuser to minimize vibration transmission from the fan/motor. The diffuser, which is used to recover the pressure of the exiting flow, is 1.651 m long and has an angle of 588° and area ratio of 1.9. At the 19 downstream end of the diffuser, a fan-diffuser adapter converts the square cross section of the diffuser into a circular area to match the inlet geometry of the fan. 2.2 Axi-symmetric Backward-facing-step Model Figure 2.2 shows a picture of the axi-symmetric backward-facing-step model supported on a traverse table. The corresponding schematic diagram of the model is plotted in Figure 2.3 to show the dimensions of all components. The total length of the model is 2.383 m. At its upstream end, the model is fitted with an aluminum half- spherical nose made with a radius of 63.5 mm. The 2D section is made from 127.0 mm diameter aluminum tubing with a wall thickness of 6.35 mm. The step section is also made from the same aluminum tubing with 101.6 mm outside diameter and 6.35 mm in wall thickness. The difference between the tubing diameters creates a nominal step height (H) of 12.7 mm between the rotating and step sections. Downstream of the latter, the motor section is a thick steel tubing with 101.6 mm outside diameter and 69.9 mm inner diameter. The bottom surface of the tubing is flattened and seated onto the top of a support bar and the two parts are attached firmly together using eight 3/8-inch screws. Finally, the downstream end of the model is terminated with conical tail section to avoid abrupt termination of the geometry. The dimensions of the steel support plate are 19.05 mm thick, 482.6 mm long and 501.65 mm high. The bottom of the plate is connected to the top of the traverse table, providing a rigid support of the model at the downstream end. At the opposite end, four piano wires located at about 90° with respect to each other are attached to the 2D section of the model, as shown in Figure 2.4, to prevent deflection due to the long overhanging 20 portion of the model. The wires pass through narrow slots in the windows of the test section and are straightened with tension and connected to a support framework outside the tunnel. Since the diameter of the wires is only 0.99 mm, their effect on the flow is anticipated to be negligible. This will be verified later by checking the azimuthal symmetry of the flow around the model in the flow region of interest. Tail i\-‘Iotor Section Step Figure 2.2 Picture of the axi-symmetric backward-facing-step model The traverse table has two angle adjustments: yaw, along the spanwise direction, and tilt, along the vertical direction. In addition, three translation adjustments are available in the vertical, streamwise and spanwise directions. These positioning flexibilities ensure that the BFS model is installed at the center of the wind tunnel and aligned with the freestream velocity direction. 21 Ea as 222.8% as, 388 mam 2e .8 23883 2 8am mdwmm III teem ll TI :3 1! N132 I deh IVA. m.omm 1L 0.2: G 062 9 mac m 22 Piano wires Figure 2.4 Piano-wires and boundary layer trip downstream of the model nose The boundary layer is tripped using sandpaper just downstream of the nose section (see Figure 2.4) and develops to 2D (axi-symmetric) turbulent boundary layer state at the separation point. The development length of the boundary layer was selected by examining the study of Driver and Johnston (1990) of non-separating two- and three- dimensional turbulent boundary layers over an axi-symmetric model. Based on this examination, a length of about 1.2 m was found to be sufficient for the transition of the boundary layer to a turbulent state for the range of flow speeds anticipated. More specifically, the length from the tip of the half sphere nose to the edge of the Step is 1.219 In. On the other hand, the distance from the step edge to the support plate edge is 487.7 mm, or 38.4 step heights. For reference, Eaton and Johnston (1981) found the reattachment length for a planar two-dimensional backward-facing-step flow to fall in the range of 4.9 - 8.2 step heights. As will be seen later, the largest reattachment length found in the current measurements is approximately 4.9H. Hence, the support plate is almost 23 eight reattachment lengths, or 40H, downstream of the step. The measurements on the model were taken primarily in the region extending from the step to 127.0 mm (i.e. 10 step heights) farther downstream. Figure 2.5 shows a magnified view of the measurement region. Within this zone, four sensor covers are located at the top, bottom, left and right of the cylindrical model surface. There are static pressure taps in sensor covers numbered 2, 3, and 4 in Figure 2.5. The slot for sensor cover 1 could be used to install two oscillating-hot—wire sensors for two-point wall-shear measurements (as shown in Figure 2.6) or static pressure taps mirroring those in cover 3. The static pressure taps were primarily used to verify the alignment of the model parallel to the fieestream direction. For the sake of the alignment process, 24 static pressure taps are installed in sensor cover 2. The first of these is located at 4.76 m away from the edge of the step with all taps equally spaced at an inter-sensor spacing of 4.76 mm. For the remaining sensor covers, only eight static pressure taps are used in each cover, with the first tap positioned at 19.05 mm downstream from the edge of the step. The corresponding distance between two neighboring taps is 9.52 mm. All of the signal wires and static pressure tubes are routed inside the outer model shell, coming out from two holes at the upstream junction between the outer shell and support plate. The cables and tubes then run the height of the support plate, covered from the flow by a shroud in front of the plate as shown in Figure 2.2. Finally, all static-pressure tubes are connected through a 48-port scarrivalve to a Setra Model 239 pressure transducer with a range of 0-5 inch water (i.e. 1245 Pa). 24 (a) Sensor Cover 2 (b) A-A Cross-section View Sensor Cover 2 with 24 pressure taps Sensor Cover 3 with 8 pressure taps Sensor Cover 1 with OHW sensors (or 8 pressure taps) Sensor Cover 4 l with 8 pressure I taps Figure 2.5 Magnified view of the measurement zone showing wall-sensor covers: (a) 3D view, (b) Cross-section view 25 Figure 2.6 Picture of the sensor cover containing two OHW sensors The model was aligned by adjusting the tension of the piano wires and the position of the traverse table. First, coarse alignment of the model was done using a level gauge and tape-measure in order to position the model centerline at the center of the test section and make it parallel to the nominal flow direction. This was followed by fine alignment, which relied on the static pressure measurements from sensor covers 1- 4. The static wall pressure along the azimuthal direction of the model should be equal at the same streamwise (x) location if the model is perfectly aligned. Figure 2.7 shows static pressure measurements after alignment of the model for a fi‘eestream velocity of 15 m/s. The labels of east, north, west and south in the legend refer to those measurements from sensor covers 1 —— 4, respectively. The abscissa represents the streamwise location downstream of the step normalized with the reattachment length (x,), the determination of which will be discussed in details in Chapter 4. The ordinate represents the pressure coefficient Cp defined as follows: 26 Cp =-—————ps'p’2 (2.1) I/Zono where p, is the surface static pressure along the model, p, is a reference pressure measured at the entrance to the test section, ,0 is the air density, and U... is the freestream velocity. The four curves in the figure collapse at most x positions. The largest discrepancy is found at x/x, around 1 and is less than 8% of the total Cp variation. It is believed that this discrepancy is caused by small increase in the pressure on the lower (south) side of the model resulting from flow stagnation on the Shroud. 0.15 . -6- North o1-e—South “-A—East 0.05» Figure 2.7 Static pressure distribution downstream of the step at four azimuthal locations and U... = 15 m/s To further examine the alignment of the model, the azimuthal uniformity of the flow just upstream of separation was checked using a single hot-wire probe. The sensor was positioned at a certain height y above the wall and less than 0.5 mm upstream of the 27 step to measure the mean streamwise velocity U at different azimuthal angles ((1)) while maintaining the same height. To rotate the probe around the centerline of the model, a ring holder was fabricated with an inside diameter that matches the outside diameter of the model downstream of the step. The step ring was made of two halves that could be split and re-assembled around the model such that the ring's center coincides with that of the model, as shown in the photograph in Figure 2.8. As also evident from the figure, the hot-wire probe was mounted on the ring holder, which is located well downstream of the step to avoid interference with the flow at the point of measurements. By rotating the ring in the azimuthal direction it was possible to move the probe to different azimuthal angles. : Step V Traversing ring Hot-wire probe Figure 2.8 Picture of the setup for checking azimuthal symmetry of the flow Azimuthal measurements for four different heights are illustrated in Figure 2.9. The heights are normalized by the local boundary layer momentum thickness 19 28 (determination of 19 will be discussed in section 4.1). For each height, U is plotted versus the azimuthal angle (with zero angle corresponding to probe position on the top side of the model) and plotted as percent of the average velocity over all angles. At the three lower y positions, the velocity varies within i10% of the average value of the measurements. However, since these measurements were within the boundary layer, the error could also stem from probe-height variation caused by inaccuracies of the azimuthal traversing method. At y/0 = 10.59, which is almost at the edge of the boundary layer, the variation is less than 4%. 120.4 MW, ~ , ‘e',L/9,=1'65 . —91 we = 5.;9 910" § E ---------- ~ I z ; I . 80 ,,-,,L i__, .4 80' , h , . 7 1,.-. ,, 0 90 18 270 360 0 90 180 270 360 11, (degree) 4) (degree) 1207 120‘ ~ if, ., : i ,. , —A—Wy/6= 7.94 II I —1— we = 10.59 o 90 180 2Io 360 0 ’9‘0 7180 270 860 4, (degree) :1 (degree) Figure 2.9 Azimuthal symmetry of the flow at separation for U... = 15 m/s 29 After alignment of the model, the ceiling of the test section was adjusted to obtain zero pressure-gradient condition in the freestream upstream of the step. For this purpose, two 3.2 mm thick by 12.7 mm wide plates instrumented with 16 static pressure tapes were fitted into the probe-access slots along the ceiling centerline. The plates were flush with the ceiling surface. In addition, 66 micron-thick kapton tape was used to cover any possible gaps between the inserts and the ceiling. The location of the first static pressure tap (Cl) was 19 mm downstream of the tip of the model. The pressure taps C1 through C7 were 76.2 mm fi'om center to center. Because of a discontinuity in the probe-access slot, the distance between taps C7 and C8 was 299.2 mm. The remaining taps (C8 - C15) were also spaced 76.2 mm fiom center to center. The results for freestream velocities of 5, 10 and 15 m/s are given in Figure 2.10. The x locations of the pressure taps are _ normalized as follows: (x-xJ/Q where x is the distance to the tip of the model, x, is the position of the step, and 6 is the momentum thickness of the boundary layer at 0.5 mm upstream of the step. The two vertical broken lines included in the figure indicate the positions of the model tip and the step ((x-xJ/Q = 0). The error bars are used to indicate i 10% variation of the measured values. The results demonstrate that a nominally zero pressure gradient, corresponding to a constant pressure value to within 3:10%, is established for at least 3006 upstream of the step. The initial pressure drop around the nose of the model is presumably caused by flow acceleration associated with the fairly small model blockage of 3.4% (where blockage is defined as model cross sectional area upstream of the step divided by test section flow area). 30 I _ um: 5rn/s nor '0.“ ' o __ _l_ ____ oe-ooe - -093. -o.1, .V2L_ _, L_._.-______ -11..__.._n__.-. . izz.,,x,AJL_ 0'12 -500 -400 -300 -200 -100 0 o. _.- . .. .____..- . — I —_J_. -0.02 -o.04I I (ft-0.06: I -o.oa» I -0.1L -0.12 bison—ELIOTLSM -360 Ciao—”T0 I -o.o2I I: 1 MI ' 0 0° -o.061 -0.08r LI -500 .460 4430 4’66“ i1—00_—_0 x-x, 19 -0.12 Figure 2.10 Freestream static pressure distribution upstream of the step 31 2.3 Boundary Layer Profiles Boundary layer mean-velocity profiles were obtained at 0.5 mm upstream of the step using a single hot-wire probe. The hot-wire probe was installed on a traversing mechanism driven by a TMG Model 56188-01 hybrid stepper motor. The traverse system was calibrated against a dial indicator with a 12.7 um resolution. The calibration results Showed that the system traversed a distance of 3.3 pm per motor step, leading to a very fine control of the probe motion. The initial position of the hot-wire probe above the surface was measured using an imaging system. The systemconsisted of a 768 x 494 pixel Sony XC-75 CCD camera connected to a PC computer via a National Instruments IMAQ PC1-1411 frame grabber. A Perkin Elmer MVS-2060 strobe light sent a short duration, high intensity light pulse to illuminate the probe tip and step area from the opposite side of the camera (as depicted in Figure 2.11) every time when an image was acquired. The distance between the tip of the wire and top of the model was found in pixels and then converted to m using a calibration constant. The latter was found by moving the probe a known distance of 0.5 mm and finding the number of pixels between the probe's initial and final locations. The calibration constant was typically about 4-6 urn/pixel. Once the initial position of the probe was found, it was moved up to the freestream and calibrated against a pitot-static tube. The stagnation and static ports of the pitot tube were connected to the high and low inputs, respectively, of a Setra Model 239 pressure transducer. Two transducers were available: one with a pressure range of 0.5 inch of water (i.e. 124.5 Pa), and the other with a higher range of 5 inch of water, or 1245 Pa. Both transducers had an output of five volts at the full-scale pressure. Thus, the 32 former transducer, with its smaller range but higher sensitivity, was used for flow velocities less than 14 m/s, while the higher range transducer was used for velocities above 14 m/s. The flow temperature was monitored by an Omega DP25-TH thermistor indicator for use in calculation of the air density and temperature compensation of the hot-wire response. (a) (b) Hot-wire prong Figure 2.11 Typical image of the probe tip and step area used for hot-wire positioning near the wall: (a) Wire at the initial position, (b) Wire at 0.5 mm above initial posrtron After calibration, the hot-wire probe was traversed back down to the initial location near the wall. The accuracy of re-positioning of the wire was verified using a dial gauge attached to the traversing system. The probe was then traversed away from the wall to 33 different heights, or y positions. At each height the mean and rms (root mean square) velocity were recorded for ten seconds at a sampling rate of 1000 samples/s. The traversing steps followed a logarithmic distribution with smaller values in the near-wall region where the mean-shear is high, and larger ones closer to the freestream. After the experiment, a “post” hot-wire calibration was done to check any drift in the sensor's 33 response. If there was more than 5% error between pre and post calibration, the data were discarded and the experiments were repeated. The results for the boundary layer profiles and associated characteristics will be analyzed in section 4.1. 2.4 Couette Flow Calibration Facility In the current study, a new hot-wire technique was used to measure the wall-shear stress. Therefore, a facility for calibration of near-wall hot-wire sensors was required to be deve10ped. Some of the common methods to calibrate wall-shear sensors include the use of a firlly-developed laminar or turbulent channel flow or comparison against a direct measurement methods, such as oil-film interferometry. Here, a Couette-flow facility was used for the calibration. This choice was motivated by the fact that the facility met certain Special considerations that are required for successful measurements of the velocity very close to the wall (within 100 pm) in a separating/reattaching flow environment. First, the calibration range must extend down to velocity, or shear-stress, values that are very close to zero. A Simple extrapolation of King’s law-type calibration to lower velocity values is not valid since the heat transfer mode from the wire switches fiom forced, where King’s law is applicable, to natural convection as the velocity approaches zero (in addition to the increasing importance of heat conduction to the prongs near zero velocity). Second, the calibration facility should have provisions to reverse the flow direction in order to verify and document the Wire’s directional sensitivity. Third, since ultimately the measurement of interest is 1..., one needs to know the wall-shear—stress value at different calibration velocities. The idea for the Couette flow calibration facility was first discussed by Khoo et 34 al. (1998). The facility consists of two discs: one of the discs is stationary and the other is spinning. Ifthe gap between the discs is small enough, the azimuthal flow velocity varies linearly across the gap; that is, establishing a Couette flow field. The flow direction could also be reversed easily by switching the rotational direction of the spinning disc. A similar device was designed and constructed in this study as demonstrated in Figure 2.12. The picture gives an overview of the device. Figure 2.13 shows a cross section of the facility along with relevant dimensions. The rotating disc is made of aluminum and the stationary disc is fabricated from plexiglass. Both discs have a diameter of 317.5 mm. The rotating disc is connected to a steel drive shaft seated in a bearing assembly. Two NTN high-precision angular contact bearings (model number: 7006CDB/GNP4) at the right end of the bearing assembly are fixed in position in back- to-back arrangement by trapping the inner and outer races of the bearings using shoulders or nuts. A deep groove ball bearing (model number: 6204—2RSl) is floated at the other end of the bearing assembly by fixing its inner, but not outer, race. The purpose of the floating arrangement is to allow the shaft to deform due to thermal expansion. A timing pulley (model number: Browning 40XLBO37) at the end of the shaft was connected through a timing belt to another pulley (model number: Browning 20XLBOB7) on the shaft of a 12/24-volt DC motor (model number: Dayton 4Z144). The pulley on the main shaft is twice as large as the pulley on the motor. In this fashion, it was possible to run the rotating disc at a speed as low as 18 rpm to calibrate the sensors at very low speeds. 35 Rotating Disc Bearing ISDtlitéonary Assembly I \SDCHI ('1 ‘Sensor Figure 2.12 Picture of the Couette Flow Facility Rotating Stationary Disc Spring-loaded support screw Bearing Assembly Timing Belt \ f7 T g ~ Supporting 190.5 bracket L I _ T _- . ~ 476.3"* ' * Figure 2.13 Cross Section of the Couette Flow Facility (dimensions in nun) 36 The Couette facility motor is driven by a Sorensen DCR 40-13B DC power supply. Experimentation with the motor-driving unit showed that when driven by the DC power supply, the motor has lower running noise and less vibration than when operated using a SCR DC motor controller. However, the disadvantage of a DC power supply is that there is no over-load protection of the motor. This was remedied by monitoring the Sorensen power supply's current meter to ensure that the current drawn by the motor does not exceed the rated value. At all speeds, the current was less than 80% of the full load current of the motor. Finally, the rotational speed of the motor was measured using an Omron EE-SBSV photo-detector switch. The sensor output was switched off whenever a piece of black tape adhered to the rotating disc passed in front of the sensor. This resulted in output of one 5V pulse per revolution from the switch. During calibration, the pulses were sampled, counted and divided by the sampling time to yield the average rotational speed of the disc. Proper-rotation and vibration of the rotating disk was checked by a Kaman KD- 2300 hall-effect displacement sensor, the sensitivity of which was 0.5 pm per milli-volt, that measured the distance from the surface of the disk to the tip of the sensor within the measurement range (005 mm). The sensor was installed in the facility's supporting bracket on a three—dimensional traversing mechanism and adjusted to be perpendicular to the rotating disc with the sensor tip about 0.25-mm away. The maximum variation measured in the distance between the disc surface and the sensor tip was 27 pm over the range of rotational speeds employed here. This variation represented less than 5% of the 600 um gap between the two plates. The latter was set using feeler gauges while ensuring parallelism of the two plates by adjusting three spring-loaded support screws 37 (see Figure 2.13). The resulting uniformity of the gap was verified during rotation using high-magnification video imaging of the gap. A sample frame from the video recording is given in Figure 2.14. Examination of different frames of the video demonstrated that the gap variation was less than 37 um. Rotating / disc Stationary disc Gap Figure 2.14 Image of the gap between the rotating and stationary discs Khoo et al. (1998) calibrated near-wall hot-wires successfully in a similar type of facility as that used here and verified their operation in fully-developed laminar and turbulent channel flow. The same authors also did a Direct Numerical Simulation (DNS) to solve for the flow field between the two discs and compared the DNS results to the Couette flow assumption. These results are displayed in Figure 2.15, where three plots showing the radial (V, , azimuthal (Va) and wall-normal (Vy) velocity profiles across the gap are given. The symbols in the plots depict DNS data while the solid lines represent asymptotic solution obtained by Stewartson (1953) for the flow between two infinitely wide discs and small Reynolds number based on the gap size: 38 52-112 V Re}, = (2.2) where Q is the rotational angular velocity, It is the gap between the two discs and v is the kinematic viscosity. Also included in the figure is a dotted line indicating the typical. height above the wall for the sensors used here during calibration. A few important conclusions may be drawn from Figure 2.15. First, the agreement between the DNS and analytic solution for the radial and azimuthal velocities is good for all Reynolds numbers providing confidence in both solutions. For the wall- normal velocity, the agreement between the two approaches is limited to Re h of 3 or less. However, the latter velocity component is more than three orders of magnitude smaller than the other two components, and therefore its influence on the hot-wire calibration can be ignored. A second important conclusion is the fact that V9 is in fact found to vary linearly across the gap (middle plot) as expected for Re In of 4.5 or smaller. This provides confidence in that the Couette-flow assumption made here is valid for the entire calibration range since the largest Re h value encountered during calibration is equal to 4.7. For the purpose of the calibration, the hot-wire sensor was installed in a 63.5 mm- diarneter wall plug whose top surface was flush with the inside surface of the stationary disc. The center of the plug, where the sensor's center is also located, was at a radial position (r...) of 120 m. If the wire height above the wall is denoted by y, then the velocity ( V9) and wall shear (I...) seen by the wire are given by: Q-rw V0 = y- (2.3) 39 dV .Q-r Twzfl dyg=y hw (2.4) where u is the dynamic viscosity. The facility was typically operated from 0 - 1800 rpm, leading to wall-shear—stress variation within the range of 0 to 0.66 Pa. Prior to calibration, the wire height was determined using a high-magnification video system. The wall plug was installed on an ULTRAlignTM traversing mechanism (model number: Newport M-462) with a resolution of 2 um. Using focused strobe light illumination, the image of the wire and surface of the wall plug was viewed on a video monitor. The wire position in the image was first marked on the monitor. The wire height above the wall was then determined fi'om the micrometer reading of the traverse of the translation distance necessary to traverse the plug surface within the view of the microscope to the marked location of the wire. The calibration velocity/ wall-shear values were selected to cover their actual range in the BFS flow. Specifically, the minimum wall shear is zero in this case, while the maximum mean-skin friction-coefficient is never above 0.003 in the literature; e.g., Jovic and Driver (1995) and Spazzini et al. (2001), which corresponds to a wall-Shear value of 0.42 Pa at 15 m/s, the highest velocity of interest. Two separate curve fits were used: polynomial fit for the low velocity /wall-shear values (about 0 —l .2 m/s or 0 — 0.21 Pa) where natural convection effects may be important, and King's law fit for high velocity /wall-shear values (about 1.2 — 3.8 m/s or 0.21 -— 0.66 Pa). There were about ten points in each calibration range to ensure reliability of the curve fit. Also, each calibration data point was obtained from averaging the wall-shear magnitude and hot-wire voltage over at least 50 rotations of the rotating disc in order to "average out" the effect of any background noise such as that caused by motor vibration. 4O Vr/(w-r) 0.06 f . Symbols: numerical results: 9 Reh = 0.5 0.04 1' '. Line: series solutions ‘ I Re}, : 1.0 A RE}, I 1.5 0'02 . F . x Reh : 3.0 0 e ‘ etc Reh :45 , o R) : 6.0 -0.02 1 ‘h -0.04 4 -0.06 0.0 0.2 0.4 0.6 0.8 1.0 y/h V9 my - r) 0.0 0.2 0.4 0.6 0 y/h V}, /(a) - r) 6.0E-5 4.0E-5 2.0E-5 0.0E+0 -2.0E-5 -4.0E-5 -6.0E-5 -8.0E-5 0.0 0.2 0.4 0.6 0.8 1.0 y/h Figure 2.15 Comparison between DNS results and Couette flow assumption (Khoo et al. (1998) 41 CHAPTER 3 OSCILLATING-HOT-WIRE (OHW) TECHNIQUE The main measurements in this thesis were conducted using a new oscillating-hot- wire (OHW) sensor developed in the Flow Physics and Control Laboratory at Michigan State University for the purpose of this work. The sensor is able to capture the near-wall velocity/wall shear signature of unsteady direction-reversing turbulent flows. The concept and validation of the operation of a prototypical OHW sensor suitable for use in flows containing unsteady frequencies of the order of tens of hertz have been demonstrated in Li and Naguib (2003). However, near-wall flows behind a BFS involve significant flow unsteadiness occurring at frequencies of the order of hundreds of Hz to a few kHz. Thus, oscillation frequencies of the order of a few kilohertz are desired for successful operation of OHW sensors. The development of such an OHW sensor is described in this chapter. Section 3.1 reviews existing techniques for measuring the direction-reversing wall shear in air flows. Section 3.2 introduces the basic principles of the OHW sensor, while the operation and characteristics of the sensor are presented in Sections 3.3 through 3.5. 3.1 Review of Techniques for Measurement of Direction-reversing Wall-Shear Stress The near-wall flow behind a backward-facing step is highly turbulent and flow reversal occurs frequently. Therefore, it is difficult to conduct measurements, especially, of the wall-shear stress, in this region. The conventional wall-hot-wire sensors fiequently used for wall-shear measurements would not work here for their inability to discern the 42 shear direction. A brief review of the techniques that are able to measure direction- reversing wall Shear follows. Hanratty and Campbell (1983) reviewed six principal methods to measure the local wall shear and found only the direct method (i.e. floating element) and sublayer fence were able to determine the direction as well as magnitude of the wall-shear stress. Femholz et al. (1996) updated the review of the floating element and sublayer fence and also added two more methods for reverse wall-shear measurements: oil-film interferometry and wall pulsed wires. The floating-element is suspended by beams, with some gap between the element and surrounding walls. The wall shear acting on the surface of the element causes beam deflection that could be measured to obtain the force (i.e. wall shear) acting on it. But the requirement of a relatively large surface area for the small shear-stress values typically encountered in low-speed flows limits its usage. More recent studies attempted to resolve this issue through the development of microfabricated floating elements with extremely small area ( for example, Padmanabhan (1997) and Schmidt et al. (1988) ), but sensors of this type with acceptable signal-to-noise ratio, sensitivity and reliability are yet to be demonstrated. The sublayer fence is mounted on the wall, with the height of the fence less than five wall units to ensure that the fence is immersed within the viscous sublayer. The measured pressure difference between the front and back of the fence is used to obtain the wall-shear stress after appropriate calibration. However, the fence technique is still limited to primarily measuring the mean skin friction coefficient. It is unclear if the technique could be adapted for time-resolved measurements. 43 The oil-fihn interferometry is another technique to directly measure the direction- reversing wall shear. An extremely thin oil film is spread on the solid wall and illuminated using a laser beam, or some other monochromatic light source. Interference of the reflected light fiom the top and bottom of the film is used to measure the film thickness, which is related to the wall shear acting on the film through well known analytic equations. Although this is probably the most accurate method available today for mean skin fiiction sensing, it can't be used for unsteady measurements because the fihn doesn’t respond quickly enough to instantaneous wall-shear-stress variation. Pulsed-wire anemometry was first developed by Bradbury and Castro (1971) for velocity measurements in separated flows. A Similar approach for measuring wall-shear stress was later developed by Westphal et al. (1981). In pulsed anemometry, the probe consists of a central heating wire surrounded by upstream and downstream cold wires. The central wire, which is typically oriented at 90 degrees with respect to the cold wires, is heated periodically. The velocity is measured from the time of flight of the hot tracer fluid from the central wire to the downstream or upstream wires, for forward or reverse velocities, respectively. There are several difficulties and limitations of pulsed anemometry. First, to avoid thermal diffusion effects, the sensing volume size is typically no better than one to two millimeters on the side, limiting the spatial resolution of the sensor. This sensor separation also limits the frequency response to tens, or a few hundred hertz. Second, in flows with large velocity gradients, such as near walls, the measurements need to be corrected as discussed recently by Schober et al. (1998). Finally, pulsed wires require elaborate and careful calibration that renders them inconvenient for applications involving array measurements. A different variation on pulsed anemometry was introduced by Downing (1972) who used three parallel wires to measure the velocity in a one-dimensional pulsating flow in a pipe. In this approach, the central wire was operated as a conventional constant- temperature sensor and used to measure the magnitude of the velocity. The flow direction was deduced by incorporating the two outside wires, which were located upstream and downstream of the central wire, in opposite legs of a Wheatstone bridge to form a thermal tuft similar to that used by Eaton et al. (1979). Although this method overcomes some of the disadvantages of the time of flight technique, the fi'equency bandwidth remains limited because of the separation of the thermal tuft sensors and their thermal inertia. Most recently, Spazzini et al. (1999) constructed a direction-sensitive wall-shear sensor using two S-um diameter tungsten hot wires that were mounted on top of a small l-mm diameter cavity while being separated by a small gap in the direction parallel to the main flow direction. The sensor was able to detect the flow direction since the output of whichever wire was in the downstream direction was affected by being exposed to the thermal wake of the other wire. Hence, by measuring the signal from both wires simultaneously it was possible to measure the shear-stress magnitude and direction, after appropriate calibration. Although the sensor of Spazzini et al. (1999) has better spatial and temporal resolution than that achieved by earlier methods, it is fundamentally based on thermal wake effects like pulsed anemometry and variations on it. Therefore, the temporal response of the sensor is limited by the separation of the two hot wires, rather than by the much faster bridge circuit response. 45 3.2 OHW Principle From the above review, it is evident that a wall-shear-stress sensor with fast frequency response for use in direction-reversing flows is still not available today. Li and Naguib (2003) demonstrated a prototype of an oscillating hot-wire sensor for measurements in unsteady direction-reversing flows. As shown in Figure 3.1, the hot- wire prongs are attached to an oscillating base driven by a sinusoidal voltage. Ideally, the oscillating frequency is much higher than the unsteady flow frequency. Measurements of the velocity are then acquired at the oscillation center, where the oscillation velocity is the largest. In this manner, two measurements are acquired during one cycle of oscillation: one when the wire moves to the right, referred to as the backward measurement, and the other one, or forward measurement, when the wire moves towards the left. Direction-Reversing Flow ”f _ Hot Wire Prong 5 Oscillating support @ UoSin (Zizfot) Wall Side View End View Figure 3.1 Schematic diagram illustrating the OHW sensor concept In Figure 3.2(a), the flow direction is towards the right resulting in an effective cooling velocity equal to the Stunmation of the oscillation velocity (U0) and flow velocity (Hf) for the forward measurement since the wire moves against the flow. On the contrary, for the backward measurements, since the wire moves with the flow, the effective cooling 46 velocity of the sensor is the difference between the two velocities. In this case, the forward is greater than the backward measurement. However, if the direction of flow velocity reverses as shown in Figure 3.2(b), the opposite scenario takes place, where the backward is greater than the forward measurement. Thus, the flow direction could be determined by comparing the two measurements. On the other hand, the magnitude of the flow velocity/wall shear could be obtained from the measurement with the wire moving against the flow. Since only one velocity magnitude and direction is obtained during one cycle of oscillation, the oscillation frequency of the OHW sensor is also the sampling frequency of the measurements. The prototype sensor used by Li and Naguib (2003) was oscillated at a frequency of 490 Hz and U, of 0.5 m/s using a 63.5 mm-long by 31.8 mm-wide by 0.58 mm-thick piezoelectric cantilever oscillator. Hence, it is suitable only for use in flows containing unsteady frequencies of the order of tens of Hz. A great effort has been taken to build a higher frequency and smaller size OHW sensors for measurement at higher frequencies. In order to reach this goal, characteristics of the piezoelectric oscillator has to be understood first. Figure 3.3 shows a two-layer piezo-element clamped at the left end. The length and width of the element are L and W, respectively. The deflection (Ad) caused by the driving voltage (V) is proportional to L2 and V, and the resonant frequency of the element is proportional to l/LZ. Thus, in order to increase the resonant frequency, the element length has to be shortened. But that would sacrifice the oscillation amplitude. Theoretically, the amplitude of oscillation velocity should not change since it is proportional to the product of the oscillation frequency and amplitude. However, in practice, a small drop in U, was observed since the added mass at the tip of the element, for example, fi'om the prongs and epoxy, had larger effect on 47 smaller piezo—elements. Although the oscillation amplitude could be increased by a higher driving voltage, the voltage is generally limited by the characteristics of the element to below 120 volts. (a) Flow direction is towards the right Flow Flow ‘2’ “I :5 “1 %, we, Vfwd=luf|'Uo wad=qu|+Uo (b) Flow direction is towards the left Flow Flow :1 14f (1: “f %, w. Vfwd=lqu+Uo wad =|uf|-U0 Figure 3.2 Schematic diagram illustrating OHW sensor response for different flow directions: (a) Flow direction is towards the right, (b) Flow direction is towards the left Figure 3.3 Schematic of a two-layer bending piezo-element 48 Figure 3.4 shows a photograph of the OHW sensor developed for use with the axisymmetric BFS model. The main dimensions are given in Figure 3.5. The sensor is mounted in a 7.62 mm-wide wall plug with a curved top surface with a radius of 50.8 mm to match the curvature of the outer shell behind the step of the test model. The prongs of the hot-wire are attached using epoxy on one end of a 12.7 mm-long by 6.35 mm-wide by 0.58 mm-thick piezoelectric beam, protruding through 0.51 mm-diameter holes in a plexi-galss insert in the wall plug. The surfaces of the piezo-element are coated with a thin layer of nickel so that it is electrically conductive for supplying the driving voltage to the piezo-element. To isolate the hot-wire signal from the driving signal, a thin layer of 3M electric tape is placed between the prongs and piezo-element as shown in Figure 3.6. At the opposite end of the prongs, the piezoelectric beam is rigidly fixed using a clamp made out of phenolic to electrically isolate the piezo element’s electrodes from the plug. In turn, the clamp is supported on two dowel pins that are force fitted to the wall-plug body. The pins provide rigid support to the piezo element in the direction of oscillation while allowing the element’s position to be adjusted in the wall—normal direction to place the hot wire at heights (y) up to 5 mm above the wall. Once at the desired location, two set screws are used to fix the sensor’s position. Dowel pin Sensor holder Set screw Piezo-element Sensor clamp Figure 3.4 Picture of OHW sensor plug 49 Piezo-element Hot-wire Sensor holder Sensor clamp ' F, < r 7.62 < 33.0 F e - . 59-24 - Side View End View Figure 3.5 Cross-section view of OHW sensor plug (dimensions in mm) Details of the wiring connections of the OHW sensor are illustrated in Figure 3.6. One end of stripped gauge 32 wires is soldered to the bottom end of the prongs. The other end of the wires is soldered onto a printed circuit board that is adhered using epoxy to the bottom of the piezo—element, with a thin layer of self—adhesive copper film sandwiched between the board and element. The adhesive side of the copper film is in contact with the PC board, while the other side of the film, which is electrically conductive, touches the piezo-element's electode to provide the connection with the piezo-driving signal. On the other hand, connection to the hot-wire anemometer circuit is made using two 76 mm long 28 AWG hook-up wires, which are also soldered onto the PC board to make electrical connection. The use of the PC board to transfer the electrical connections to the prongs prevents the outside cabling from interfering or influencing the motion of the free- end of the pizo element. For the same reason, the piezo-driving cables are soldered onto the copper films instead of the piezo-element. This prevents any distortion of the oscillation of the piezo-element because of "cable tugging", which would have adverse 50 effects on the performance of the sensor. More specifically, the element oscillates at a high frequency but with very small oscillation amplitude. For example, at fi'equency of 4800 hertz, the amplitude is only of the order of the wire diameter, making it susceptible to any minor interference. In earlier-version OHW sensors that did not incorporate similar cable-separation provisions, wiggling the hot-wire or piezo-driving cables distorted the sensor's output signal. Hot-wire / Prongs Piezo- element Hot-wire Cable Sensor Clamp Sensor/ Clamp Copper Piezo-driving Film cable Figure 3.6 Sketch showing wire connections to the OHW sensor An Agilent 33120A function generator is used to supply the piezo-driving signal. The output of the function generator is amplified with a P1000 Hafler audio amplifier before feeding the power-amplified signal to the secondary side of a 1:186 step-down transformer to increase the voltage amplitude of the signal. In this fashion, the voltage amplitude of the sinusoidal driving Signal could be adjusted up to 180 V peak to peak. 51 The choice of the driving circuitry is determined from the power required to drive the piezo- element, which is estimated as follows: the current 1' through the piezo-element at time t is d V i(t) =C7t- (3.1) where V is the voltage across the piezo-element and C is the capacitance of the piezo- element. Since V is a sinusoidal wave with an amplitude A, then V(t) = Av sin(27rfl + (D0) (3.2) where f is the frequency of the sine wave and (Do is the initial phase angle at t = 0. The power P is given by P=imxvm=7r-f-Av2-C (3.3) For a capacitance C of 6 nf, a typical value for the piezo-element, even for as high of a voltage as 120 V and frequency of 5000 Hz, the power consumption is only 1.36 watts. This is well within the power capacity of both the power amplifier and transformer (110 and 4.7 watts, respectively). The sensing element of the OHW sensor is a 3.75 rim-diameter tungsten wire. The two ends are electroplated with copper and soldered onto the probe prongs. The sensing length (l); i.e., unplated portion, of the wire is 0.8—1.2 mm and is centered in the middle of the 4 mm-gap between the prongs. The resulting I/d ratio (where d is the wire diameter) of the wire is greater than 200. In implementing the OHW sensor, one concern is that vibrations of the tungsten wire could be excited by the imposed wire motion. Such vibration will not only result in measurement inaccuracy but also in shortening the life of the wire. In order to verify that 52 the wire did not vibrate during oscillation, a microscope composed of Sony XC-75 CCD camera and VZMTM Model 1000 lens, with a magnification factor that is adjustable in the range 2.5X - 10X, was used to visualize the oscillation of the wire. The resulting image resolution was estimated to be 1.68 pm. That is, the system is capable of detecting oscillation amplitudes of the order of the wire diameter. An MVS-2060 fiber-optic strobe light that can be flashed in the frequency range of 0-60 hertz was used to illuminate the wire during oscillation in order to provide recording of an alias of the oscillation frequency so that the high frequency oscillation could be observed visually. Figure 3.7 shows a sample sequence of images captured during the sensor oscillation. In Figure 3.7(a), the wire clearly remains straight, moving like a rigid beam. Figure 3.7(b), on the other hand, depicts a case where substantial wire vibration was seen at an oscillation frequency of 4000. The wire oscillation is excited because the oscillation frequency of the piezo-element is close to the resonant frequency of the wire, which depends on the residual stress of the wire and the plated and unplated length of the wire. The problem could be easily remedied by changing the oscillation frequency or replacing the wire. 3.3 OHW Response Although typical hot-wire sensors are known to have wide bandwidth of 20 — 40 kHz, evidence of substantial reduction in the dynamic response of the hot wires near zero mean-flow velocity exists in the literature. Elger and Adams (1987) used a calibration facility that created an oscillating, zero-mean flow by a reciprocal motion of a piston driven by a loudspeaker in a cylinder. In their experiments, the flow oscillates at 25, 45.2and 70 hertz. And the oscillation velocity amplitude is in the range 0.22 — 5.60 m/s. 53 The dynamic Nusselt number was calculated from the measured anemometer output voltage. The mean value of the dynamic Nusselt number corresponding to zero-mean flow increased with fluid motion amplitude, indicating that the heat transfer mode is not purely natural convection. This suggested that a dynamic calibration is needed for the oscillating flow measurement. But the authors also concluded that the peak flow velocity of an oscillating flow could be accurately obtained from a steady calibration. Perry and Morrison (1971) also found that the dynamic calibration method was more accurate and consistent than the conventional procedure for measurement of absolute turbulence intensities. Therefore, with the use of kilohertz OHW, it is important to calibrate and check the hot wire ability to respond to the imposed oscillation in a controlled environment where the mean velocity could be adjusted to very small value and the direction of the flow could be reversed. The Couette flow facility mentioned in Section 2.4 served this purpose. An adapter plug as shown in Figure 3.8 is used to adapt the curved top surface of the OHW sensor plug into a flat surface to match the shape of the inner surface of the stationary disc of the Couette flow facility. Figure 3.9 shows the main dimensions of the plug. The OHW sensor plug is inserted into the slot of the adapter and aligned with reference marks as Shown in Figure 3.8 to make the top surface of the sensor plug flush with that of the adapter. Once aligned properly, the sensor plug is fixed with two set screws fiom opposing sides. Two pieces of 66-um thick kapton tape were used to cover any discontinuity between the sensor and adapter plugs at the edges of the slot. 54 (a) Wire is straight during oscillation (dash line is used as a staight-line reference) Figure 3.7 Video image sequences of the hot wire during oscillation: (a) Wire is straight during oscillation, (b) Wire bends during oscillation 55 OHW sensor Kapton tape Reference mark Plug Set screw R5080 ’. , . . \ $2.84 I ‘ . ' , . ' §\\\‘ 3 evenly ’ ' I N spaced ' *I 4.76 holes 7.62 14.28 Front View Side View Figure 3.9 Dimensions of adapter plug for calibration of OHW sensors (dimensions in mm) To examine the response of the OHW sensor, it was first desired to verify that the imposed oscillation does not influence the wire calibration; i.e., the oscillation merely results in linear amplitude modulation of the sensor output. To this end, the wire was first kept stationary and calibrated over the velocity range of 0.1 - 4 m/s. Subsequently, the calibration was repeated when the OHW was driven by i 30 Volt ac signal while 56 varying the frequency from 1.1 to 4.8 kHz. Figure 3.10 shows the hot-wire calibration obtained without wire oscillation (solid line), compared to that obtained from the mean hot-wire voltage for wire oscillation frequencies of 1.1, 2, 3 and 4 kHz. The no- oscillation calibration extends down to a velocity of 0.1 m/s. On the other hand, data obtained during oscillation are shown for a velocity range that is larger than 0.3 m/s. In this range, a good agreement is evident between the results with/without oscillation. This verifies that the high-frequency small-amplitude oscillation of the wire does not result in an anomalous or non-linear behavior that may affect the Wire’s calibration or its accuracy in retrieving the velocity magnitude. 0.95I STAT f=1100 Hz f=2000 Hz f=3000 Hz f=4000 Hz ’50” 0.9 I I I Irf+ 0.85»- 0.75» I J 0.7l i 1 1 4' 0 1 2 3 4 5 U(m/s) Figure 3.10 Magnitude calibration of the hot wire with/without oscillation To examine the unsteady response of the wire to the imposed oscillations at different frequencies and flow velocities, the mean-removed hot-wire output was 57 acquired and phase averaged with respect to the OHW driving voltage. The phase average aids in removing the effect of any background disturbances related to motor vibration influences or other sources of noise. A portion of the resulting signal is shown in Figure 3.11 for an oscillation frequency of 2 kHz and different mean velocities in the range 0.34 — 3.8 m/s. The reduced dynamic-response ability of the wire with decreasing speed is evident from the data as the amplitude of the measured oscillation velocity (um) is attenuated and an increased phase lag is observed with the reduction in the mean flow velocity. 0.1514 .. . U=0.34m/s uo,m I u=o,53m/s 0.1~ U=0.8lm/s (m/s) « U=l.20m/s I U=l.62 m/s . . U=2.03 m/S ‘ . U=2.49 m/s ‘~ U=2.9l m/s 7 U=3.36 m/s U=3.80 m/s 0.05‘» -o.05:7 -o.1I,V‘ -o.15. 'o' ' 7 "19” W ’26 T .30-.. " 40 SampleNumber Figure 3.1 1 Phase averaged hot-wire output at the oscillation frequency For flow speeds larger than 1 m/s, um is fairly constant. However, when the flow velocity drops below 1 m/s, larger attenuation is found. The reduction close to zero flow velocity is presumably due to shifting of the primary heat transfer mode of the wire from 58 forced to natural convection and increasing importance of heat transfer through the prongs by conduction, leading to a larger thermal time constant. Although the Constant Temperature Anemometer (CTA) circuit compensates for the sensor’s thermal response, the compensation generally works well for flow velocities larger than roughly 1 m/s where the thermal time constant apparently does not vary significantly. Below 1 m/s there seems to be an appreciable change in the time constant, for which the CTA does not adapt. The reduction in the dynamic response should not cause a problem in operating the OHW sensors as long as the amplitude attenuation is not too large for the Sinusoidal modulation of the wire output to be detectable at very low velocities. In Figure 3.12, the modulation Signal at the oscillation frequency was detectable at flow velocities as low as 0.03 m/s. In addition to amplitude attenuation, as the flow velocity decreases, a phase-angle delay is observed between the imposed and measured oscillation. The delay becomes progressively larger with decreasing flow velocity. Similar effects on the phase response of hot wires were also made by Egler and Adams (1989) and Antonini et al. (1976). They explained the reason for the phase lag was because the temperature distribution around the wire could not follow the wire oscillation. In Figure 3.11, the biggest phase lag is less than 90°, which is found at the lowest flow speed. It is well known that the thermal response of the hot-wire sensor could be modeled as a first-order dynamical system. Even as the oscillation frequency approaches infinity, the phase lag reaches only —90°. 59 (a) Modulated velocity signals —e~ Positive I I ~ _._ Negative r C‘F‘i‘fiif‘ i CJMCC' u 0 6 .. or" n.) , “XII“ : C . ............ u ———————————— a ..................... 1:...pr ........... (gen—I N,‘ I ‘ g; Cir". ,l/./‘ .‘ ‘., ,1 Q :2 h /. '\. a J _ (m/s) . - 1,. . U— 0 55 m/s at - _ 1. . ~ ‘ o O 5 t- . . 13:72. _ .. ...... .o 3 ..{‘.’29,.. . “of.“ I .. fl” .'. ... .. I 1:78 ' ~‘l . . “k¥;7f;7 . . T F “fli'fifrk ‘. W .. ‘e ,..k ' 9‘ ‘~- ‘A "‘ 0.2 -~~ 3'81: ----, lCWC sat-v“ t ...................... ......... .4—0—0 44‘. ‘ I \‘ 9,3..332; ' ’0 .. > 4 I.» 1““ _ 0" 0‘ _ “e. ., - e “a e ‘e-" U — 0 03 m/ 0' .~::»3'~«”~4:..E.~i*’9‘*~e - . -' ~‘ *“*’“""ff~»-'«-«.:j - en ‘6’ “’1‘ ” e" W’ "fir ' S . ‘trf'fl - - ‘yo‘e‘ ‘ (b) OHW driving signals 0 . 8 ______.__ e d I (V) Sample Number m_fi__ _.__~_._ _ ,.._ —-:::+— POSIIIVG I _._ Negative ‘——- v” Sample Number Figure 3.12 Demonstration of the effect of the flow direction on the phase of the OHW modulation signal: (a) Modulated velocity Signals, (b) OHW driving signals 60 3.4 OHW Measurements (Phase Method and Phase Calibration) In Section 3.2, it was demonstrated how the flow direction information is extracted from comparison of the forward and backward measurements. However, as seen from the previous section, the dynamic response of the OHW sensor leads to a phase shift at low flow velocities, which further results in smaller difference between the forward and backward measurements. This may be seen from Figure 3.11: the two dotted lines indicate the oscillation phase corresponding to the forward and backward measurements, respectively. When the flow velocity is U = 0.34 m/s, the difference between the two measurements is only about 0.08 m/s. In a real flow situation, this small difference may be easily masked by unsteady flow-velocity variation. Moreover, because of the amplitude attenuation of the OHW signal at low velocity, it is not possible to obtain the exact magnitude of the flow velocity by subtracting the oscillation velocity from the measured cooling velocity as outlined in section 3.2. To overcome the above problems, a different method was used to recover the flow information from the OHW time series. In particular, it can be shown as follows that the phase difference between the wire output and the piezo-driving voltage, hereafter denoted by mm at the oscillation frequency experiences a 180° jump when the flow switches direction. In Figure 3.1, the instantaneous effective cooling velocity of the wire is: luf+ Unsin(27y'ot)| (The absolute value is used here because a hot-wire sensor can’t determine the flow direction.) If the flow velocity is positive, the measured velocity is W + Uosin(279’0t). On the other hand, the measured velocity is l-uf + U,sin(279€,t)| = luji - Uosin(279’0t) if the flow direction switches. There is 1800 phase difference between the former and latter measured velocities. 61 To demonstrate the concept, the OHW sensor was used to take measurements at different velocities under clockwise as well as counter—clockwise rotation of the rotating disc in the Couette flow facility. The direction of flow velocity in this facility depends on the rotational direction. Two different rotational directions produce two opposite directional flows. The results are plotted in Figure 3.12 for the velocities of U = 0.03, 0.2 and 0.55 m/s. For each speed, two data sets are plotted (open circles and dots) representing two opposing flow directions. For orientation purposes, the bottom portion of Figure 3.12 contains a plot of the driving voltage of the OHW. The data clearly demonstrate the reversal of phase of the wire sinusoidal modulation signal as the flow changes direction for all flow speeds. It is noted here that for the lowest flow velocity and negative direction, small (about 0.01 m/S) negative values of the sensor output are seen in Figure 3.12. Since this is the “raw” output of the sensor before demodulation and extraction of direction information, no negative values should be observed. The discrepancy is of the order of the accuracy of the calibration process in the vicinity of zero velocity. From Figure 3.11, it is evident that in a given flow direction, the phase of the oscillation in the hot-wire output changes with decreasing velocity. Because this phase change does not exceed 90°, it does not create a problem in direction determination via the phase method. In other words, in the absence of phase lag, rial... would jump from, say, 0° for positive to 180° for negative flow velocity. This would be the same in the actual situation at high flow velocities where no limitation on the Wire’s dynamic response has been observed. On the other hand, at very low velocities, or wall shear stress, values, Ito)... would jump from 90 to 270 instead of 0 to 180 degrees because of the 62 phase lag. Albeit this change, positive and negative flow directions could still be distinguished without ambiguity since in one direction e50)", falls in the range 0 — 90 degrees, which does not overlap with its range in the opposite direction (180° - 270°). A problem would have been encountered if the Wire’s phase lag at low Speeds exceeds 180 degrees, resulting in an overlap in Ito}... values in the positive and negative directions. However, as mentioned before, the wire’s phase lag is limited to less than 90°. Itch... is unique for each OHW sensor because it depends on the dynamic response of the piezo-element as well as that of operation of the anemometer and the hot-wire sensor at different velocities. Therefore, Ito)... response must be verified for individual sensors by calibrating against the shear stress in the positive and negative directions. An example of such a calibration for a wall-shear OHW sensor is shown in Figure 3.13. For reference, the ideal ¢0hw response, exhibiting 180° phase jump with flow reversal, is also illustrated in the figure using a broken line. It is clear that 410;... values in one direction do not overlap with those in the other, confirming that the direction can be determined without ambiguity. In implementation, ¢ohw is found from calculation of the cross spectrum between the piezo-driving signal and the hot-wire output at the oscillation frequency for every cycle of oscillation. The process of detecting the flow direction is simplified by realizing that Ito}... values in the positive shear direction correspond to a negative real component of the cross spectrum, and vice versa. Thus, for every cycle of oscillation, the Sign of the real part of the cross spectrum is used to determine the flow direction. Finally, it should be noted that it is believed that the deviation between the ideal and actual ¢ohw behavior (see Figure 3.13) is primarily caused by the limitation of the wire dynamic response near zero velocity, as discussed above. In addition, it is also 63 believed that at these low velocities, low noise levels at the oscillation frequency (resulting from capacitive coupling of the driving signal to the hot wire prongs) may influence the imposed modulation of the wire output. Understanding of the latter requires additional investigation, which was not done here since the observed at»... response indicate that the problem does not result in ambiguity in shear direction determination. 200;? " I f ——. 150 100» 1w (Pa) Figure 3.13 A typical direction calibration of the OHW sensor To determine the magnitude of the velocity from OHW measurements, the sensor output was low-pass filtered to remove the sinusoidal modulation due to wire oscillation. The cut-off frequency of the filter was set at half the oscillation frequency. Since the latter is also the sampling frequency of the shear information, the filter also prevents aliasing due to subsequent downsampling of the data. The low-pass filtered signal is then converted to wall-shear stress using the magnitude calibration. Figure 3.14 provides a block diagram summarizing the overall operation of the OHW sensor. 64 Low-Pass Filter I: E : Ind f,/2! K _. U @ Constant E Temperature Phase f0/2: IM Detection UrDirection I (Cross Spectrum) - - H' -P F'lt INWVI Ojglilatw” 1gb ass 1 er To Oscillating river Oscillation Driving Signal Support Figure 3.14 OHW Operation schematic diagram 3.5 Effect of Wire Height above the Wall For a wall hot-wire sensor to measure wall-shear stress, the wire must be located within the viscous sub-layer; i.e., y+ < 5 (or y < 5v/u,). Referring to the measurements by Tihon et al. (2001), Spazzini er al. (2001), and Jovic and Driver (1995), the maximum reverse and forward skin-friction coefficient is less than 0.004 at the highest Reynolds number. For a given step height, the highest Reynolds number corresponds to the largest wall shear and friction velocity, and therefore the largest y+ for the same distance (y). Consequently, the largest wire height above the wall in viscous units is actually found at the highest Reynolds number. Combining U... = 15 m/s, which corresponds to the largest Reynolds number in the present study, with Cf = 0.004, the maximum wire height was estimated to be 112 pm. This provided a guide for initial placement of the wire above the wall since the fiiction velocity was not known a priori. 65 In order to experimentally check proper positioning of the wire above the wall, measurements were conducted with the height of the sensor set to three different values: 48, 96 and 124 um. The x location for the measurements was chosen to be well beyond the expected x, position (x = 9821-!) such that t... has recovered near its largest value after reattachment. The freestream velocity was set to 15 m/s, corresponding to Reynolds number based on step height of Re” = 13000. At each height, the experiment was repeated for three times with the wind tunnel turned off in between runs. This was done in order to compare the agreement between the data for different wire heights against the measurement repeatability from a given height. The data were acquired for one minute for each run. No wire oscillation was imposed during this set of experiments since the measurement location is well downstream of the expected reattachment position, and hence there is no directional ambiguity of the flow. The work of Tihon et al. (2001) and Spazzini et al. (2001) provide supporting evidence that the flow at the chosen measurement location is always in the downstream direction. More importantly, the flow-direction—probability results from the current study, given later in 4.2.4, verify that the flow is in fact always in the downstream direction at x = 9.82 H. Figure 3.15 and Figure 3.16 show the power spectra and normalized histograms, respectively, of the wall-shear stress measured from three repeated runs at y = 48 pm. The normalization of the histogram refers to the fact that the sum of the histogram values equals unity. The data show that the power spectra exhibit a high degree of repeatability. On the other had, some scatter is observed in the vicinity of the histogram peak. However, the mean values of the wall-shear stress obtained from the histograms are 0.277, 0.279 and 0.282 Pa, respectively. The differences in these values are less than 66 1.8%. This small difference and the repeatability of the spectra demonstrate the precision of the current measurement procedure. Figure 3.17 and Figure 3.18 display the power spectra and normalized histograms for data acquired at three different wire heights, respectively. The data corresponding to the three different cases collapse well in both plots. The mean values of the wall-shear stress calculated from the histograms are 0.277, 0.276 and 0.275 Pa. The small variation of the measurements with the wire height indicates that all three positions are within the viscous sublayer, where the velocity is linearly proportional to the wall shear stress. In fact, if the highest observed '1:w value of 0.282 Pa is used, one wall unit will be 31.3 urn. The corresponding wire heights (y+) are equal to 1.5, 3.1 and 4.0; i.e., all positions are within the viscous sublayer. Based on this, the wire heights were set at 82 and 97 pm for the two OHW sensors used in the two-point measurements. .Lg I 1O _.4 —A7.—¥ . _A.~._._.~ ~_‘._‘_A_I.' ,,.. _..._.. 'h-CL #4."; _..-_L i. 2'. —A—. ~—.‘__‘._A_ _L..#1 1o 10 1o3 10 f(Hz) 4 Figure 3.15 Power spectra of three different runs for wire height y = 48 um 67 r (Hz) Figure 3.17 Power spectra for wire heights y = 48, 96 and 124 um 68 I‘ +h=48pm I +h=96um I —- h=124umI 0.3;» MI ._ ,_. ._._...... Histogram O L. .O 9!- - N .0 A 0.05’ Figure 3.18 Histogram for wire heights y = 48, 96 and.124 urn 3.6 Effect of Wire Oscillation on the Flow Although highly unlikely, it seemed appropriate to check whether the oscillation of OHW sensors disturbs the unsteady direction-reversing flow or not. For this purpose, data obtained at ReH=13000 and x/x, = 0.06, 0.92 and 1.54, i.e. before, close to and after the mean reattachment point, were utilized. At each location, measurements were conducted when the oscillation was turned on as well as off. The former case corresponded to an OHW sensor with the oscillation frequency of 2.8 kHz, and the latter represented a conventional hot-wire sensor. The premise of the test is based on the fact that notwithstanding its directional blindness, the conventional hot wire output voltage should provide the correct magnitude of the wall shear stress. Therefore, the low-pass- filtered voltage signal of the OHW, which also provides the magnitude information of the wall shear, should yield the same voltage output magnitude as the conventional hot wire. 69 If the OHW operation does disturb the flow, then one would expect that the frequency spectrum, or some other statistical measure, of the low-pass-filtered OHW voltage would deviate from that obtained with the stationary wire. I Figure 3.l9(a), (b) and (c) show a comparison between the power spectra of the raw voltage data of the oscillating- and conventional-hot-wire sensors, labeled as “HW” and “OHW” in the legend. A low-pass filter with a cut-off frequency of 1.4 kHz, or half of the oscillation frequency, was also applied to the OHW signal. Since the sampling frequency of the OHW sensor is also the oscillation frequency, this filtering eliminates any possible aliasing in the OHW signal. The spectra obtained fi'om the filtered signal are also included in the plots in Figure 3.19. Inspection of the figure clearly shows that within the pass-band of the filtered signal, the voltage spectrum of the OHW and HW sensors agree quite well for all three measurement locations. The imposed oscillation on the OHW sensor is also evident from the harmonic peak at f = 2.8 kHz. Therefore, one may conclude that the oscillating hot wire operation does not generate any flow anomaly. It is also noteworthy that the energy level of the oscillation signal is at least two orders of magnitude larger than that of the flow at the oscillation fi'equency. This yields a strong signal-to-"noise" ratio, which is necessary for successful determination of the flow direction. Another important observation in Figure 3.19 is that the “flow” energy contained in frequencies above 1 kHz is only very small component of the total energy. This may be seen fi'om the fact that the Spectrum level at frequencies higher than 1 kHz is two orders of magnitude or more lower than the spectral peak value. Hence, the selected oscillation frequency should be high enough to resolve the wall-shear signature of the 70 most energetic flow structures. (a)%r = 0.06 .4 10 :--l,-._---- . ~ . o HW .5 g — OHW 10 I + OHW Filtered .5 Z $107 I (210“3 .. $103 . 104°: .11: 10 .- 10° 10"»F ,W - .-..H, fizz .mvflzuz ..,-. fizz..- 7. .. _..E.I I O HW I 10.5 ' — OHW I 2‘ + OHW Filtered 3. .5 3 NA‘IO I 2, 2 E - . 2107 g I ‘21043 i a) if 7 g _ : : 8‘109 10'”; . l -11I - 10 . ----n---------_--------- w 10° 0 102 103 10‘ f(Hz) 10“ {E e We ___.r__._ a -.j I O HW ; «6 l — OHW .- 10 E + OHW Filtered I -6 7‘ . «“10 I E 2 E 3 gm” «(310“3 i . gl £10 I 104°?» 1 10'11I i _r A l A lint , --.M.__+ HEMI 10° 101 102 103 10‘ f (Hz) Figure 3.19 Comparison of power spectra of OHW & conventional hot wire (HW) at different x positions: (a) x/x, = 0.06, (b) x/x, = 0.92, (c) x/x, = 1.54 Further comparison between the oscillating and stationary hot wire measurements was conducted by comparing the normalized histogram results obtained in both cases. As evident in Figure 3.20(a), (b) and (c), the “OHW” and “HW” histograms are consistent with each other, at the same x locations used for the spectral comparison in Figure 3.19. This suggests that the two measurement methods produce the same statistics of the voltage, and hence shear-stress, magnitude. In summary, the above comparison demonstrates the ability of the OHW to operate successfully without disturbing the flow field. Moreover, the selected oscillation frequency appears to be high enough to capture the majority of the energy in the flow. 72 —i— OHW Filtered if —e— HW —l— OHW Filtered ,,, . ,_ fi—r Histogram 73 Y—e—HWY —+— OHW Filtered E 0.5 4 0.4» E a 30.3 E I 0.2» .0 _a. 0.9 (9.75 3 E (V) Figure 3.20 Comparison of histogram of OHW & conventional hot wire (HW) at different x positions: (a) x/x, = 0.06, (b) x/x, = 0.92, (c) x/x, = 1.54 3.7 Specification of OHW Sensors Use in the Present Experiments Two OHW sensors (OHW 1 & 2), similar to that pictured in Figure 3.4, were developed for the current study. The height of the wire above the wall for OHW] and OHW2 was set to 97 and 82 pm, respectively. The length of the sensing portion of OHWl sensor was 1.259 mm, and its cold resistance was 6.01 D at room temperature of 22.7 °C. Similarly, the active sensing length of OHW2 sensor was 1.0 mm with cold resistance of 5.26 0 at room temperature of 22.8 °C. The overheat ratio of the two sensors was 1.6 and the oscillation frequency was set to 2800 hertz. As shown in Figure 3.21, the two sensors were assembled together with spacers to form a sensor cover fitted into the slot behind the BFS (see Figure 3.22). One of the 74 sensors was used as a reference sensor by keeping its x location fixed for all measurements. The other one could be mounted at seven different x locations on one side of the reference sensor by using different-width spacers. By flipping the plug 180 degrees, seven additional positions could be easily obtained. Table 3.1 lists the width and arrangement of the spacers and sensors corresponding to the seven positions of the movable sensor. In the table, M and R are used to denote the movable and reference sensors, respectively, and the width of each element in mm is given in brackets. Table 3.2 lists the measurement locations for the movable sensor normalized by the step height. The corresponding reference sensor location is 5.17H. OHW sensor OHW sensor (movable) (ref) Sensor cover 1L Spacer 59.69 127.0 Figure 3.21 One of the combinations of the sensor cover (in mm) 75 OHW] OHW2 Sensor cover Figure 3.22 Picture of the sensor cover during insertion in the test model Table 3.1 Combinations of the sensors and spacers corresponding to different measurement locations of the movable probe Combinations for the sensor plug Spacer 1 Mgfiazblle Spacer 2 Rigegce Spacer 3 1 M (7.62) Spacer (52.07) R (7.62) Spacer (59.69) 2 Spacer (13.97) M (7.62) Spacer (38.10) R (7.62) Spacer (59.69) 3 Spacer (21.59) M (7.62) Spacer (30.48) R (7.62) Spacer (59.69) 4 Spacer (29.21) M (7.62) Spacer (22.86) R (7.62) Spacer (59.69) 5 Spacer (36.83) M (7.62) Spacer (15.24) R (7.62) Spacer (59.69) 6 Spacer (44.45) M (7.62) Spacer (7.62) R (7.62) Spacer (59.69) 7 Spacer (52.07) M (7.62) R (7.62) Spacer (59.69) 76 Table 3.2 Measurement positions of the movable sensor Position 1 2 3 4 5 6 7 x/H 0.31 1.45 2.07 2.69 3.31 3.93 4.55 Position 8 9 10 11 12 13 14 x/H 5.79 6.40 7.02 7.64 8.26 8.88 10.02 77 CHAPTER 4 ONE-POINT MEASUREMENT RESULTS This chapter contains two main sections: one on the boundary-layer velocity profiles at separation, and the other containing a discussion of the single-point characteristics of the wall-shear stress beneath the separating/reattaching flow. In the second section, wall-shear results pertinent to the reattachment length, mean/rms skin- friction coefficient, probability density function, power spectra, and autocorrelation are discussed and compared with the literature. 4.1 Boundary-layer Velocity Profiles at Separation The boundary-layer characteristics at the point of separation have significant effects on the separating/reattaching flow over a BFS, as demonstrated by Eaton and Johnston (1981), Armaly et al. (1983) and Westphal and Johnston (1984). For example, the boundary-layer thickness weakly affects the reattachment length, while the boundary layer state (turbulent or laminar) has a significant effect on the reattachment length. In the current study, the mean and ms streamwise velocity profiles (u) were documented at a location that was 0.5 mm (or approximately 15-20% of the boundary layer momentum thickness for all three Reynolds number investigated) upstream of the step. Figure 4.1 displays the mean velocity profiles at three different velocities: 5, 10 and 15 m/s. The distance above the wall (y) plotted on the ordinate is normalized by the momentum thickness (0), which is defined as follows: 78 _ My). -2312. i9— [EDT/:[l U00] dy (4.1) where u(y) is the mean velocity at position y. The value of 6? was found by numerical integration of the measured velocities using the trapezoidal rule (also known as Newton- Cotes formula). The outcome is shown for all cases in Table 4.1. The corresponding Reynolds numbers based on 6 are also listed in the table, along with the boundary-layer thickness (6), displacement thickness (6') and shape factor (HI; = 576). Table 4.1 Parameters for the boundary layer at separation U..(m/s) 5 10 15 5(mm) 28.5 24.5 21.6 6' (mm) 4.03 3.54 3.05 0(mm) 3.12 2.77 2.40 H]; 1.29 1.28 1.27 Reg 1139 2057 2739 As seen from Table 4.1, the Reg range of the boundary layer is on the lower end required for the establishment of a turbulent state. The lowest known Reg value of a turbulent boundary layer is 670, which was obtained in the first Direct Numerical Simulation of a turbulent boundary layer by Spalart (1988). In the laboratory, Naguib (1992) studied turbulent boundary layers that spanned a Reg range starting from approximately 1500. The boundary layer profiles shown in Figure 4.1 have large velocity magnitudes in the near-wall region (more than 40% of the freestream velocity at the lowest y position 79 of 0.355 mm). This is reminiscent of the 'fuller' nature of turbulent velocity profiles, where the more effective momentum transport across the layer, brought about by the turbulent eddies, causes penetration of the high-momentum fluid deep into the boundary layer. A comparison between one of the current profiles with a Blausius and turbulent velocity profiles at a similar Reg is shown in Figure 4.2. The data demonstrate the turbulent character of the separating boundary layer in this study. aoI—u- * ~- - E Q l _e.Re=1139 \ 25I 129-2057 A i —1— e6- _A_Ree=2739 20+ 15} l 10E 5:» of __ 022"” 04 Figure 4.1 Mean velocity profiles of the boundary layer at separation The turbulent nature of the separating boundary layer may also be put on quantitative basis by considering the boundary layer shape factor (H12). Hg is a useful quantity to describe the shape of velocity profile. For example, in a zero-pressure- gradient turbulent boundary layer, the shape factor is given as 1.3 in Schlichting (1979) and 1.31 in Morris (2002). For a laminar boundary layer, H12 is 2.59, which is obtained from the Blasius solution. The shape factor values for the boundary layers established in 80 this study are nearly equal to those fi‘om Schlichting ( 1979) and Morris (2002) (see Table 4.1). They are also in good agreement with those in some of the BFS flow studies: 1.29 in Chun and Sung (1996) and 1.26 in Baton (1980), as examples. 301-——-~ ~—.————— --___x-_ W“ I _____ I -e.- Ree=2739 . Q 25; O Reef2953, Nagu1b(1992) I >5 . "‘— Blasms J zoI 15$ U/Um Figure 4.2 Comparison of a mean velocity profile from the current study to that of turbulent and laminar boundary layers. Conventionally, the friction velocity (u, = 41,, / p ) is used to non- dimensionalize the mean velocity (U+ =U/u,) and the inner, or viscous, length scale to normalize y ( y+ = yur /u) in plots of the boundary-layer profiles. A 'canonical', or 'fully-developed', turbulent boundary layer is known to posses the so called 'logarithmic region' where the mean-velocity varies with height according to: U + = 2.44 ln(y+ ) + 5.0. The constants in the last equation are determined empirically. Generally speaking, there is a good deal of scatter in the values reported in the literature, with 2.44 and 5.0 being 81 the most commonly cited ones. The primary difficulty in obtaining the constants of the 'log law' is the ability to measure the wall-shear stress, or fiiction velocity, accurately. One typical, albeit inaccurate method, to determine ur is known as the Clauser fit. In this procedure, the log law is assumed to have a known value for the slope or intercept, which leads to the value of ur by comparing to the slope or intercept of a least-squares line fit to the logarithmic portion of the mean velocity profile. However, the accuracy of the method is generally no better than 10 - 15% because of uncertainty in determining the limits of the log region. Recently, Osterlund et a1. (1999) conducted careful measurements of the wall-shear stress in a turbulent boundary layer over a wide Reynolds number range using oil-filrn interferometry. Their results showed that appropriate implementation of the Clauser-fit requires the fit to start above y+ = 200. This would leave too little data to obtain a reliable fit at Reynolds numbers comparable to those investigated here, and therefore this method was not employed. No attempt was made to measure the wall-shear stress at the point of separation directly, since the main focus of the current study is the separated region downstream of the step rather than the separating boundary layer. In fact, Morris (2002) showed that in geometrical separation of turbulent boundary layers, a 'sub-shear layer' controls the dynamics and subsequent development of the flow downstream of separation. Only the near-wall vorticity of the boundary layer participates in the initial instability of the sub- shear layer. An effective momentum thickness of ten wall units was found to characterize the width of the region significant to the separating-shear-layer instability. This is well below the log region, and hence the specific constants of the velocity profile are not expected to have important bearing on the separated flow. However, for the 82 purpose of qualitatively examining the behavior of the current boundary layer profiles under the traditional scaling, the friction velocity was estimated using the empirical correlation of Ludwieg and Tillmanm (1949) for the skin-friction coefficient for an equilibrium turbulent boundary layer (see Hinze (1959), page 486): C f = 0.246x 10‘0-678” 12 Ref-263 (4.2) The mean velocity profiles are normalized with the fiiction velocity obtained using Equation 4.2 and u7 = ‘/(CfU3, ) / 2 . The results are plotted on a semi-logarithmic scale in Figure 4.3. All profiles seem to collapse fairly well using the traditional inner scaling, with the exception of data near the edge of the boundary layer where the classical Reynolds-number-dependent deviation from the log law (know as the 'law of the wake') is observed. The data also exhibit a 'log-law—like' behavior from approximately y+ = 30 to 300. The best fit to these data is u+ = 2.141n(y+) + 5.65 , which yield constants that are about 13% off fiom 2.44 and 5.0. The deviation may be attributed to a number of reasons, including the error in determining the friction velocity, the transverse curvature effects of the axi-symmetric shape used in the current study, or the development length of the boundary layer, which is shorter than that employed in most studies aimed at establishing equilibrium turbulent boundary layers. Finally, the streamwise rms velocity profiles are shown in Figure 4.4 for all Reynolds numbers. The data exhibit the classical behavior of a turbulent boundary layer with the largest rms value located very close to the wall. Collectively, the results discussed in this section indicate that the separating boundary layer is turbulent. It is uncertain, however, whether it is in what is considered in the literature to be a 'canonical' 83 state. As discussed above, this issue is not of central significance to the objectives of the current study. To Ree=1139 + Ree=2057 201 A Ree=2739 — Fit 10E . l 5» \ I u+ =2.14ln(y+)+5.65 1 01-0 . ...2 22.2...2222i 2222222222...” 2___2.. 2.2.2.... 222.23 222—22- .. 2.2.22.2 4 10 10 10 10 10 y+ Figure 4.3 Inner-scaled Mean velocity profiles of the boundary layer at separation 30, __. 2 2. 2 2 2 2 22 , 2 22 m . —e- Ree=1139 R 25} _+_ Reo=2057 I .4. Ree=2739 ..k. . 2 0 " 0.02 0.04 “6.06“ ." urine/U00 Figure 4.4 Rms velocity profile of the boundary layer at separation 84 4.2 One-point Wall-shear Measurements in the Separated/reattaching Flow 4.2.1 Mean Reattachment Length At the mean reattachment location (x,), the mean wall-shear stress equals zero. Thus, the streamwise distribution of the mean skin-friction coefficient (Cf) could be used to directly find x,. Using the results from the movable sensor at different streamwise locations over the range x = 0.3OH to 9.84H (the specific positions are specified in Table 4.2 section 3.7), Figure 4.5 was obtained for three Reynolds numbers of Re”: 4300, 8700, and 13000. Note that the plot also contains data from the reference sensor at x = 5.07H. The vertical axis represents the mean skin-fiiction coefficient, and the horizontal axis indicates the corresponding normalized x location. The data for all Reynolds numbers have the same trend. -3 5x10? f 4 22.22.22.222 —9— ReH = 4300 t . 4:2. ReH = 8700 c = ; 3,_,_ ReH=13000 ,, ’ I 2. 0 ' 0‘1 0 e" -1. » -2. -31.. 22. . 2 2 222 _222 2 0 1 5 6 7 8 9 10 x/H Figure 4.5 Streamwise distribution of the mean skin-friction coefficient 85 The mean skin-friction distribution in Figure 4.5 displays two positions where Cf crosses zero. The first position close to the step defines the location of the mean secondary separation point or corner eddy length as discussed above. The position, where Cf: 0 at the farther downstream location, gives the mean reattachment length. The values of x, for the Reynolds numbers investigated are tabulated in Table 4.2. Another method to find the mean reattachment point is to use the distribution of forward/reverse flow probability, as demonstrated in Baton (1980), Westphal and Johnston (1984), Spazzini et al. (2001) and Tihon et al. (2001). For example, a forward flow probability (hereafter referred to as FFP) of 100% means the flow is always towards the downstream direction. On the contrary, FFP = 0% means the flow is always towards the upstream direction. At the mean reattachment point, FFP = 50%. Figure 4.6 displays three curves representing the FFP distribution for all three Reynolds numbers, which exhibit the same trend. Consistent with the distribution of the mean skin-fiiction coefficient, there are also two positions where F FP = 50%. The farther downstream one indicates the mean reattachment location, the values for which are listed in Table 4.2. As depicted in Table 4.2, the difference in determining the reattachment length using the Cf or FFP is less than 1.5%. The reattachment length slightly increases with rising Reynolds number over the range of investigation. The same trend was also found by Spazzini et al. (2001), where Rey was varied from 3500 through 16000. The increase could be attributed to the variation of the boundary-layer thickness. Both Eaton and Johnston (1981) and Adams and Johnston (1988) found that the reattachment length increased slightly with decreasing boundary-layer thickness. Since the Reynolds number is increased by increasing the freestream velocity in the present study, the thickness of 86 the boundary layer at separation does in fact decrease with increasing Reynolds number (the specific values of 6 may be found by referring to Table 4.1). 100I— * FFP (%) 0| 9 25 —e— Re=4300 I + Re=8700 I —l—- Re=13000 0! — 22 — 2 — 4— 0 1 2 3 4 5 6 7 8 9 10 Figure 4.6 Streamwise distribution of forward flow probability Table 4.2 Reattachment length values x/I-I xr/H x,/H . ReH = = o Drfference (%) (Cf 0) (FFP 50 /o) Mean 4300 4.26 4.33 1.43 4.30 8700 4.60 4.67 1.45 4.64 13000 4.84 4.88 1.00 4.86 Another important observation is that the reattachment length (x,/H) reported here is smaller than that in the bulk of the literature; for example, x/H = 5.39 at Re” = 5100 (Spazzini et a1. (2001)), 5.1 at Re” = 4800 (Tihon et al. (2001)), 6.00 at Re” = 5000 87 (Jovic and Driver (1995)) and also smaller than the minimum x, (4.9 H) reported in the review by Eaton and Johnston (1981). But in all of these studies, either a 'two- dimensional' planar BFS geometry or sudden expansion of a pipe was used. Although the flow in the sudden expansion is also axi-symmetric, the geometry has concave curvature, and, the boundary layer characteristics at separation may be different. Therefore, it is possible that the shorter reattachment length found here is caused by the geometry difference. Kim et al. (1996) also found that the reattachment length in laminar boundary-layer separation over an axi-symmetric BFS, composed of an elliptic-cone- shaped obstruction placed in front of a cylinder tube, was always shorter than that of the planar BF S flow. They attributed this phenomenon to the effect of transverse convex curvature. To examine the reasonableness of this hypothesis further, the reattachment mechanism needs to be scrutinized. As argued by Chapman et al. (1958) and Eaton (1980), the reattachment physics could be attributed to the mass balance between the fluid re-entrained into the shear layer and the pressure-driven backflow. For example, if there is more pressure-driven backflow, the reattachment length will increase so that the shear layer has a longer length to entrain the backflow, and the pressure gradient decreases leading to a reduction in the backflow. On the contrary, increasing entrainment rate to the shear layer will decrease the reattachment length, which has been verified in the control experiments of the backward-facing step flow; for example, Chun et al. ( 1999) and Lai et al. (2002). It is conjectured here that because of the axi-symmetric geometry of the BFS, the vortical structures at separation are substantially more coherent in the transverse 88 (azimtuthal) direction than in the planar geometry, and more capable of sustaining higher entrainment rates than in the planar geometry, leading to a shorter reattachment length. In the planar geometry, the transverse coherence, or two-dimensionality, of the vortices is influenced by side-wall effects directly, as well as indirectly, through the imperfect two- dimensionality of the mean flow itself. The coherence (and energy) of the vortices may be enhanced through artificial excitation of the flow. Chun and Sung (1996) employed local forcing to control the turbulent flow separating over a planar BF S. The reattachment length was reduced from 7.8H to 5.0H. It was shown that the forcing enhanced the shear-layer growth rate and produced roll-up of a large vortex at separation, leading to higher entrainment rate and shorter reattachment length. In a similar manner, one could imagine a highly coherent vortex ring shedding from the transverse convex curvature at the step and growing quickly to a large size through pairing interactions, affecting a higher entrainment rate. Another factor to consider in examining x, results is the boundary-layer-thickness to step-height ratio. In the current study at Re” =13000, Reg = 2739, (SH =1.72 and x,/H = 4.88. This may be compared with two well-documented experiments in a planar BFS flow: Re” = 33000, R39 = 1340, (WI =0.41 and x,/H =7.8 for Chun and Sung (1996) and Re” = 39455, Reg = 852, 5’H =0.23 and x/H =7.95 for Eaton (1980). The boundary layer in all studies is turbulent. The current 8’11 is the biggest, so the reattachment should be the smallest. 89 4.2.2 Mean/ms Skin-friction Coefficient and FFP The mean skin-fiiction coefficient distribution is re-plotted in Figure 4.7 after normalization of x using x,. Data from the literature are also added to the figure for the purpose of comparison. The measurements by Spazzini et al. (2001), Tihon et al. (2001), and Jovic and Driver (1995) were obtained using a "double" hot-wire type skin-friction probe, an electro-diffirsion technique and laser-oil-interferometry (LOI) technique, respectively. The first two methods infer the shear stress from heat/mass transfer fiom the sensor and require calibration, while the last one measures the shear stress directly and does not require calibration. LOI is the most accurate wall-shear measurement technique known today, but it can only provide mean-shear measurements. The mean skin-friction plots in Figure 4.7 are divided into two parts: (a) for Re” < 10000 and (b) for Re” >10000, in order to show the data points clearly. Despite differences in the experiments, such as geometry and Reynolds number, all data sets exhibit the same trend. From the separation point to x/x, z 0.33, the value of the mean skin fiiction is very small and slightly above zero. This region is referred to as the secondary recirculation zone, or comer eddy. Kim et al. (1996) studied the axi-symmetric separation and reattachment of a laminar boundary layer and were able to identify the comer eddy only when the diameter ratio of the up and downstream cylinders at the step was less than 2.0. In the current study, the diameter ratio is 1.25, and hence the finding of the comer eddy here is consistent with the conclusions of Kim et al. (1996). 90 (a) Re” < 10000 ReH =4300 ReH =8700 ReH =3500 (Spazzini et al. (2001)) Re” =5100 (Spazzini et al. (2001)) Re” =3860 (Tihon et al. (2001)) Re... =8840 (Tihon et al. (2001)) Re” =5000 (Jovic & Driver (1995)) Re” =6800 (Jovic & Driver (1995)) DxV+DO>O 91 (b) Rey > 10000 ‘60 0:5 1 “1.5 i if 2 2.5 x/xr ReH =13000 ReH =10000 (Spazzini et al. (2001)) Re" =16000 (Spazzini et al. (2001)) Re” =11900 (T ihon et al. (2001)) Re” =10400 (Jovic & Driver (1995)) XVDOO Figure 4.7 Comparison of the mean skin friction coefficient results with other studies: (a) ReH< 10000, (b) ReH> 10000 From x/x, z 0.33 — 1.0; i.e., within the primary recirculation region, the mean Cf is negative and the maximum "reverse" Cf is at x/x, z 0.68. Downstream of the reattachment, the mean Cf rises slowly to a steady value at x/x, z 2. The magnitude of mean Cf in Tihon et al. (2001) is generally higher than all other measurements. Hence, these data are not used for comparisons concerning C; magnitude. The data at R811 =13000 agree well with those at Re” = 10400 (Jovic & Driver (1995)), and Re” =10000 92 (Spazzini et al. (2001)) except for the region around the maximum negative shear, where the present data are closer to those from Spazzini et al. (2001) than the others. Similarly, the data at Rey = 8700 collapse on those at Re” = 5100 (Spazzini et al. (2001)) and Re” =6800 (Jovic & Driver (1995)) except for the region around the maximum shear reversal (x/x, = 0.68). The magnitude of Cf from the current data is somewhat smaller than that obtained from the other studies around x/x, = 0.68. Finally, at the lowest Reynolds number of Re” = 4300, the results are consistent with Rey = 3500 data (Spazzini et al. (2001)) upstream of reattachment. However, farther downstream, the data at Rey = 4300 are obviously larger than those at Re” = 3500 (Spazzini et al. (2001)). The discrepancy might be caused by the differences in the geometry and/or the separating boundary layer state. In Spazzini et al. (2001), the step has planar geometry and the flow upstream of the point of separation is a two-dimensional fully-developed channel flow. Figure 4.8 depicts the magnitude of the minimum skin-fiiction coefficient (Cflm), which is the maximum skin-friction coefficient of the backflow, against the Reynolds number. More data points from J ovic and Driver (1996) are included in the figure for the purpose of comparison. All data points collapse fairly well onto a straight line with —1/2 slope when plotted using logarithmic scale. This kind of power-law relationship between CM”. and Re” was also found in Devenport and Sutton (1991), Jovic and Driver (1996) and Tihon et al. (2001). Tihon et al. (2001) included data fi'om Jovic and Driver (1996), Westphal et al. (1987) and Devenport and Sutton (1991). They concluded that the flow near the wall in the primary recirculation region is viscous-dominated or laminar-like, where the local skin-friction coefficient is inversely proportional to square root of the Reynolds number. 93 l 5.103»- I 2‘." ”l 0 present l l A Jovic & Driver (1996) . — -1/2 slope 10“ 2 2 . . 2 2 . 2 2 2 2 . . . . 103 104 105 ReH Figure 4.8 Minimum skin-friction coefficient versus Re” Despite differences in various experiments, it is of interest to note that a universal empirical correlation between Cm». and the Reynolds number provides a fairly good agreement with all data sets. Tihon et al. (2001) investigated the skin-friction coefficient inside a water channel using a two-dimensional BFS geometry, J ovic and Driver (1996) studied the air flow over a two-dimensional BFS geometry while the current study is concerned with the air flow over an axi-symmetric backward-facing step. Figure 4.9 shows a comparison with published results for the forward flow probability. All of the measurements show FFP value that is close to 100% just downstream of the step. This value decreases rapidly to 50% at x/x, z 0.33, which marks the boundary between the comer eddy and primary recirculation region. Downstream of this position, the FFP decreases to 0% followed by an increase to 50% again at x/x, z 1, or the reattachment position. The FFP continues to rise until it reaches 100% around x/x, z 1.33. The x/x, 94 location corresponding to 50% F F P at the point separating the primary and secondary recirculation zones, becomes smaller with increasing Reynolds number. This suggests that the size of the secondary recirculation zone becomes smaller relative to the primary recirculation zone as the Reynolds number increases. : 6 80 (3+ 8‘ 9 g Q g 60» v ‘6 8 s. E u. 40. 9; j 20' $27.- l t ‘ 05” 0 1 ” ”1.5 ‘ 2 25 Mr 0 Re” =4300 ‘ Rep-1 =8700 0 Re” =13000 o ReH =3500,Spazzini et al. (2001) A ReH =5100,Spazzini at al. (2001) v Re" =10000,Spazzini at al. (2001) 0 Re“ =16000,Spazzini et al. (2001) + Rel. =3860,Tihon et al. (2001) Figure 4.9 Comparison of forward flow probability results with other studies The observations from Figure 4.7 and Figure 4.9 are quite consistent. Positive values of the mean skin-friction coefficient correspond to FFP > 50% (i.e., most of the time the flow is in the downstream direction), while negative Cf is found at locations 95 where FFP < 50%. At FF P = 50%, sz 0: thus the consistent agreement found earlier in determining x, from F FF and Cf(see Table 4.2). Finally, it is interesting to note that the location of the maximum reverse shear (x/x, z 0.68) is different from the location of FFP z 0 (x/x, = 0.57). Figure 4.10(a) and (b) display the rms skin-friction coefficient from the current measurements and those from the selected studies. The trend of C;’ is the same for all measurements. It increases monotonically with increasing x, reaching a maximum in the vicinity of the mean reattachment point, followed by a slow decay thereafter. (a) Re” < 10000 x 10‘3 5 , + . - 4 1 +++ l l + 1 4 ++ 1 ++ + 3I + +++ ~ ... I + o +++++ l 1 ‘ ‘ 8 O I + “ ‘AAAAA A I l O 2 2 2 l 0 0 5 1 1.5 2 2 5 x/x , 0 Re“ =4300 A ReH =8700 0 Re... =3500,Spazzini et al. (2001) A ReH =5100,Spazzini et al. (2001) + ReH =3860,Tihon et al. (2001) 96 (b) Re” > 10000 -3 5 x_1_0__ f *7 4. 3 I. {r Re" =13000 ReH =10000,Spazzini et al. (2001) Re“ =16000,Spazzini et al. (2001) Re” =37300, Westphal et al. (1981) Re“ =36000, Adam & Johnston (1988) XUDOO Figure 4.10 Comparison of rms skin-friction coefficient results with other studies: (a) ReH< 10000, (b) Rey> 10000 Qualitatively speaking, the Cf’ results bear similarity to that associated with the rms wall-pressure fluctuation in Lee and Sung (2001) and Hudy (2001), where the rms pressure rises with increasing downstream distance till x/x, z 1 and continues to decay afierwards. This suggests that the driving mechanism of wall-shear fluctuations is linked to that of wall-pressure fluctuations. Such a link would be consistent with the conclusions of Devenport and Sutton (1991) that the fluctuations of the streamwise velocity near the 97 wall were driven by pressure gradient fluctuations imposed by the separated shear layer. One component of this pressure gradient is that associated with the low-frequency flapping of the shear layer. The other is that generated by the momentary passage of the vortical structures above the point of 1;. observation. The detailed characteristics of the shear-layer vortices were examined recently by Kostas et al. (2002) using PIV measurements within the separated flow region. They also proposed that small-scale vortical structures were responsible for Reynolds stress and turbulent kinetic energy upstream of reattachment while large-scale structures were dominant downstream of reattachment. Moreover, Scarano et al. (1999) found that the vortex formation started from x/x, z 0.17, and that the vortices approached the bottom wall at x/x, z 0.58 in their PIV investigation of a BFS flow. Owing to their smaller size and larger distance above the wall, the smaller vortices found near separation presumably have less of an effect on the wall-shear than the larger ones, farther downstream. The gradual increase in C;' with x/x, is, thus, partly a manifestation of the increase in energy of the passing vortices and their movement closer to the wall. Near the step, the very small skin-friction fluctuation could be caused by the unsteadiness of the corner eddy, which is associated with the flapping of the shear layer. Downstream of the reattachment, Cf’ decays slowly due to the lack of free shear layer activity there. The energized vortical structures from the separated shear layer decay and diffuse as they travel downstream. Comparing the magnitude of Cf' from the vaious investigations,, the data from Tihon et al. (2001) are higher than others, as was the case with the mean Cf results (see Figure 4.10). Although the data from Westphal et al. ( 1981) and Adam and Johnston 98 (1988) are close to those from Spazzini et al. (2001), they were obtained at much higher Reynolds number. Except near x,, the current rms skin-friction coefficient values at Rey = 4300 fall between the data at Re” =3500 and 5100 from Spazzini et al. (2001). At Re” = 8700 and 13000, the current data are smaller than those at Re” =10000 and Re” =16000, respectively. However, the reader is reminded of the axi-symmetric geometry of the back step used here, which may be the cause of some of the observed differences. Additionally, with the exception of the data from Tihon et al. (2001), all other investigations employed one or more variation of thermal-wake tagging by a heated wire to detect the flow direction. The present investigation is the first to employ flow direction detection that is not based on thermal-wake detection. It is also interesting to examine the location of the peak rms skin-fiiction coefficient. The data from Westphal et al. (1981) and Spazzini et al. (2001) peak at x/x, z 0.72, Tihon et al. (2001) at x/x, z 0.76, Adam and Johnston (1988) at x/x, 211.25, while the current results exhibit the peak at x/x, = 1. It is well documented that the peak rms wall pressure occurs slightly upstream of the reattachment point in two-dimensional BF S flow and splitter-plate flow. Driver er al. (1987) found the peak rms wall-pressure at x/xr z 0.91, Heenan and Morrsion (1998) at 0.91, Lee and Sung (2001) at 0.93, and Hudy et al. (2003) at 0.92. Due to the limited number of measurement locations, there is no sensor data available from x/x, = 0.9 — 1.0 in the current study. It is possible that the maximum rms skin-fiiction coefficient found here could be located at x/x, = 0.9 —1.0, which would be consistent with the wall-pressure measurements. 99 4.2.3 Probability Density Function (pdf) The normalized pdfs of the fluctuating wall-shear stress are shown in Figure 4.11 (a), (b) and (c) for all Reynolds numbers. Consistent with the mean/rms skin friction coefficient, at x/x, = 0.07 and Re” = 4300, the values are very close to zero. Deviations from the mean are very small, signifying a very low level of fluctuations. The shape of the distribution is generally similar to those at the higher Reynolds numbers. At x/x, = 0.6, which is close to the position of maximum wall shear in the reverse direction, most of the data have negative values and the distribution spreads over a wider range, corresponding to large fluctuation level. (a) Re” = 4300 3000 . 400 A XIxr=0'07 j I 300 2000» '3 200 1000i 100 j- o 2_ 2 2 _ ___22 2l -0.01 O 0.01 -0 01 300 2 400 300» 200 100 , 0 222 2 -0.01 100 (b) Re” = 8700 101 0.01 (c) Re” =13000 4000 2. 1000 -— 2 2 — ~—— — x/xr=0.06 ' I x/xr=0.54 3000 1 - _ “g 2000 I 500I I 1000 l . . ___l ol I I 49.01 0 0.01 -0.01 0 0 01 1000 r—---—---— 1 800 .. . I VX5032 x/xr=1.54 ' I 600 I R 5°°I I 400 I I . I 200.; j , 49.01 0 0.01 49.01 0 A 0.01 of of Figure 4.11 Wall-shear pdf at different x positions: (a) ReH= 4300, (b) ReH= 8700, (c) Re” = 13000 At locations near x/x, =1, i.e., the mean reattachment location, the histogram exhibits a bi-modal character. Such a histogram is characteristic of quasi-periodic, or limit-cycle, type signals. The “bi-modality” is related to the movement of the instantaneous point of reattachment up/downstream of x z x,, and may be indicative that such movement is the result of a limit cycle resulting from global instability of the flow. Hudy et al. (2003) suggested that the flapping of the shear layer, which would cause the reattachment point movement, is driven by an absolute instability of the primary re- circulation bubble, leading to self-sustained, or limit-cycle, expansion/contraction of the 102 separation bubble. The bi-modal character of the wall-shear histogram is in agreement with this scenario. Finally, downstream of the mean reattachment location, x/x, = 1.54- 1.75, most of the measurements are positive, which is in agreement with the fully attached flow well beyond x,, as found earlier from the FFP results. 4.2.4 Power Spectra The power spectra of the fluctuating wall-shear stress at different locations are plotted in Figure 4.12 (a), (b) and (c) for Reynolds numbers of 4300, 8700, and 13000, respectively. The original time series was divided into small records (4096 points per record). The spectra (Pu) were obtained by averaging the spectra of these records as follows: M5 X(f)*C0nj(X(f)) 1 (4.3) mez Pxx(f) = " where f is frequency, m is number of record, N is number of points in each record, and X0) is Fourier transformation of data record x(t). To avoid clutter of the plot, each spectrum was shifted up by one decade relative to that one x location upstream of it. The spectra and fiequency are normalized as follows: 4- Pxx Pxx = ___—1 2 (4.4) IEPUiI :1- f-x, = 4.5 f U0, ( ) 103 Two primary observations may be made from Figure 4.12. First, for all Reynolds numbers, the spectrum value decays with increasing frequency at all x locations. This is fundamentally different fi'om wall-pressure spectra, which exhibit a spectrum "peak" switch with increasing x. To demonstrate, wall-pressure spectra measured by Hudy et al. (2003) on the same model and Re” = 8700 are shown in Figure 4.13. Unlike rw' spectra, the peak of the wall-pressure spectrum switches from being within the low-fiequency range near separation to a frequency of f z 0.5 for x/x, > 0.9. This discrepancy will be discussed in the following chapter (section 5.3) when presenting a unified picture that accounts for both one- and two-point observations. Although no 239' spectrum peak is found around f z 0.5 in Figure 4.12, there is a clear "filling" of the spectrum in the vicinity of this frequency as x increases. This may be identified by the decrease in the spectrum slope in the middle range of frequencies as identified in the figure. For x approximately less than 0.3x,, a slope of —4/3 is found. At larger x locations, the slope becomes —2/3. Also at locations sufficiently far downstream, x/x, > 0.8, the spectral slope at the high-frequency end exhibits a -5/3 slope. This is indicative of the establishment of an inertial subrange, presumably associated with the re- development of the boundary layer after reattachment. The rise in the spectrum energy in the middle range of frequencies as x approaches xr (around f z 0.5) may be seen more clearly by plotting the spectra on a semi-log plot, as seen in Figure 4.14. For these plots, the horizontal axis is log(f"'), and the vertical axis is linear and represents the product of P; x f. In this fashion, the geometrical area under the spectra represents the energy level. For example, the area under the spectra around x/xr = 0.07 is smallest, which corresponds to very small energy 104 - i.e., very small fluctuation of a. On the contrary, the area under the spectra around the reattachment position is largest, where the fluctuation is highest. (a) Re” = 4300 -6 10 222 .22 222222222 1o“3 ’- -10 10 42 .O_§ 1o 2 -14 10 ~16 10 -1a 10 9.2... 2. 222 ......._k_.22k2._ ___22 2. ___“- hux-._~._2 2;. _A2 A*L.C~A b—b ___—___.L .a.__._4._¢ A_._r 10'2 (b) Re” = 8700 .5 10 10" - -10 10 .12 .032 10 2 .— x/x, = 0.57 _. <— x/x, = 0.31 10 10' (c) Re” =13000 10.6 IF— ‘—'_'—'“”“*_'fi afi h} 10'8 } -------------- M ‘ --------- : .213 1o“°~ \ <-x/x,= 1.67 . . - (3.3—Vb: 1.17 § 1042' 'm " \ <—x/x,=0.79 “ 10-14” ““““““““ «X/xr=0.54 A <—x/x,=0.29 i 10'“? .: 18 <-x/x,=0.06 j 10 . ....l W W: . ALA 10'2 10'1 10° 101 102 f. Figure 4.12 Wall-shear power spectra at different x locations: (a) ReH= 4300, (b) Re”: 8700, (c) Rey = 13000 (DPP 109 - ”Rs. 4— x/x,z 2.3 106 . 4— x/x,z 1.8 103 - "2.x {— x/xrz 1.3 1 """""""""""""""""" <— x/x,z 0.9 00 _ -.""-\..~.‘ ‘ 1 p - 4— x/x,z 0.6 'f- a l... . __fig/M; 10'2 10’l 10° 101 102 Figure 4.13 Power spectra of the fluctuating wall pressure at different x locations (Hudy (2003), private communication) 106 (a) Re” = 4300 “'1 E I 4 I 5x 10“3 { , ‘ . — x/xr=0.07 , .-°-'°': --- x/xr=0.33 r 4' '° -.-. x/xr=0.61 :: ...... xlxr=0.90 33» . a: ‘ ' Q 2f r.,\—l°V-‘ k; I. 3'. :_ I “. °-. .1 9‘. 3 1% I" ‘3. i \3. ,. oL—d ""' ""‘ 10'2 10'1 10° 101 (b) Re” = 8700 FT — xlx =0.07 {:3 -..- xlx =o.31 -.-. x/x =o.57 ...... x/xr=0.83 q n-I . C ,,«- $5.: 1 .n 3 °. . I"! ‘ j ... .’ 0.. I .. .’o ~"-\’AA”‘ "‘ v ‘ "\ ‘ :— T "“ “w_-__ _ -1 0 1 10 10 10 107 (c) Rey =13000 .9 2.5§_1_0__,_,_,,,,,, , - ,t._-s-__w,-._- , V! _ — xlxr=0.06 i :,‘: --.. xlxr=0.29 r 2 gigs-z, -.-. x/xr=0.54 J .5‘ 3"- ...... x/xr=0.79 " 15‘ t3 ‘ 'u— . i s a. X 1 .’ 5' \ t o I” F” W ‘ #702 Figure 4.14 Wall-shear power spectra plotted on semil-log scale for different x locations: (a) ReH= 4300, (b) ReH= 8700, (0) Re” = 13000 At x/x, z 0.07, most contribution to the total energy of the fluctuating components of the wall-shear stress comes from low frequencies for all three Reynolds numbers. This may be seen more clearly from Figure 4.15, where a different vertical scale is used to magnify the view of the spectra at x/x, = 0.07. Ignoring the harmonic peak at f z 1 and the lowest Reynolds number, which is believed to be associated with noise, the energy content is concentrated around a frequency of f z 0.1. A peak at f z 0.16 was found in the pressure measurements of Hudy et al. (2003) upstream of O.25x,. Spazzini et al. (2001) also detected a peak at f z 0.1 in the wall-shear spectra measured upstream of O.28x,. For the locations downstream of x/x, z 0.30, the build-up of energy up to the reattachment location clearly falls in a frequency range that is centered around an f value 0.65 (see Figure 4.14). This frequency range encompasses the passage frequency of the 108 shear layer vortices, as reported in the literature (see next paragraph for specific details). Thus, it is apparent that the increased level of wall-shear fluctuations is a manifestation of the increased energy and proximity of these vortices as they approach the reattachment point. The two frequencies of f = 0.1 and 0.65 were also identified in the limited wall- shear literature available. For example, Spazzini et al. (2001) found the low and high ' frequencies to be f = 0.08 and 1.0, respectively. More abundant observations of these frequencies are available from wall-pressure measurements. Peak frequencies of f = 0.12 — 0.18 and 0.6 - 0.9, were found by Hudy et al. (2003) in a fence/splitter-plate model, f' = 0.11 and 0.50, by Lee and Sung (2001) in BPS flow, and f = 0.18 and 0.6, by Driver et al. (1987) in BFS flow. —e— YRef:4300fi 1‘ —A— Re=8700 —+— Re=13000 4' xf 3r .0}? 10'2 10'1 10° 101 102 f' Figure 4.15 Wall-shear power spectra plotted on semi-log scale at x/xr z 0.07 109 4.2.5 Temporal Autocorrelation Function The autocorrelation coefficient (R (1’) obtained from single-point measurements is generally used to identify the dominant time scales of the signal of interest. R H is essentially the inverse Fourier Transform of the power spectrum. In the summary, Rf,» may be calculated using the following equation: 1ti.(t)rl.(t - 9) (4.6) Rr‘r'(9) z , (t)2 1w where the overbar denotes time averaging, Sis the time delay, and tw'(t) is the fluctuating wall-shear time series. Figure 4.16 (a), (b) and (c) display RH for three Reynolds numbers at x/x, z 0.07, 0.7, 1.3 and 1.8. These positions correspond to those immediately after the step, in the vicinity of x, and farther downstream of x,, respectively. The general trends in the data are the same for all three Reynolds numbers. As expected, R” equals 1 at 9 = 0 and decreases monotonically with the increase in the time delay. At x/x, z 0.07, Rf’f has a long tail and approaches a zero value at SUw/xm i 15 for Re” = 4300 and i 8 for Ken = 8700 and 13000 (see Figure 4.17), while it decreases much more quickly and drops to zero at SUm/x, z i 2 at the rest of the locations. The longer correlation time means that larger scale (or lower frequency) motion dominates the fluctuation of the wall shear. Therefore, it may be concluded that low-frequency disturbances are dominant immediately after the step and higher-frequency ones are more dominant farther downstream. This result is consistent with the power spectrum analysis in the previous section, where it was found that most of the energy was contained at the low-frequency end of the spectrum at x/x, z 0.07, while most of the contribution to the total energy came fi'om higher frequencies downstream of x/x, z 0.3. 110 (a) Re” = 4300 Rfr’ -37 1577-4732771) '2' it, is ’8 SUOJx, (b) ReH=8700 1'7 7 #Ti #5 7'7 7 7"— Rff ‘ — X/Xr=0 07 ...j. x/xr=0.70 03} -.-. $0421.36 " 1 x/xr=§1.88 ‘ 0.6) ~ 1 0.4) i 0.2) 111 (c) ReH= 13000 Rr’r’ -8 -6 41’ -i2 707772 4 6" 3 SUa/x, Figure 4.16 Autocorrelation coefficient of the fluctuating wall-shear stress at different x locations: (a) ReH= 4300, (b) Re”: 8700, (c) Re” = 13000 1: 7e 7‘ #¥ n.7, a 7,, 7 a Rh: — ReH=4326 , l ...... ReH=8653 08" " ' " ' ' ;.‘.‘.'ReHi12979 ‘ 46‘ ‘7 7 :11)” ‘7 "_o—TV ' To" W 2‘0 SUq/x, Figure 4.17 Autocorrelation coefficient of the fluctuating wall-shear stress at x/xr z0.07 112 Contour plots of the autocorrelation coefficient at all x locations are displayed in Figure 4.18 (a) — (c) to show a filler picture of the variation of Rf,» in space. The horizontal axis represents the streamwise distance from the step normalized with x,, and the vertical axis gives the non-dimensionalized time delay 9U../x,_ The color indicates values of RH. The contour line with Rh: value of 0.1 (labeled as R0,; in Figure 4.18) is used arbitrarily here to mark the width, or integral scale, of the autocorrelation. At the most upstream location, the width of R” extends approximately from SUw/x, = -5 to 5. As x increases, the correlation width decreases fairly rapidly till x/x, z 0.5. Farther downstream, the correlation remains practically unchanged. These variations in the integral time scale of the wall-shear stress are consistent with those fiom the wall- pressure measurements by Hudy et al. (2003). These authors found that the fluctuating wall-pressure autocorrelation was wide at x < 0.25x,., but it narrowed-down quickly over the range x z 0.25 - 0.5x,, remaining relatively unchanged downstream of that position. Unlike the wall-pressure autocorrelation fimction, R” does not posses a region with appreciable negative correlation. The wall-shear autocorrelation approaches zero with increasing time delay, oscillating with very small amplitude around zero. In contrast, the wall-pressure autocorrelation of Hudy et al. (2003) crosses zero, reaching a negative peak of -0.1 to -0.2, before oscillating around zero. The negative correlation peak is inferred to be a manifestation of the quasi-periodic nature of the separated shear-layer vortical structures, which dominate the wall-pressure signature in the vicinity of reattachment. Although these vortices are also expected to have a significant impact on the wall-shear signature as well, it is believed that the influence of viscous damping on the near wall flow attenuates this signature; hence, obscuring the negative correlation 113 peak. A more detailed comparison of wall-shear and wall-pressure results is given in section 5.3. (a) Re” = 4300 10 .1 SUw/x, 08 R1411 5 -, ,:0.6 0 .0.4 _5 fl ‘ 0.2 0 —10 x/x, (b) ReH= 8700 10 .1 SUa/x, - 0 8 Rev 5 20.6 0 0.4 -lO x/x, 114 (c) Re” = 13000 10 S Ugo/x, x/x, Figure 4.18 Contour maps of the auto-correlation coefficient of the fluctuating wall- shear stress: (a) Re”: 4300, (b) Rey: 8700, (c) ReH = 13000 115 CHAPTER 5 TWO-POINT MEASUREMENT RESULTS In this chapter, analysis of the space-time characteristics of the wall-shear stress in the BF S flow is presented. Specifically, the results fi'om the two-point cross- correlation and coherence between the time-series obtained fiom the reference and movable OHW sensors are examined. Subsequently, convection velocities of the dominant flow structures are obtained through the use of two methods: one in the time domain (based on the cross correlation) and the other in the frequency domain (based on the phase of the cross spectrum). Since there is no literature involving two-point measurements of the wall shear in separating/reattaching flows, the current data are exclusively compared with those from wall-pressure measurements. 5.1 Two-point Cross-correlation 5.1.1 Two-point Cross-correlation with the Reference Sensor Located near x, The spatial correlation function is a useful tool for identifying the length scales of the turbulent structures. In the current study, two OHW sensors separated from each other in the streamwise direction were used to acquire the wall-shear stress simultaneously for the purpose of obtaining the two-point correlation function. In particular, the two-point cross-correlation coefficient (Rt. t. ) is defined as l 2 Rea, (9.xm;xo) = 7”’("°”)‘”2(x’""‘9) (5.1) l \lx'w1(x0,t)2 Jam", ,t)2 116 where 1“,; and m represent time series of the wall-shear stress obtained from the reference and movable sensors, respectively, and S is the time delay. Negative/positive 9 values correspond to advancing/delaying 1“,; relative to 1w] in calculation of R . . . 1"172 In the present study, one sensor (hereafter referred to as the reference sensor) was placed at a fixed position (xmf) and the other sensor (hereafter termed the movable sensor) was moved along the streamwise direction. Two cases were tested. One set xref at 5.07H, i.e., in the vicinity of the reattachment position. The corresponding xrq/x, is 1.18, 1.09 and 1.04 for Rey= 4300, 8700 and 13000, respectively. The other placed x"; at 2.28H, i.e., in the middle of the reattachment zone. The corresponding xm/x, = 0.47, 0.44 and 0.42, in order of increasing Reynolds number. For the former reference position, the movable sensor locations extended from the step to x/x, z 2.0. For the latter reference sensor position, the movable sensor was placed at locations extending from downstream of the reference sensor to x/x, z 1.0. Figure 5.1 shows the two-point cross-correlation coefficient for all Reynolds numbers and selected movable sensor positions with the reference sensor close to the reattachment point. In the figure, the abscissa represents the normalized time delay, the ordinate indicates the cross-correlation coefficient, and the different lines correspond to different movable sensor positions. The black, red and blue color lines correspond to movable sensor positions that are upstream of the reference sensor; that is, x/x, = 0.76, 0.90 and 1.04 in Figure 5.1 (a), x/x, = 0.70, 0.83 and 0.96 in Figure 5.1 (b), and x/x, = 0.67, 0.79 and 0.92 in Figure 5.1 (c). On the contrary, the cyan, yellow and green color lines are used to represent the cross-correlation results for movable sensor positions 117 downstream of the reference sensor. A positive time offset means that the feature giving rise to the correlation between the two sensors registers at the movable sensor earlier in time before registering at the reference sensor, and vice versa. The largest cross-correlation coefficient is found at the closest spacing between the two sensors and decreases with the increasing spacing. Upstream of the reference sensor, the main peak of Rt. 1:. occurs at smaller time lead with decreasing distance 1 2 between the reference and movable sensors. Similarly, downstream of the reference sensor, the main peak of RT.1" takes place at a time lag that increases with increasing 1 2 spacing between the two sensors. This is a manifestation of a downstream-traveling disturbance. The shape of the Rt- T. curves at the first two upstream positions is different from 1 2 the rest. In addition to the main peak of the cross-correlation, these curves exhibit a second peak of the same order of magnitude as the main peak but at negative time delay. These secondary peaks are marked by open circles in Figures 5.1 (a) - (c). It is interesting to note that the time delay corresponding to the secondary peak is consistent with a downstream-propagating disturbance at Re” = 4300, a stationary disturbance at Re” = 8700 and an upstream-propagating disturbance at Reg =13000. The significance of this secondary peak and the reasons why the convective characteristics of the disturbance it represents change with increasing Reynolds number are not known. 118 (a) ReH = 4300 (b) ReH = 8700 ’1': 119 (c) ReH = 13000 0.25177 3 . r}. ! —; x/xr'-f0.67 12 0.2'1 "is ”4:919- . -.L. x/xr:}o.92 . Q15 . f .-.-...z.X/Xr¥_1.29.h .."H' ‘3 x/xr=1.79 . 0.1 ' i f '_.; \‘ x/xr=9 09 0.05 0 -o.o5~ - j --------- — -o.1 SUa/x, Figure 5.1 Cross-correlation coefficient of the fluctuating wall-shear stress with the reference sensor located near x,: (a) Re”: 4300, (b) ReH= 8700, (c) Re” = 13000 Figure 5.2 (a) — (c) show contour plots of the two-point cross-correlation results at all x measurement locations. At the reference sensor position, the results are the same as the auto-correlation with a peak value of 1.0 at zero time delay. Note that the color scale used to represent the correlation value only spans the range — 0.1 through 0.5 to make the lower cross-correlation values visible. The most obvious feature of the correlation maps is inclined-lobe-shaped contours with negative slope. This pattern of contour lines is consistent with that produced by a downstream-propagating wall-shear disturbance. Furthermore, the slope of a locus of the correlation peaks along the center of the lobes (highlighted with broken line in the figures) is inversely proportional to the mean convection velocity of the disturbance. 120 (a) Re“ = 4300 SUm/x, qr, (b) ReH = 8700 9 U m/x, 7172 x/x, 121 (0) Reg = 13000 SUx/x, . . 0.4 x/x, Figure 5.2 Contour maps of the cross-correlation coefficient of the fluctuating wall- shear stress with the reference sensor located near x,: (a) Re”: 4300, (b) Re”: 8700, (c) Re” = 13000 At the lowest Reynolds number, the lobe-shaped region over which most of the correlation exists extends from about 0.7x, to beyond 2x, in the streamwise direction and from time delay of - 2SUw/x, to + ZSUm/xr. As the Reynolds number increases, the area of the lobe appears to become somewhat larger. This apparent, but small, increase in the lobe size is caused by the fact that the reference sensor position decreases fiom 1.18x, to 1.04x, as the Reynolds number increases from 4300 to 13000. As a result of this upstream movement in the non-dimensional location of the reference sensor, the entire correlation contour plot shifts in the upstream direction; thus extending to x/x, locations below that found at the lowest Reynolds number. On the other hand, the increase of the temporal extent of the lobe area reflects the slight increase in the width of the auto- 122 correlation function, or normalized integral time scale, in the vicinity of x/x, = 1 with increasing Reynolds number (see Figure 4.18). Another observation from the contour plots is that the extent of the lobe in the streamwise direction is not symmetric about the reference position. The correlation distance is shorter upstream than downstream of the reference position. No significant correlation is found upstream of 0.6 - 0.7x,. On the contrary, downstream of the reference position, the wall-shear fluctuations are still correlated with the reference signal as far as 2x" and possibly farther. It is not surprising to see this long extending correlation downstream of the reference point. It is well known that the coherent vortical structures are sustained over fairly long distance beyond the reattachment point. (see Kostas et al. (2002), Lee and Sung (2002)). However, the shorter correlation distance on the upstream side is somewhat different from the wall-pressure results, where, for example a correlation coefficient of 0.1 extends to x/x, z 0.5 and in addition, there is substantial negative correlation even further upstream, in comparison with x/x, z 0.9 in the wall- shear case. Because the shear layer vortices are initially small and not too energetic near the separation point, their influence on the wall pressure is generally not felt close to the step. However, as they travel downstream, the vortices grow in size, move closer to the wall, and become more energetic and begin to influence the wall-pressure signature downstream of x/x, z 0.25 (Hudy et al.(2003)). For reasons that will be discussed in section 5.3, it is believed that the wall-shear signature of the vortices is even weaker than the wall-pressure one. Hence, they would need to travel farther downstream before influencing the wall-shear stress: a suggestion that is consistent with the short correlation width upstream of the reference probe location. 123 One way to obtain the convection velocity of the wall-shear disturbances associated with the vortex passage is from the slope of the peak locus of the cross- correlation contour plot. Figure 5.3 contains a plot of the correlation-peak time delay at different locations of the movable sensor for all Reynolds numbers. The horizontal axis represents the streamwise position of the movable sensor and the vertical axis gives the corresponding time delay of the correlation peak. The symbols display the experimental data, which are used to arrive at a linear least-square fit to obtain the convection velocity from the slope of the fit. Note that this approach assumes that the convection velocity does not vary with x, since the slope of a straight line is the same at all x locations. It is noted, though, that the data do exhibit some deviation fi'om the straight line, suggesting that an x-dependent convection velocity may be more appropriate. The slope (ms) of the line may be related to the mean convection velocity as follows: flea; (5.2) U.no ms The resulting mean convection velocities are 047,045 and 0.46Uao in order of increasing Reynolds number. Compared with the wall-pressure measurements, the downstream convection velocity is 0.57U... in Hudy (2001), 0.6U,o in Lee and Sung (2001), 0.5 — 0.6U,o in Heenan and Morrison (1998), 0.5U0° in Cherry et al. (1984), and 0.5U,o in Kiya and Sasaki (1983). All of these values, which are obtained in various configurations (fence-with-splitter-plate, thick splitter plate and BFS), consistently fall in the range of 0.5 - 0.6Uw. In addition to the mean convection velocity (averaged over all scales of motion and x locations), the local convection velocity may be estimated fi'om the inverse of the 124 slope of the correlation peak locus (broken line in Figure 5.2). Figure 5.4 shows the local convection velocities for all positions. The observed values decrease as the vortical structures approach the reattachment point, then increase again. 1.5 -- M——-—r , , —~—— .. _Tm- -__.____ __ SUm/x, A 0 Re” = 4326 1 1 A ReH = 8653 j + ReH = 12979 . 0.5: _ fit 1 i 0.. -0.5L -1. l 4.5) 1 -2. _: '2'35 "i T" "'1‘? "—3 ' 2.5 Figure 5.3 Correlation-peak time delay at different x location and associated line fits The deceleration of the vortices as they approach the separation point may be clarified through examination of the mean streamwise pressure gradient. Figure 5.5 displays the local convection velocities (He) and mean-pressure coefficient (Cp) in the same plot. The ordinate scale on the left side represents Uc normalized by U00, and that on the right side indicates Q, values. In the range x/x, z 0.5 — 1.0, the recovery of the mean pressure results in the establishment of an adverse pressure gradient. It is believed that this pressure gradient results in the deceleration of the vortices as they approach the reattachment point. The stronger this pressure gradient is, the stronger the deceleration effect. Since x, for the current flow is shorter than that found in other studies of planar 125 BFS flows, the pressure recovery occurs over smaller distance, leading to stronger adverse pressure gradient. This apparently results in the lower convection velocity as mentioned earlier. 1. , L , . i____.. -_.__.____ _ _--. .. —— o ReH = 4326 ; Uc/U” A ReH = 8653 1 08.l ... ReH =12979 ,; 0.6 J +Al' Q +A 0 °" 91: o 0.4;L 400 1" o 1 ”also «mo 1 0.2-- Figure 5.4 “Local” convection velocity at different x positions The reason for the acceleration of the vortices downstream of reattachment (Figure 5.5) is not clear. However, the observation is consistent with that from the wall- pressure measurements of F arabee and Casarella (1986) which shows the convection velocity to continue to increase downstream of reattachment up to an asymptotic value of approximately 0.7 Ugo. This may be partly caused by the continued increase in the size of the vortices associated with increase in the thickness of the attached shear layer with increasing x. As a result, the vortex center moves farther away from the wall, where the local mean velocity is larger and the decelerating-influence of the image vortex becomes 126 smaller. The latter presumably also decreases monotonically with increasing x, since the energy of the vortices, and hence their ability to induce velocity in the Biot-Sevarat sense, continues to decrease as they travel downstream. Uc/ Um -0.2 Figure 5.5 Local convection velocity and pressure coefficient versus streamwise locations 5.1.2 Filtered Two-point Cross-correlation with the Reference Sensor Located near x, In section 4.2.3, the two frequencies of f z 0.1 and f' z 0.65 were detected in the power spectra of Tw'. These frequencies are associated with the global flapping motion of the shear layer and the passage of the vortical structures, respectively. As just found in section 5.1.1, the latter convects downstream at a mean velocity of 0.45 - 0.47Uw, It seems interesting to see how the cross-correlation shape changes if the high frequency component is filtered out. To this end, a low-pass filter with a cut-off frequency f = 0.25 127 was used to filter the wall-shear data from both OHW sensors. The cross-correlation of the filtered signals is displayed in Figure 5.6 (a) — (c). Compared to the results in Figure 5.1, the first observation from Figure 5.6 is that the correlation-coefficient curves are smoother and, at large time delays, fluctuate around the zero correlation value with larger amplitude and period. The second finding is that the values of the correlation coefficient are larger; that is, the signals are more coherent. The reason for the above observations could be attributed to the removal of the effect of the high-frequency (small-scale) eddies. Thus, the correlation is higher over longer distances because of the remaining large-scale structure. It is interesting to note that, albeit influences of the filtering to remove the vortex passage effects, the cross-correlation shape still reflects the dominance of downstream- moving disturbance, as clearly illustrated in the contour plots of Figure 5.7(a) — (c). The inclined lobe structure with negative slope is eminent in the plots. However, the contours are more organized and the size of the lobes is also larger than that in Figure 5.2, consistent with the removal of smaller-scale influences by filtering. As mentioned above, the primary motivation for the filtered-signal analysis is to check the nature of the global flapping motion with frequency f' z 0.1 by removing the higher-frequency influences of the energetic vortical structures. Kiya and Sasaki (1985) proposed a model for the mechanism of surface-velocity fluctuation: a traveling wave with downstream convection velocity for the large-scale vortical structures and a standing wave with a node near the middle of reattachment zone for the low-frequency flapping motion. This standing wave was also observed by Hudy et al (2003). However, the contour plots in Figure 5.7 are still dominated by the convection motion. Possible 128 reasoning for this is as follows: although the passage frequency of the dominant vortices is f z 0.6, a fairly broad range of frequencies is associated with the passage of vortices of different sizes. This is reflected in the fact that the energy in the spectral plots in Figure 4.14 in the vicinity of f z 0.65 is contained in a broad, rather than narrow (harmonic), peak. It appears from the results in this section that the influence of vortex passage extends down to frequencies that are below the cut-off frequency of the filter. Kiya and Sasaki (1983) also found that very large-scale vortices were shed with frequencies less than f' z 0.2. Therefore, the wall-shear signature of the flapping motion in the vicinity of x, might still be masked by that of the vortices even for f less than 0.25. (a) ReH = 4300 0-4 " *7 "fir—+1 Rm; 1 9",. § —1— x/xr=0.76 ‘ ~"‘;’; i, ‘ --- x/xr=0.90 0,3» ~ ,' i\ ‘- -. .xlxrt.04 ‘ ‘ :.’ i 3. \_ -... x/xr:1.46 ,- ,- -, 3 --- x/xri1.89 t)... .. x/xr=z.u.5 1 129 (b) R6“ = 8700 0.4 l I . 1 LT Rn}; I 1 S —- X/Xr=:o-70 - ' 1 .-, i --- x/x =‘o.83 : ; I \ I'\ : r 0.3‘ i I. I .‘ -l-I XIV -0.95 ‘ 3i -’ t '\ .. '- i! -‘ ." -.-- X/Xr-:1.36 O ql . ! I. ‘\ ‘. --- X/xr:'1.88 -“ I “'5‘; ‘.‘ “ .. x/xr==z.1z Ii 3 .A‘ ti‘ " I '.. I ‘ ' \ . l 1 «I 1".- U \ \ | - 0'... . " 1‘. '- ‘ 0.1 ’ .' ’ )' I I. \I 41‘! i/ ‘/-— \ .- I - '- \ . ... F”“‘°.: ----- ’l I' ’4’ ‘. '\ $¥K--:‘\. '- --.'>.—’ 3 ' ‘1’ /'\._ 0' --’°"".X§, 3&3?! 0.9;..— AAAAAA ...... w _ 32.... ' 7 —o 1* i 3 #1 I-8 -6 -4 -2 0 2 4 6 8 SUao/Xr (0) R6“ = 13000 0.4fi AF . Rriryz 1 — X/Xr:=0.67 l --é- x/xr==0.79 ; 0.3 -. . x’x“0.92 ‘ ---- X/Xr==1.29 --- x/xr==1.79 0.2% .. X/Xr==2.UZ 0.1 “E‘~ s. .- a 0.” -"'.I ";;:".*‘¢‘.:;. _ i ‘ i 0"18 -6 -4 -2 O 2 4 6 8 Figure 5.6 Cross-correlation coefficient of the low—pass filtered fluctuating wall-shear stress with reference sensor located near x,: (a) Re” = 4300, (b) ReH= 8700, (c) Re” = 13000 130 (a) Re“ = 4300 .0.5 SUa/x, a R .t ,‘ e 0.4 {"1 0.3 "II—“P 0.2 x/xr (b) Re“ = 8700 4 SUw/x, . . 2 -2 131 (c) Re“ = 13000 SUaz/x, rlrz 0.5 1 1.5 2 x/x, Figure 5.7 Contour maps of the cross-correlation coefficient of the low-pass filtered fluctuating wall—shear stress with reference sensor located near x,: (a) ReH= 4300, (b) Re” = 8700, (c) Re” = 13000 5.1.3 Two-point Cross-correlation with the Reference Sensor Located near 0.5x, Since the wall-shear stress at the previous reference position did not show appreciable cross-correlation with that from the movable sensor upstream of 0.7x,, a second reference-sensor position around 0.5x, was employed. Specifically, the new reference location was 0.47, 0.44 and 0.42x, in order of increasing Re”. The movable sensor was placed at five x locations extending from the reference position to the reattachment point. Figure 5.8 shows the cross-correlation coefficient between the two sensors for all five locations of the movable sensor (represented by the different-type lines, as depicted in the legend). The largest cross-correlation coefficient is about 0.1 when the movable 132 sensor is placed closest to the reference probe. This value is smaller than that obtained at the same sensor separation but with the reference probe located near the reattachment position. As shown in Figure 5.1, the latter is about 0.15 — 0.2 depending on Reynolds number. By increasing the sensor separation to the second smallest one (x = 0.76x, at Rey = 4300, etc.), the correlation-peak value drops substantially to 0.05. Additional increase in the sensor separation appears to have little effect on the magnitude of the correlation. The most prominent feature in the cross-correlation results in Figure 5.8 is a sinusoidal-like signature that shifis towards negative time delay with increasing x location. The signature, which is depicted for all Reynolds numbers, is evident for all but the most downstream position. The time scale of the disturbances responsible for the generation of this signature may be estimated as twice the time delay between the negative and positive peaks of the signature, which is ZUw/xr. The corresponding frequency is f' z 0.5, which is approximately the same as that of vortex passage. This suggests that the wall-shear influence of the shear layer vortices is detectable as early as near 0.5x, (the approximate location of the reference probe), which is farther upstream than implied by the correlation results with the reference probe located near reattachment. Finally, Figure 5.9 displays contour plots of the cross-correlation coefficient relative to the new reference location. The downstream convective character of the wall- shear signature is evident from the negative-inclined broken lines tracing the positive and negative correlation peaks of the sinusoidal-like signature. It is interesting to recall here that the negative correlation values were absent in the results based on the reference 133 probe position near x,. It is also worth noting that the negative correlation-peak value becomes progressively smaller with increasing Reynolds number. (a) Re“ = 4300 Rtir} ‘ ‘ 'f- X/Xr-':‘O.61 0.15: , ,, , - -- i ‘ --i- ”#0154 , : -.t. x/xr:0.90 0_1 .-..-.,X/XrT-f.1.04..a .. .. x/xr=1.18 0.05 (b) ReH = 8700 R,, :er (c) ReH = 13000 S U aa/x, Figure 5.8 Cross-correlation coefficient of the fluctuating wall-shear stress with the reference sensor located near 0.5x,: (a) ReH= 4300, (b) Re”: 8700, (0) Re” = 13000 (a) Re“ = 4300 (a) Re” = 4300 x/x, 135 (b) Re” = 8700 SUm/x, x/x, (0) Re“ = 13000 SUa/x, . . 0.4 R’l’z 0.5 0.6 . 0.7 0.8 0.9 1 x/x, Figure 5.9 Contour maps of the cross-correlation coefficient of the fluctuating wall- shear stress with reference sensor located near 0.5x,: (a) ReH= 4300, (b) ReH= 8700, (0) Re” = 13000 136 The difference in the correlation-function shape for the two different x choices of the reference probe location warrants closer examination. When the reference probe is located near x,, it has ahnost equal probability of being up or downstream of the instantaneous reattachment location, depending on the phase of the flapping cycle. This results in switching of the sign of the shear signature in a manner that is clearly not related to the vortex passage. Hence, it is not too surprising that this effect would cause “de-correlation” of the vortex-passage signature, which in turn smears out the negative correlation peak. This also would be in agreement with the disappearance of the sinusoidal-like signature in the cross-correlation in Figure 5.8 at the most downstream location, which is in the vicinity of x,. 5.2 Frequency Dependence of the Convection Velocities The cross-correlation function analyzed in section 5.1 is the result of the non- discriminant influence of all scales of motion. The cross-spectrum (Pt't' ), on the other 1 2 hand, may be used to identify the frequencies corresponding to the turbulent motions responsible for the correlation of the turbulent wall-shear stress at two different points. P. . is calculated as follows: Tl"2 P. . (10:2!)1 mm; (f) (5.3) TITZ where f is the frequency, X1. is the Fourier transform of the reference sensor signal, 1 and X :' is the conjugate of the Fourier transform of the movable sensor signal. P .1. (f) is generally a complex number composed of phase angle and magnitude. The T1 2 137 angle of Pro 1.2 (f) indicates the average phase difference between the two signals at 1 frequency f. On the other hand, the magnitude of Pr. 1.2 (f) normalized as follows 1 defines the coherence (1‘1. 1' ) between the two signals: 1 2 5162“" 11'; (f )= (5-4) 12 Ptir'l(f)xpt'21'2(f) where Pt- I- and P1. 1' are the auto power spectra of the wall-shear fluctuations obtained 1 l 2 2 from the reference and movable sensors, respectively. Figure 5.10 shows the coherence plots for all three Reynolds numbers and the reference sensor position near the mean reattachment point. The location of the movable sensor corresponding to each plot is specified in the legend. The selected positions, for which data are shown, are the same as those in the cross-correlation plots of Figure 5.1 and Figure 5.6. The first three positions are upstream of the reference position while the last three are on its downstream side. In Figure 5.10, the abscissa indicates the normalized frequency (f') and the ordinate gives the coherence. Consistent with the cross- correlation results, the smaller the separation distance between the two sensors, the larger the coherence. At the Re” = 4300 and the most upstream and downstream locations, the coherence is practically zero for all fi'equencies. At the remaining four positions, a distinguishable but small coherence value is depicted for frequencies below f z 1. This frequency range encompasses the wall-shear—stress influences of the energetic vortical structures (f' z 0.65) and the global flapping motion (/ z 0.1). At Re” = 8700, the 138 coherence results are qualitatively similar but the magnitude is a little larger than that at Re” = 4300 for the same separation distance. At Re” = 13000, the value of the coherence increases further. Even at the most up and downstream positions, some coherence is now observed at low frequencies. The increased coherence at these locations far away from the reference sensor suggests an increase in the level of the low-frequency (large-scale) fluctuations with increasing Reynolds number. (a) ReH = 4300 139 (b) Re“ = 8700 ’l‘: ,. 7 ..V- l 0.3L 0.2; 1O 10 140 (c) Re“ = 13000 r——————— —- "—w——— —— ——«———.‘ F. . . * '1': 03 - x/xr=o.67 *: 0.3 ‘ XIX =0.79 0.2% 4' 0-2? 0.1 ; 0.3L 0.23» 0.1} « . . \ Ol____ . r , ___) Ot__,, __,,__”H, __4,_ . 10’2 10 f f' Figure 5.10 Coherence of the fluctuating wall-shear stress with the reference sensor located near x,: (a) ReH= 4300, (b) ReH= 8700, (c) Rey = 13000 The coherence level of the wall-shear fluctuation is smaller than that of wall- pressure fluctuation for the same separation distance. Again, this is attributed to the effect of local small structures and viscous dissipation on the turbulent wall-shear stress. However, it is still clear from Figure 5.10 that the coherence level at f < l is distinguishable from that associated with the uncorrelated motion at high frequencies. Since the coherence at a given frequency is a measure of the degree of phase locking between two signals at that frequency, it seems possible to calculate the phase difference between the signals obtained from the two shear sensors with some reliability for normalized frequencies less than 1. 141 The phase information is of interest as it can be used to determine the convection velocity at specific frequencies. To understand this, consider a traveling wave equation in the form F ( x,t ) = AT sin( 27r( ft i kx)); where Aris the amplitude, f and k are frequency and wavenumber, respectively, and “ i ” sign indicates up- or downstream propagation. The wavenumber is related to the convection velocity by k = f/Uc and the phase difference (‘1’) of the traveling wave at two different x locations is given by 2nkAx (Ax being the difference between the x locations). Thus, the convection velocity Uc = 2anx/ 4) for a traveling wave, and if one is to plot 4) (which is obtainable from the cross spectrum) versus the separation distance between the movable and reference sensors, the slope of the phase line at any desired frequency could be used to determine the corresponding convection velocity. Figure 5.11(a) — (c) display plots of the phase versus x location of the movable probe for frequencies of f z 0.06, 0.1, 0.2, 0.4, 0.6, and 0.8 and all three Reynolds numbers. Different symbols represent the phase calculated from the cross-spectrum at the different frequencies. A least-squares line fit of 45 -vs-x data is obtained at each frequency and displayed using a solid line in the figure. If the slope of the line is m¢, the convection velocity is obtained from: Uc 24f (5.5) Note that the phase at the reference sensor location is zero and that phase jumps of more than 21: between successive data points have been removed using unwrapping algorithms. Also phase data are only shown for the x range extending from 0.8 to 1.8x, where coherence values distinguishable above the background noise are found. As 142 l‘i shown in Figure 5.11, the data exhibit less scatter at f z 0.06, 0.1 and 0.2 because of the higher level of coherence at low frequencies. The fitting line with positive slope indicates downstream convection of the wall-shear disturbances for all fi'equencies. In agreement with wall-pressure results, no upstream convection was found because the analysis is limited to x > O.8x,. The upstream convection at f z 0.12 was observed by Hudy et al. (2003) from wall-pressure measurements upstream of O.5x,. This was obtained fiom phase results relative to a sensor positioned immediately downstream of the separation point (x/x, = 0.023). (a) Re“ = 4300 8 ___ _n ,_ k ___fl* _ WW9 i o f‘=0.06 i A f-=o.12 { ‘ 6r + f‘=0.18 <> r=o.41 ’ , 42» D f‘=0.59 o i <1 f‘=0.82 , 4 — fit > ’ 9 2t 4’ O ”...—— ,/ *3 '2i 5,4 I z .4 143 (b) ReH = 87001 ¢(rad) 8T (c) Re” = 13000 8 Ah, ,. 7? izm 7_ ¢(rad) 0 f'=0.04 6 A f'=o.11 i + f‘=0.20 ‘ o f‘=0.40 ‘ 4. D f"=0.60 i <1 f*=0.80 2i — m i 0i l 4 <1 x/x, Figure 5.11 Streamwise dependence of the phase angle at selected frequencies: (a) Rey= 4300, (b) ReH= 8700, (c) Re” = 13000 144 The convection velocities calculated from Equation 5.5 are plotted versus the normalized frequency for all Reynolds numbers in Figure 5.12. The average values of the convection velocities are 0.47, 0.45 and 0.43 Ugo in order of increasing Reynolds number, which are consistent with the results obtained from the cross—correlation. The broken line in the figure represents an equation relating the convection velocities to the frequency fiom the frequency-wavenumber spectrum of the wall-pressure measurements by Hudy et al. (2003) on the same model, which is given by: I UC f (5.6) U00 —1.69f‘+o.15 1r“*"‘ , #fi*#~.,_:kha A ReH= 8700 08* 0 Re": . a: .. i=__f___ ‘ 1.69 f‘ + 0.15 0.6» Pi , 0.2 i ol — — 4 +— ——~——~I ‘ 0.2 0.4 0.62 0.8 i Figure 5.12 Convection velocities obtained from phase-angle analysis The error bar at each point is estimated using one standard deviation of the calculated convection velocities from Equation 5.5. This is obtained from the standard deviation associate with the estimate of the slope of the line fit to the data in Figure 5.3 145 caused by the scatter in the data around the line (see Doebelin (1990), page 57). Notwithstanding the data scatter around the broken line, the calculated convection velocities are in general agreement with the trend predicted from wall-pressure measurements. Additionally, it is seen that the convection velocity value increases with increasing fiequency. This suggests that the high frequency (small scale) structures travel at a faster speed. Also note that the results are independent of the Reynolds number. 5.3 Additional Discussion In this section, an attempt is made to systematically compare wall-shear with wall-pressure measurements, and to examine the mechanism(s) causing differences and similarities between them based on the one- and two-point results from the present study. In both wall-shear and wall-pressure (Hudy (2003)) spectra plots, there are two dominant frequencies: f' z 0.1 immediately after the step and f“ z 0.65 farther downstream. The former is associated with the flapping of the separated shear layer, and the latter corresponds to quasi-periodic passage of the shear-layer vortical structures. One of the features associated with the flapping motion is the displacement of the instantaneous reattachment position. This motion leads to variation of the dividing- streamline curvature; thus inducing the pressure fluctuations. Devenport & Sutton (1991) proposed that the near-wall streamwise velocity fluctuations are related to the streamwise gradient of such shear-layer-imposed pressure disturbances through a one-dimensional model. The model is based on uni-directional simplification of the x momentum equation. More specifically, for an infinitesimally small distance above the wall, the streamwise 146 velocity (u) is assumed to be the dominant component and its gradient along the y direction is assumed to be much greater than that along the x and 2 directions. The momentum equation can then be simplified to: 2 ea. 16mm. (5.7) Devenport and Sutton (1991) assumed a sinusoidal pressure-gradient fluctuation of the fo ° 1 6p = A em" , with frequency 0) and amplitude Ap. For such a harmonic disturbance, Equation 5.7 produced the following solution of the instantaneous velocity field: A __2’. u = —p sin(wt) —e 50 sin[a)t __y_] (5.8) 0) 60 where 60 = "-21 is the oscillating boundary layer thickness. Devenport and Sutton (0 (1991) compared the predicted near-wall turbulence-intensity profiles using Equation 5.8 to experimental data in a sudden expansion in a pipe to verify the reasonableness of the a [—7 ”2 one-dimensional model. As shown in Figure 5.13, when 8+ (6+ = —0[ r; /p] ) is 8, v corresponding to f = 0.18 near the step and f = 0.74 near the reattachment point, the predictions are in good agreement with the measurements. It is noticed that these two frequencies are close to the two dominant frequency values identified in the current study. 147 J:- I’le IVVV'IVVI'YrerVIII‘IIVI'vIII'VIII ._a TDATUM FLOW ‘ FLOW WITH CENTERBODY ‘ ‘ D x/xr=.16 <> x/xr=.96 ‘ ‘7 x/xr=.181 o x/xr=.77 i x x/x,=.32 o x/xr=1.12 l - x/x,=.362 a x/x,=.9o71 G x/xr= .498 A x/xr= 1,043‘ A x/xr — .48 x x/xr = 1.28 + x/xr = .64 El x/xr = 1.6 , o X/xr = .634 V x/xr = .8 A x/xr = 1.95 Figure 5.13 Comparison of near-wall turbulence—intensity profiles between measurement results and one-dimension model prediction (Devenport and Sutton (1991)) Since the one-dimensional model appears to be reasonable, Equation 5.8 was used to predict the corresponding wall-shear stress associated with a pure harmonic pressure- gradient disturbance as follows: 611 Ap . 7! w t = u— = —._ an — 5.9 r () 6y y=0 0 sm( + 4) ( ) 148 A The corresponding ms of the wall-shear stress is p , which indicates that the wall- 420a) shear fluctuation is directly proportional to the amplitude, while being inversely proportional to the square root of the frequency of the wall-pressure—gradient fluctuations. Thus, for equal-magnitude pressure-gradient disturbances, the induced shear-stress fluctuations would be attenuated more at higher frequencies. This may explain the qualitative difference between the wall-pressure and wall-shear spectra observed earlier in section 4.2.4. The difference in the spectral shape is more evident if both spectra are plotted on the same graph. Figure 5.14 depicts such a graph, showing the power spectra of the wall shear and pressure for two similar positions at Re”: 13000. Each spectrum is normalized by its own peak value to fit the two spectra on the same graph. It is clear that the shape of the wall-shear spectrum is different from that of the wall-pressure one. The wall-shear spectrum decays with increasing frequency, while that of the wall-pressure exhibits a broad peak at f z 0.4. It is suggested here that a similar peak is not found in the wall- shear spectrum due to the inverse-square-root attenuation resulting from the Davenport and Sutton model, which results in damping of higher frequency 1w fluctuations relative to lower fi'equency ones. Notwithstanding the attenuation of high-frequency 7w fluctuations, the shear-layer vortices appear to be responsible for most of the 2"“, activity. One evidence of that is the two-point correlation results in section 5.1, which demonstrated the downstream convective character of the wall-shear fluctuations. Such a convective field is clearly associated with the vortical structures and not the shear layer flapping. Furthermore, 149 even though a low-pass filter with cut-off frequency of f = 0.25 was used to remove the influence of the passing vortices on the cross-correlation, the results still reflected the eminence of downstream-traveling disturbances. The power spectra analysis in section 4.2.4 also demonstrated that the amplification of the wall-shear rms value with increasing x primarily corresponded to a rise in the spectral energy within the frequency band associated with vortex passage. To underscore this observation further, a high-pass filter with cut-off frequency of f = 0.25 was used to remove the influences of the low- fi'equency flapping motion on the wall-shear stress signal. As discussed above, the energy within the pass band of the filter does not account for all contribution due to vortex passage since some of this contribution comes from frequencies below f of 0.25. Thus, the rms calculation from the high-pass—filtered signal is likely to underestimate the total contribution of vortex passage to the wall-shear fluctuations. — X/Xr=1.67 (tw') ...... XIXr=1.80 (pw') i J" l Normalize power spectra A o 10 A 10° 10 Figure 5.14 Comparison between wall shear and pressure power spectra 150 Figure 5.15 displays a comparison between the filtered and unfiltered rms values of the fluctuating skin-friction coefficient, i.e.; C;' and thf, respectively, while Figure 5.16 plots the percentage of the high-frequency contribution to the total z’W energy. The percentage increases continuously till x/x, = 0.54 and then remains practically constant with a value of 65 - 70%. This confirms that the r“, disturbances associated with the vortices are dominant downstream of x/xr = 0.54, which is the streamwise domain in which the cross-correlation measurements were conducted. x10- lr—'—-- e-r — --‘-T——-———— r -— C . —e— Yunfilterd l f + filtered 0.8- 0.6 0.4; Figure 5.15 Comparsion of rms skin-fiiction coefficient from the unfiltered and high- pass-filtered signals 151 80 S : E7 60: s3: 1 4oi r; 20 « T‘— 1 0L__ _ _ _ ___—L 1 ._ ._.._.;.__ ___J _ ___! 0 0.5 1 1.5 2 2.5 x/x, Figure 5.16 Percent z"w energy associated with vortex passage L” 152 CHAPTER 6 CONCLUSIONS AND RECOMMENDATIONS 6.1 Conclusions I. Development of a new wall-shear sensor A high-frequency oscillating-hot-wire (OHW) sensing technique for measurement of the flow-velocity/wall-shear-stress magnitude and direction was developed for the first time in the current work. The mechanism for direction determination by the sensor is based on a 180-degrees phase jump of the modulation signal, resulting from wire oscillation, when the flow switches direction. On the other hand, the shear-stress magnitude could be obtained in a manner similar to a conventional hot wire, after filtering the output signal to remove the modulated portion. Magnitude and direction calibration of the OHW sensor was conducted in a Couette-flow facility that was especially designed and constructed for the purposes of the present study. The facility provided a wall-shear calibration range of 0 — 0.7 Pa. In this range, a good agreement was found between the magnitude-calibration results with/without oscillation of the wire. This verified that the high-frequency oscillation of the wire did no result in an anomalous or non-linear behavior that may affect the wire’s calibration or its accuracy in retrieving the shear-stress magnitude. It was also found that decreasing the flow velocity (shear stress) resulted in amplitude reduction of the sensor’s modulation signal at the oscillation fi'equency. This reduction was associated with increased phase lag between the imposed and measured oscillation velocity of the OHW sensor. Both amplitude and phase effects did not 153 influence the measured shear-stress magnitude, the determination of which was based on _ the filtered, modulation-free, output of the sensor. On the other hand, the phase lag resulted in deviation of the direction response of the wire from the ideal one outlined above. However, since this phase delay did not exceed 90 degrees, it did not cause the modulation signal phase values to overlap for opposite shear directions. Thus, the flow direction could still be determined without ambiguity, afier appropriate direction- response calibration in the Couette-flow facility to determine the range of modulation- signal phase corresponding to each flow direction. 11. Reattachment length in the axi-symmetric backward-facing-stepflow The wall-shear signature behind an axi-symmetric BFS flow was investigated using the new OHW sensors while oscillating at a frequency of 2.8 kHz. Measurements of the wall-shear stress were conducted at different streamwise locations downstream of the step for three Reynolds numbers (based on step height) of 4300, 8700 and 13000. The reattachment length was determined using two methods: one based on the mean skin- friction coefficient and the other on forward flow probability. The difference between the results fi'om the two methods was less than 1.5%. The specific x, values varied from 4.3H to 4.86H with increasing Reynolds number. These values are smaller than those reported in the literature for planar BFS flows. It is hypothesized that the reduction in the reattachment length is caused by enhanced entrainment of the separated shear-layer in the axisymmetric case and/or transverse curvature of the test model used here (i.e., 6/R ratio effect; 5 is the boundary layer thickness at separation and R is the model radius). 154 HI. Mean shear-stress characteristics The streamwise distribution of the mean skin-fiiction coefficient (Cf) was found to be qualitatively consistent with existing literature on wall-shear measurements in BF S flows, including those that are based on direct measurements using oil—film interferometry. The value of Cf is positive but very small in the region extending from the separation point to x/xr z 0.33. From x/x, z 0.33 — 1.0, Cf is negative, with the maximum “reverse” Cf (Cflm) located at x/x, z 0.68. A universal empirical correlation of the form Cflm = aReH'o‘s, which was suggested in earlier studies, provided good representation of data obtained here and those from the direct shear measurements of Jovic and Driver (1996). Downstream of the reattachment position, the mean skin- fiiction coefficient slowly rises, reaching a steady value at x/x, z 2. IV. Turbulent shear-stress characteristics The rms skin-friction coefficient (C 'f) increased monotonically with increasing x, ‘ reaching a maximum in the vicinity of the mean reattachment location, followed by a slow decay thereafter. Examination of the power spectra revealed that the rise in C 'f with increasing streamwise distance from separation to reattachment is concentrated around the “middle” range of fiequencies that is associated with the passage of shear-layer vortices. The identification of vortex passage frequencies was achieved from examination of wall-pressure literature. In addition to the influence of the passing eddies at an average frequency of f =fx,/Uco z 0.65, shear-stress unsteadiness at a lower characteristic fi'equency of f z 0.1 was found to be dominant very close to the step. This 155 signature was attributed to quasi-steady modulation of the mean shear stress by low- frequency “flapping” of the shear layer. Consistent with the above observations from the power spectrum results, an autocorrelation analysis revealed a large integral time scale immediately downstream of the step. This integral scale decreased sharply in the region extending fi'om 0.25 — 0.5x,, but remained practically unchanged with further increase in streamwise distance. Overall, observations from both the power spectrum and autocorrelation of the wall-shear stress were qualitatively consistent with those documented in the wall-pressure literature. V. T wo-point characteristics of the turbulent shear-stress A unique feature of the present work is the compilation of what is believed to be the first two-point database of the wall-shear-stress field in a BFS flow. This was accomplished using one fixed, or reference, OHW sensor at xre/x, = 5.0711 and another movable sensor that was located at 14 different x locations ranging fiom 0.3H to 10H. The compiled data set was used to obtain the two-point space-time cross-correlation function of the wall-shear stress. For all Reynolds numbers investigated, the shape of the cross-correlation function reflected the dominant influence of the downstream-traveling shear-layer vortices on the wall-shear-stress field in the vicinity of the reattachment location. The downstream convection velocity of the vortical structures was obtained from the cross-correlation results. The resulting value, which represents an average over all time scales and x locations, was approximately 0.45 Um. A “local” determination of the convection velocity revealed that the vortical structures decelerated as they approached 156 the reattachment point. This deceleration was attributed to the mean adverse pressure gradient resulting from the pressure recovery in the range x/x, z 0.5 — 1.0. Because of the shorter x, found here relative to other studies of planar BFS flows, it is believed that the adverse pressure gradient is stronger, leading to a more pronounced deceleration effect. This apparently results in a lower convection velocity in comparison to values reported in the literature (0.5 — 0.6 Uco from wall-pressure data in various geometries). Additional information on the convection velocity was obtained through the use of phase analysis. This analysis yielded the dependence of the convection velocity on fi'equency. In agreement with wall-pressure results from data obtained on the same model (Hudy (2003)), the convection velocity exhibited a rise in magnitude with increasing frequency. However, the wall-shear-based phase data experienced a good deal of scatter, particularly at higher frequencies, because of the generally low values of the associated coherence (i.e., low degree of phase locking between the measurements obtained at two different locations). VI. Some difl'erences between the wall-shear and wall-pressure signatures Although there were a number of features found to be common between the wall- shear-stress field measured here and the wall-pressure in BFS and related flows in general, some differences were also observed. One difference related to the “spectral- peak switching” from the lower frequency corresponding to shear-layer flapping to the higher frequency associated with vortex passage as x increases beyond the reattachment location. This peak switching, which was found in pressure data measured on the same model and at similar Reynolds numbers (Hudy (2003)), was not manifested in the wall- 157 shear spectra. This is believed to be caused by the stronger attenuation of higher- frequency near-wall velocity (and hence wall-shear stress) fluctuations. The attenuation was clarified through the use of a uni-directional flow model that was found by Davenport and Sutton (1991) to provide a good description of near-wall velocity fluctuations in separating/reattaching flows. Another difference that was noted between wall-pressure and wall-shear measurements is that pertaining to the values of the two-point correlation and coherence. Generally, the correlation and coherence values decayed more rapidly in the wall-shear case with increasing spacing between the reference and movable sensors. This was attributed to the frequent change in shear-stress direction in the vicinity of the reattachment point, where the reference sensor was located. Because a wall-pressure sensor located at the same position would not be influenced by the near-wall flow- direction change, it should not suffer from the de-correlating effects of local direction changes. 6.2 Recommendations Notwithstanding the successful realization and use of the new OHW sensors in this study, some extra work is required to fine-tune their operation. In particular, the issue of noise induced from coupling of the piezo-driving signal to the hot-wire sensor requires some attention. As discussed in Chapter 3, this noise is partly responsible in the deviation of the direction-response of the sensor from the ideal ISO-degree phase jump. For the 2.8 kHz wire-oscillation frequency used here, the deviation was too small to influence the shear-direction determination. However, at an oscillation fiequency of 5 158 kHz, one of the .two sensors utilized here could operate properly, while the other one exhibited strong noise coupling fiom the driving signal that completely obscured the direction-response of the wire. Thus, there is a clear need to understand the nature of the coupling between the driving and sensing signals, the dependence of this coupling on the oscillation frequency and other geometrical/physical parameters of the sensor, and the best ways to minimize/eliminate this coupling. This should result in the ability to construct sensors that can operate at tens of kilohertz to resolve even wider bandwidths flows. Another direction in the development of the OHW sensors is to reduce their size and construct them in array configurations. Actually, a MEMS-based OHW wall-shear sensor has already been developed by the author for this purpose. More than one hundred such sensors were produced on a 100-mm diameter silicon wafer. This work verified the feasibility of microfabricating the sensors. However, it is anticipated that development of a few generations of these MEMS sensors will be required before the realization of a robust sensor and sensor arrays that can be used for detailed measurements such as those conducted here. Finally, it is recommended that some of the current measurements be repeated using smaller sensors, or even sensor arrays, to capture the cross-correlation at smaller sensor spacing than that realized here (0.15x,). This is motivated by the fact that the value of the cross-correlation drops significantly over the smallest spacing used here, suggesting that the influenCe of small-scale disturbances on the correlation could not be captured (e.g., the curvature of the correlation at zero separation, or Taylor microscale). It will also be interesting to examine the wall-shear signature in the region extending 159 from the step to 0.25x,, where Hudy et al. (2003) found an upstream convection velocity of wall-pressure fluctuations. Because of the sensor-size limitation, only one point of wall-shear measurements in this region could be realized in the current study. Thus, the use of smaller sensors with smaller separation is needed to overcome this limitation. 160 BIBLIOGRAPHY Adams, EW; Johnston, JP (1988) Effects of the separated shear layer on the reattachment flow structure Part 1: Pressure and turbulence quantities. Exp Fluids 6: 400—408 Adams, EW; Johnston, JP (1988) Effects of the separating shear layer on the reattachment flow structure Part 2: Reattachment length and wall shear stress. Exp Fluids 6: 493—499 Armaly, BF; Durst, F; Pereira, JCF; Schonung, B (1983) Experimental and theoretical investigation of backward-facing step flow. J Fluid Mech 127: 473—496 Bardbury, LJS; Castro, IP (1971) A pulsed-wire technique for velocity measurements in highly turbulent flows. J Fluid Mech 49: 657—691 de Brederode, V; Bradshaw, P (1972) Three-dimensional flow in nominally two- dimensional separation bubbles. 1. Flow behind a reaward-facing step. Dept. of Aeronautics. Imperial College of Science and Technology. Rept. 72-19 Coles, D; Wadcock, A (1979) Flying hot-wire study of flow past an NACA 4412 airfoil at maximum lift. AIAA J 17: 321—329 Chun, KB; Sung, HJ (1996) Control of turbulent separated flow over a backward-facing step by local forcing. Exp Fluids 21: 417- 426 Devenport, WJ.; Sutton, EP (1991) Near-wall Behavior of separated and reattaching flows. AIAA J 29: 25-31 Doebelin, E0 (1990) Measurement systems application and design. New York: McGraw-Hill, Inc. Downing, PM (1972) Reverse flow sensing hot wire anemometer. J Phys E Sci Instr 5: 849—85 1 Driver, DM; Seegmiller, HL; Marvin, JG (1987) Time-dependent behavior of a reattaching shear layer. AIAA J 25: 914-919 Driver, D; Jovic, S (1995) Reynolds number effect on the skin friction in separated flows behind a backward-facing step. Exp Fluids 18: 464-467 Eaton, JK (1980) Turbulent flow reattachment: an experimental study of the flow and structure behind a backward-facing step. Ph.D. Thesis. Stanford University 161 l Eaton, JK; Jeans, AH; Ashjaee, J; Johnston, JP (1979) A wall-flow-direction probe for use in separating and reattaching flows. J Fluids Eng 101: 364—366 Eaton, JK; Johnston, JP (1981) A review of research on subsonic turbulent flow reattachment. AIAA J 19: 1093-1099 Elger, DF; Adams, RL (1989) Dynamic hot-wire anemometer calibration using an oscillating flow. J Phys E: Sci Instruments 22: 166-172 Farabee, TM; Casarella; MJ (1986) Measurements of fluctuating wall pressure for separated/reattached boundary layer flows. J of Vibration, Acoustics, Stress and Reliability in Design 108: 301-307 Femholz, HH; Janke, G; Schober, M; Wagner, PM; Warnack, D (1996) New developments and applications of skin-friction measuring techniques. Meas. Sci. Tech 7: 1396-1409 Fureby, C (1999) Large eddy simulation of rearward-facing step flow. AIAA J 37: 1401—1410 Furuichi, N; Takeda, Y; Kumada, M (2003) Spatial structure of the flow through an axisymmetric sudden expansion. Exp Fluids 34: 643—650 Hanratty, TJ; Campbell, JA (1983) Measurement of wall shear stress. Fluid Mechanics Measurements ed Goldstein, R. J: 559-615 Hinze, JO (1959) Turbulence. McGraw-Hill, Inc. New York Heenan, AF; Morrison, JF (1998) Passive control of pressure fluctuations generated by separated flow. AIAA J 36: 1014-1022 Hudy, LM (2001) Simultaneous wall-pressure array and PIV measurements in a separating/reattaching flow region. MS. Thesis. Michigan State University Hudy, LM (2003) Private Communication Hudy, LM; Naguib, AM; Humphreys, WM (2003) Wall-pressure-array measurements beneath a separating/reattaching flow region. Phys. Fluids 15: 706-717 Jovic, S (1998) Recovery of reattached turbulent shear layers. Exp Thermal and Fluid Sci 17: 57-62 Kasagi, N; Matsunaga, A (1995) Three-dimensional particle-tracking velocirnetry measurement of turbulence statistics and energy budget in a backward-facing step flow. Int. J. Heat and Fluid Flow 16: 477—485 162 Khoo, BC; Chew, YT; Lim, CP (1998) The flow between a rotating and a stationary disc: application to near-wall hot-wire calibration. Meas. Sci. Tech 9: 650—658 Kim, KC; Yang, JP; Boo, JS (1996) An experimental study on the structure of a turbulent reattaching flow behind an axisymmetric backward-facing step. ASME. FED 237: 863—870 Kiya, M; Sasaki, K (1983) Structure of a turbulent separation bubble. J Fluid Mech 137: 83—1 13 Kiya, M; Sasaki, K (1985) Structure of large-scale vortices and unsteady reverse flow in the reattaching zone of a turbulent separation bubble. J Fluid Mech 154: 463—491 Kostas, J; Soria, J; Chong MS (2002) Particle image velocirnetry measurements of a backward-facing step flow. Exp Fluids: 838—853 Lai, JSC; Yue, J; Platzer, MF (2002) Control of backward-facing step flow using a flapping foil. Exp Fluids 33: 44—54 Le, H; Moin, P; Kim, J (1997) Direct numerical simulation of turbulent flow over a backward-facing step. J Fluid Mech 330: 349—374 Lee, I; Sung, HJ (2001) Characteristic of wall pressure fluctuations in separated flows over a backward-facing step: Part I. Time-mean statistics and cross-spectral analysis. Exp Fluids 30: 262—272 Lee, I; Sung, HJ (2001) Characteristic of wall pressure fluctuations in separated flows over a backward-facing step: Part II. Unsteady wavelet analysis. Exp Fluids 30: 273—282 Lee, I; Sung, HJ (2002) Multiple-arrayed pressure measurement for investigation of the unsteady flow structure of a reattaching shear layer. J Fluid Mech 463: 377—402 Li, Y; Naguib, AM (2003) An oscillating hot-wire technique for resolving the magnitude and direction of velocity measurements using single hot-wire sensors. Exp Fluids 34: 597-606 Morris, SC (2002) The velocity and vorticity fields of a single stream shear layer. Ph.D. Dissertation. Michigan State University Naguib, AM (1992) Inner- and outer-layer effects on the dynamics of a turbulent boundary layer. Ph.D. Dissertation. Institute of Illinois Technology Osterlund, JM; Johansson, AV; Nagib, HM; Hites, MH (1999) Wall shear stress measurements in hi Reynolds number boundary layers from two facilities. AIAA paper 99-3814. 30t AIAA Fluid Dynamics Conference, 28 June — 1 July, Norfolk, Virginia. 163 Otugen, MV (1991) Expansion ratio effects on the separated shear layer and reattachemnt downstream of a backward-facing step. Exp Fluids 10: 273-280 Padmanabhan, A (1997) Silicon micromachined sensors and sensor arrays for shear- stress measurements in aerodynamic flows. Ph.D. Dissertation. Massachusetts Institute of Technology Rinoie, K; Shirai, Y; Sunada, Y (2002) Behavior of separated and reattaching flow formed over a backward facing step. Trans. Japan Soc. Aero. Space Sci 45: 20-27 Scarano, F; Benocci, C; Riethmuller, ML; (1999) Pattern recognition analysis of the turbulent flow past a backward facing step. Phys. Fluids 11: 3808-3818 Scarano, F; Riethmuller, ML; (1999) Iterative multigrid approach in PIV image processing with discrete window offset. Exp Fluids 26: 513—523 Schlichting, H (1979) Boundary-layer theory. McGraw-Hill, Inc. Schmidt, MA; Howe, RT; Senturia, SD; Haritonidis, JH (1988) Design and calibration of a microfabricated floating-element shear-stress sensor. IEEE Trans. Electron Devices 35: 750—757 Schober, M; Hancock, PE; Siller, H (1998) Pulsed-wire anemometry near walls. Exp Fluids 25: 151—159 Spalart, PR (1988) Direct Simulation of a Turbulent Boundary Layer up to Reg = 1440. J Fluid Mech 187: 61-98 Spazzini PG; Iuso G; Onorato M; Zurlo N (1999) Design, test and validation of a probe for time-resolved measurement of skin friction. Meas. Sci. Tech 10: 631—639 Spazzini, PG; Iuso, G; Onorato, M; Zurlo, N; Cicca, GMD (2001) Unsteady behavior of back-facing step flow. Exp Fluids 30: 551-561 Stewartson, K (1953) Flow between two rotating coaxial discs. Proc. Camb. Phil. Soc 49: 333-4 Thompson, BE; Whitelaw, JH (1984) Flying hot-wire anemometry. Exp Fluids 2: 47— 55 Tihon, J; Legrand, J; Legentilhomme, P (2001) Near-wall investigation of backward- facing step flows. Exp Fluids 31: 484-493 Watmuff, JH; Perry, AE; Chong, MS (1983) A flying hot-wire system. Exp Fluids 1: 63—71 164 Westphal, RV; Eaton, JK; Johnston, JP (1981) A new probe for measurement of velocity and wall shear stress in unsteady reversing flows. J Fluids Eng 103: 478—482 Westphal, RV; Johnston, JP (1984) Effect of initial conditions on turbulent reattachment downstream of a backward-facing step. AIAA J 22: 1727—1732 White, FM (1974) Viscous fluid flow. McGraw — Hill, Inc. New York 165