5 I l » ‘/ 5": —0\. ‘3! n- u. 52 £5" : it“ it. ‘ 1 fi “:54"— A w- TlJVTClq LIBRARY Michigan State University This is to certify that the dissertation entitled The Velocity and Vorticity Fields of a Single Stream Shear Layer presented by Scott C. Morris has been accepted towards fulfillment of the requirements for Ph.D. degree in Engineering MMrofessor 97690”? 0 Date 61%»? 2002, MS U is an Affirmative Action/Equal Opportunity Institution 0-12771 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 cJCIRC/DateDuepGSm. 15 THE VELOCITY AND VORTICITY FIELDS OF A SINGLE STREAM SHEAR LAYER By Scott C. Morris A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Mechanical Engineering 2002 ABSTRACT THE VELOCITY AND VORTICITY FEILDS OF A SINGLE STREAM SHEAR LAYER By Scott C. Morris Measurements of velocity and vorticity have been acquired in a large scale single stream shear layer. These data provided information regarding the nature of turbulent fluid flow at a relatively high Reynolds number. The Reynolds number dependence of phenom- ena is important given that many technological applications involve Reynolds numbers that are higher than those which are typically achieved in the laboratory. The use of the vorticity vector as an alternative variable to the velocity vector provides additional insight into the nature of these phenomena. Single and multi point measurements of both the velocity and vorticity have allowed the examination and comparison of the different scales of motion. The research project was effectively divided into two parts. The first involves the very near separation region in which a turbulent boundary layer separates at zero pres- sure gradient at a 90 degree edge. The measurements have shown the existence of a "sub- shear" layer.. The vorticity that participates in the first instability is found to originate from the very near wall region. At larger y values. the separated turbulent boundary layer convects downstream adjacent to the sub-shear layer. "seemingly unaware" of the separa- tion. The second part of the research involves measurements in the developed region of the shear layer. Several conclusions have been drawn from these data. For example, it has been shown that the dimensionless vorticity fluctuations do not scale in a self-similar way, but increase with Reynolds number. The cause of this is related to the large scale motions of the flow that lead to nonzero spatial correlations in vorticity over relatively large length scales. DEDICATED TO THE MEMORY OF MY FATHER, JOHN E. MORRIS ACKNOWLEDGMENTS I would like to thank my committee members. Charles Petty, Ahmed Naguib, Manoochehr Koochesfahani, C.Y. Wang, and John Foss for their contributions and com- ments on this research. Special thanks goes to my advisor, John Foss, for providing me with an environment in which creativity and ideas can be explored. I would also like to thank a number of my colleges whom I have worked with over the last several years. This includes Dr. Alptekin Aksan. Dr. Paul Hoke, Doug Neal, Richard Prevost, Al Lawrenz. Matt Maher. and many others who helped with the assembly of the shear layer facility. Lastly, I thank my wife Yvonne. and son John. Research would not be worth the trou- ble if it weren’t for family. TABLE OF CONTENTS List of Figures ................................................................................................................... vii List of Tables ..................................................................................................................... xii Nomenclature ................................................................................................................... xiii 1 .0 Introduction .............................................................................................................. l 2.0 Experimental Apparatus and Techniques ................................................................. 8 2.1 The Single Stream Shear Layer Facility ...................................................... 8 2.2 Hot-wire Techniques and processing ......................................................... 11 2.2.1 Calibration ..................................................................................... 11 2.2.2 X-wire probe calibration and processing ....................................... 13 2.2.3 Multi-sensor Probe Configurations and Processing ...................... 15 3.0 Turbulent Boundary Layer to Single Stream Shear Layer .................................... 28 3.1 Introduction ................................................................................................ 28 3.2 Upstream Boundary Layer ........................................................................ 31 3.2.1 Boundary Layer - Shear Layer “Communication” ........................ 31 3.2.2 Boundary Layer Statistics Near x=0. ............................................. 33 3.3 Point Statistics And The “Sub-shear Layer” .............................................. 37 3.4 Shear Layer Diagnostics From Entrainment Measurements ..................... 41 3.4.1 Convection Speed of the Coherent Motions .................................. 42 3.4.2 Spectral Properties of the Entrainment Stream .............................. 44 3.4.3 The Initial Instability ..................................................................... 47 3.5 Spanwise Correlation Field of the Near Separation Region ..................... 51 3.5.1 Example realizations in the irrotational stream ............................. 52 3.5.2 Correlation Coefficient Data .......................................................... 54 3.6 Summary and Conclusions ........................................................................ 55 4.0 The Developed Single Stream Shear Layer ........................................................... 88 4.1 Introduction ................................................................................................ 88 4.1.] Motivation and literature survey .................................................... 89 4.1.2 Measurement program ................................................................... 92 4.2 The Velocity Field ..................................................................................... 94 4.2.1 Integral constraints on the mean flow ........................................... 94 4.2.2 Velocity in Self-Similar Coordinates ............................................. 97 4.2.3 Profiles of the velocity gradient variance .................................... 103 4.3 Statistics of the Vorticity Field ................................................................. 109 4.3.1 Examples of Vorticity Time Series .............................................. 109 4.3.2 Histograms of Vorticity ................................................................ 110 4.3.3 Profiles of the Vorticity RMS ...................................................... 1 13 4.4 Analysis of the Reynolds Stress ............................................................... 116 4.5 Turbulent Kinetic Energy ......................................................................... 120 4.5.1 Dissipation Measurements. .......................................................... 121 4.5.2 Measurement of Kinetic Energy Budget ...................................... 127 4.6 The Scales of Turbulent Motion .............................................................. 129 4.6.1 Correlation measurements of velocity and vorticity .................... 129 4.6.2 Spectral representation of velocity and vorticity ......................... 137 4.7 Circulation Density .................................................................................. 143 4.7.1 Experimental Configuration ........................................................ 147 4.7.2 Circulation Density Results ......................................................... 148 4.8 Conclusions .............................................................................................. 150 Appendix A .......................................................................................................... 193 REFERENCES .................................................................................................... 197 vi LIST OF FIGURES Figure 1.1 Idealized schematic of the single stream shear layer geometry. ....................... 7 Figure 2.1 Primary flow delivery system .......................................................................... 21 Figure 2.2 Schematic of SSSL facility ............................................................................. 22 Figure 2.3 Geometry of single hot-wire probe ................................................................ 23 Figure 2.4 Schematic of Calibration Facility .................................................................... 24 Figure 2.5 Sample calibration and fit, y=0 degrees. ......................................................... 24 Figure 2.6 Geometry of the X-wire probe ........................................................................ 25 Figure 2.7 Example of the locating algorithm of the intersection point for X-wire processing. ....................................................................................................... 25 Figure 2.8 Predicted values of angle (7) from re-processed calibration data at nominal values of 18, 0, and -30 degrees. ...................................................... 25 Figure 2.9 Schematic of spanwise vorticity probe ............................................................ 26 Figure 2.10 Schematic of micro-circulation used for the spanwise vorticity calculation..26 Figure 2.11 Schematic of the streamwise vorticity probe. ............................................... 27 Figure 2.12 Schematic of the multiple X-wire probe. ...................................................... 27 Figure 3.1 Schematic representation of near separation region. ....................................... 63 Figure 3.2 Boundary Layer Mean Velocity. ...................................................................... 63 Figure 3.3 Boundary Layer Velocity RMS ....................................................................... 64 Figure 3.4 Mean and RMS of vorticity at separation. ...................................................... 64 Figure 3.5 Spectrum of vorticity fluctuations at y/5=0.6. ................................................. 65 Figure 3.6 Sample flow visualization of the near separation region. ............................... 65 Figure 3.7 Figure of flow regions; not to scale. ................................................................ 66 Figure 3.8 Mean velocity profiles for 0l< u,v,w NOMENCLATURE Coeficients of hot-wire calibratirm. Chapter 2 only. Voltage output of the anernometer, Chapter 2 only. Components of the one dimensional velocity spectra. Chapter 4 only. Preffered frequency observed in irrotatoinal flow. Coherence spectra Wave number in the x direction Ratios of velocity gradient variances. Magnitued of the velocity vector in the x-y plan Normalized correlation function. Reynolds number of the boundary layer at separation Reynolds number based on local momentum thickness Dimensionless time deley used in autocorrelation funciton Convection velocity of large scale motions measured in the irrotational stream Freestream velocity Effective freestream velocity for scalin g of the first instability Components of the velocity vector xiii X,y,Z 8* As,An wx,my,mz Velocity of the entrainment flow Carteasian coordinates originating from the sparation point Displacement thickness Separation of straight wire sensors in the vorticity probe. Size of the microcirculation domain used in the vorticity calculation. Dissipatoin Angle of the velocity vector with respect to an X-wire probe, Chapter 2 only. Circulation density, Chapter 4 only. Three dimensional wave number. Momentum thickenss of boundary layer at separation Local momentum thickness Effective momentum thickess for scaling the first instability Spreading parameter. Components of the vorticity vector xiv 1.0 Introduction The study of fluid mechanics at high Reynolds number is quite unique among the physical sciences. This uniqueness can be partially characterized by the full knowledge of the gov- erning equations of fluid motion without a complete understanding of the phenomenon, despite over a century of intense research. One source of the difficulty in turbulence research is the number of different phenomenon that occur with different boundary condi- tions. As a result, theory, modeling, or experiments which characterize a single given set of flow conditions will usually fail to provide similar insight for a geometry which is dif- ferent from the original. Most studies in fluid mechanics can therefore be categorized as belonging to one of two research “modes.” The first, is to choose a set of boundary conditions which is relevant to an important engineering problem. For example. most research in turbomachinery would fall into this category. The second mode, is to chose a small number of canonical flows in which the physics of the fluid motions can be studied in great detail. One goal of these fundamental studies is to incorporate the acquired knowledge into the applied studies, and ultimately into engineering design. This dissertation follows the second approach, and utilizes the single stream shear layer as the canonical flow field of interest. This geometry is. in the author’s opinion, a very sim- ple one which is applicable to the wide range of potential applications. An idealized sche- matic of the planar flow geometry is shown in Figure 1.1. The irrotational freestream enters the domain from the left, with a smooth wall boundary layer developing on the only bounding wall. The boundary layer undergoes transition to a turbulent boundary layer, and develops with a zero pressure gradient (constant velocity) free stream. After a specified length, the boundary layer plate ends with a 90 degree angle as shown. This defines the streamwise location of the separation point which marks the transition from a wall bounded shear flow to a free shear flow. The free shear flow is characterized by a zero pressure gradient free stream on one side, and fluid which is “entrained” into the shear layer from the other side. From this basic description of the flow field, it is considered to be apparent that many of the important flow phenomena of applied fluid mechanics can be studied in the single stream shear layer. It can be observed from Figure 1.1 that the only length scale of the problem is given by the length of the boundary layer plate. An equivalent characterization which is often preferred is the momentum thickness of the boundary layer at separation (80). This, along with the specification of the free stream velocity (U0) and the kinematic viscosity of the working fluid (v) completely characterizes the boundary value problem. The only parameter is the Reynolds number, Re“: U 0 /v. One of the unique features of the present research a 0 project is the large scale of the experimental facility which was constructed in the Turbu- lent Shear Flows Laboratory at Michigan State University. This facility, which is described in detail in Chapter 2, uses air as the working fluid and has the characteristics: UO=7.1m/s, 60:9.62mm. This provides a Reynolds number at the separation point of Reo=4650. Note that the dimensionless entrainment velocity (VG/U”) is not an independent parameter. This is because the boundary conditions specify that the free stream velocity (and pres- sure) are streamwise invariant. That is, BUD/8x = Eli/ax: 0 .There exists a “natural” value of ve/U0 that will allow the shear layer to grow in a zero pressure gradient environ- ment. The magnitude of the entrainment velocity is therefore not known a priori, and must be determined experimentally. This may seem odd at first, that the boundary conditions of an experiment are not known before the experiments are started. To clarify why this is the case, consider a finite entrainment domain with vc=(). This represents the backward facing step geometry in which the separation streamline must diverge away from the high speed fluid and re-attach to a bounding wall, leading to a recirculation region in the entrainment side and an adverse pressure gradient condition in the free stream region. In contrast, if ve were to be forced to be arbitrarily large, the shear layer would deflect towards the high speed fluid, thereby accelerating the shear layer in a favorable pressure gradient condition. The “natural” value of V6 is that which would be observed if the entrainment region were of infinite extent for a finite height primary flow. Since a truly infinite domain is not experimentally possible, the entrainment flow is “fed“ into the shear layer at the rate which provides the zero pressure gradient conditions. The experimental determination of the entrainment velocity is described in Chapter 2. There exists a large number of research papers in the literature which have focused on the single stream shear layer. In addition, there is a distinctively larger body of literature which examines the two-stream shear layer. The latter flow field uses a thin splitter plate to divide two parallel fluid streams with different velocities: U, and U2. The shear layer is created by the trailing edge of the splitter plate. It is incorrect to think of the single-stream shear layer as the limiting case of a two-stream shear layer with the slower velocity reduced to zero. The reason for this is the direction of the entrainment stream, which, in the single stream shear layer is directed perpendicular to the primary flow direction. In the two stream shear layer the entrainment phenomenon caused by the high speed flow will “pull” fluid from the second flow region, leading to a complicated and undesired flow field. The difference between parallel and perpendicular entrainment streams has been examined by Schmidt et a1. (1986). In summary, despite many useful similarities, the sin- gle stream shear layer is a distinct flow field from the two stream flow, and not simply the limiting case of U220. As already noted, there are a number of research papers in the literature which have docu— mented the physical characteristics of the single stream shear layer. These include Liep— mann and Laufer (1947), Freymuth (1966), Wygnanski and F iedler (1970), Champagne et al. (1976), Hussain and Zedan (1978), Hussain and Zaman (1985), Mehta and Westphal (1986), Haw et a1. (1989), Bruns (1990), Bruns et al. (1991), Foss (1994). Given the large amount of information already available on the subject. it is appropriate to establish the reason for the continued interest in this flow field. It is noteworthy, in this context, that Wygnanski and Fiedler in 1970 found it necessary to justify why they found it “worth- while to reconsider this flow”. Following an additional three decades of research, an updated justification is given by outlining several of the open questions which still exist in the literature whose answers are of considerable interest. The affects of the inflow boundary conditions on the downstream flow are of interest in the study of both single and two stream shear layers. The dynamics of the flow field in the very near separation region has not been documented or properly understood. As a result, most differences that are noted between the results of different researchers using different shear layer facilities are attributed to the differences in the inflow conditions. In addition, the inflow conditions for most experimental facilities are either laminar, or low Reynolds number tripped boundary layers. These facilities are therefore not representative of the flow separation phenomenon that exist in engineering applications which are character- ized by very high Reynolds numbers. The experimental facility used in the present work provides a unique opportunity to “fill the gap” in the available literature. This is because of the relatively large scale of the facil- ity, which provides excellent spatial resolution for measurements (599(xz0) z 100mm) in a moderately high Reynolds number boundary layer (Re0=4650). A detailed experimental investigation of the near separation field is presented in Chapter 3. This research project is self-contained, and distinctive from the downstream measurements. Therefore, Chapter 3 is complete in the sense that an introduction, motivation. background, literature review, experimental results. and conclusions are provided. The next portion of this dissertation, Chapter 4. describes measurements that were acquired in the self similar region of the flow field. This chapter is also self contained with introduction, literature review, etc. There are numerous outstanding questions in the research literature which have motivated these measurements. A first example regards the nature of shear layers in terms of how the stochastic variables of interest scale with down- stream distance. It has long been known that linear growth and self-similarity are standard features of free shear flows. However, the scalin g of quantities such as dissipation, vortic- ity fluctuations and their spectra have not previously been examined. Again, the present facility provides a unique opportunity to investigate these issues because of the relatively large scale, and hence adequate spatial resolution of the measurements. It will be shown in terms of the vorticity spectra that true self-similarity is not achieved. Rather, the dimen- sionless vorticity fluctuations increase with Reynolds number. In addition to the novel research issues which have been addressed in this dissertation, many of the previously documented features of shear layers have been repeated in this work. The reason for this repetition is (i) to validate the present facility, and (ii) to validate claims made in previous research efforts. In doing so, this volume represents a more com- prehensive study of the flow field than that which has been previously reported. The outline of this document can be summarized as follows. The experimental facility and measurement techniques which are common to all of the experiments are described in Chapter 2. The two “halves” of the research project are represented by the near separation region, and fully developed region, and are described in Chapters 3 and 4 respectively. UO 1“) _) We, '—> r—> ——) 1—9 I; R" _) : 11111111111111111111111111111111111111111111111111111 Ve Figure 1.1 Idealized schematic of the single stream shear layer geometry. The thick line represents the only solid boundary. The dotted lines bound the region of non-zero vortic- ity. The velocity vectors are schematic, and not shown to scale. 2.0 Experimental Apparatus and Techniques The experimental apparatus and methods used in this dissertation are described in this chapter. First, the Single Stream Shear Layer Facility (SSSL) will be described. This facil- ity was designed and assembled by the author, and is documented in Section 2.1. This information can serve as a reference for future research which will utilize the facility. Hot-wire anemometry was the primary experimental method for the measurement of velocity. The basic sensor geometry and calibration technique are described in Section 2.2.1. X-wire probes were used to measure one or two components of the velocity vector. The algorithm for obtaining the fluid speed and flow angle from the two sensors of the X- wire is outlined in Section 2.2.2. Multi-sensor arrays of wires were used to obtain velocity gradient and vorticity information. These are discussed in Section 2.2.3. 2.1 The Single Stream Shear Layer Facility The Single Stream Shear Layer (SSSL) facility is located in the Turbulent Shear Flows Laboratory at Michigan State University. The majority of the facility including the test section was constructed at an elevated level (its floor is ~2.1m) above the laboratory floor. The primary flow intake and delivery system were located at the floor level as shown in Figure 2.1. An axial fan accelerates the laboratory air from rest. The fan exit is connected to a 1m x 1m square section. Acoustic dampening material was placed 1.4m upstream from the blower entrance to limit the propagation of blower noise into the room and test section. The fan used a constant speed AC motor attached directly to the 48” diameter blades. The exit of the axial fan housing was connected directly to a diffuser section that expands to a 2m x 2m cross section. In this diffuser. flow dividers (not shown) were used to limit flow separation and provide relatively equal flow distribution. The air is delivered through two sets of 90 degree turning vanes which first turn the flow direction vertically upwards and then horizontally at the elevation of the test section. Standard Sum furnace filters were attached to the first set of turning vanes to eliminate dust from the free stream flow. A second acoustic dampening wall was placed at the back wall of the turning vane section as shown. Following the second turning vane section, a set of flow conditioners were installed to provide a uniform velocity with minimal unsteadiness to the test section. This condition- ing section consisted of a 25mm long honeycomb section with 3mm tube diameters, fol- lowed by three sections of stretched steel screen material. The screens were made from 0.060” wire mesh with 40 wires per inch (0.635mm spacing between wires). The outflow from this conditioning section is connected directly to the inlet of the main tunnel entrance. This is shown as “primary inflow" in Figure 2.2. The height (i.e., “out of the page”) of the entire working section of the tunnel was 1.96m. The width of the primary flow inlet was 1.96m wide. A planar contraction was placed downstream from the inlet which accelerates the flow to a width of 1.1m. The elliptical shaped leading edge of the boundary layer plate then splits the primary flow such that the boundary layer from the curved wall of the contraction is “bled” off and returned to the laboratory as shown. This point of the tunnel and all downstream locations are at nearly zero gage pressure with respect to the laboratory pressure. The majority of the flow moves through the 0.98m wide boundary layer development section. The boundary layer plate is 5.74m long, with a free stream velocity of 7.1m/s. The boundary layer plate was laminated with 1mm thick PVC sheeting to provide a smooth boundary condition. The boundary layer plate ends at the separation point which was created using a machined 90 degree edge. The free stream velocity was measured during all experiments by monitoring the differential pressure drop across the planar contraction. This pressure difference was acquired during all experi- ments and found to vary by less than 0.5% for the data reported in Chapters 3 and 4. The shear layer test section begins at the separation edge. The entrainment flow is pro- vided by four 48 inch (1.2m) diameter axial propeller type fans with an AC speed control- ler. This fan type is designed to move large volumes of air at relatively low pressure rise (~25Pa). For this reason, bypass holes were created such that the fans would operate at design conditions to prevent unsteadiness from blade stall. That is, most of the air moved by the fans is recirculated back to the laboratory. The “plenum” between the fans and the flow conditioning is maintained at a pressure which is slightly higher than the atmospheric value. The entrainment flow conditioning is identical to the primary flow conditioning. The length of the entrainment screens is 9.5m which defines the working length of the shear layer test section. The location of the final entrainment screen is y=-2.75m, using the coordinate system shown with origin at the separation edge. The exit of the tunnel is open to the laboratory. The nearest obstruction to the outflow is a building wall located 6.1m from the tunnel exit. This provided sufficient room to turn the exiting flow without dis— turbing the shear layer. The entrainment velocity was set using an AC frequency control unit to drive the electric motors connected to the entrainment fans. An experiment was conducted to identify the correct motor speed in order to create a shear layer with a zero streamwise pressure gradi- ent. Specifically, data were obtained using a single hot-wire probe in the free stream at two streamwise locations (x=0.5m and 4.0m) in the high speed irrotational flow. The controller 10 frequency was adjusted until the hot-wire readings at these two locations indicated the same velocity. This implies dUO/dx= 0, i.e., a zero pressure gradient condition. A traverse system was constructed to support and move the hot-wire sensors during data acquisition. This system used a wheel and track for movement in the streamwise direction (x) and a precision lead screw and stepper motor drive for the transverse (y) direction. The probes were mounted on a rotating shaft which was also motor driven. This permitted angular positioning of the sensors for calibration and data acquisition. 2.2 Hot-wire Techniques and processing This section will describe the details of the hot-wire calibration and processing methods used in this work. A schematic of a single sensor is shown in Figure 2.3. The active por- tion of the sensor was a 1mm long, 5am diameter Tungsten wire. The end regions were electro—plated with 30pm copper and soldered to stainless steel supports. The wires were connected to DISA 55m anemometers. The frequency response of the sensors at 7m/s was typically greater than 20kHz. 2.2.1 Calibration The wires were calibrated using a newly constructed calibration facility shown schemati- cally in Figure 2.4. The unit used a constant speed cross-flow blower and throttle to vary the exit velocity from 0.5 to 12m/s. Screens, filters and honeycomb were placed upstream of a contraction to achieve a low disturbance free stream for calibration. The contraction exit flow dimensions were 2.5 x 24 cm which provided sufficient area to calibrate up to 10 probes simultaneously. The probe holder was connected to a stepping motor which varied the angle of the X-wire probes with respect to the jet exit. The calibration technique used a continuously variable “quasi-steady” method. That is, the sensor was subjected to a velocity which was varied continuously from shutoff to maxi- mum flow over the duration of approximately 40 seconds. Data samples of the hot- wire(s), pressure transducer, and thermistor were acquired at a sample rate of 500Hz. These data were used to fit the equation 52 = Am + BmQ”“” 2.1 where E represents the output voltage from the anernometer, and Q represents the exit velocity, and y represents the pitch angle with respect to the probe axis. This velocity was calculated from the pressure measurement using Bernoulli’s equation. Note that this method assumes that the unsteady effects of the continuously varied velocity are not suffi- cient to cause error in the steady Bernoulli equation. This assumption was tested by per— forming several standard (steady state) calibrations for comparison with the quasi-steady results. These experiments have shown no measurable differences between the two types of calibratoins.The primary reason for using the quasi-steady method is the speed of cali- bration. For example, single wires were calibrated in less than one minute. The x-wire probes, for example, were calibrated at 13 angles which required approximately 12 min- utes of total calibration time, which include angle changes between measurements. If the sensors were calibrated with the alternative technique: establish the flow angle, record time series data for a period of 10 seconds at each flow speed, then an X-wire probe would take typically 45minutes to calibrate. An example of a wire calibration and fit is shown in Figure 2.5. The quality of the fit was determined from the standard deviation between the measured velocity and the predicted velocity from equation 2.1. In all experiments, this was found to be less than 0.05m/s. If a 12 sensor/anemometer were found to have a higher value for the standard deviation, the probe was not used. The sensors were calibrated before and after each experiment. If the predicted velocity from the two calibrations changed by more than 2%, the data were dis- carded and the experiment was repeated. In most experiments, the velocity calculated from the “pre-calibration” and “post-calibration” agreed to better than 1%. 2.2.2 X-wire probe calibration and processing X-wire probes were used for the simultaneous measurement of two velocity components. The probe geometry is illustrated in Figure 2.6. Two sensors were oriented at nominally :45 degrees to the probe axis. The spacing between the sensors was 1mm. The output from the two sensors is denoted as E1 and E2. The goal of the calibration and processing algorithm is to obtain a transfer function between these voltages and the desired variables: the speed (Q) and angle (7) of the fluid velocity with respect to the probe. The optimal transfer function, which provides the most accurate estimate of Q and 7, has been the sub- ject of considerable research. Techniques which are used in the literature vary from the very simple cosine law to very complex algorithms such as mapping techniques; these are described by Bruun ( 1995). Since all techniques seem quite accurate for small angles, say, [71 < 12 degrees, it is the largest angles where the differences between algorithms has the most effect. Because the angles found in the shear layer can fluctuate considerably, a new measurement technique was developed which provides the most reliable speed and angle estimates for 171 < 36 degrees. The present technique, is a modification of the “speed-wire/angle-wire” (SWAW) tech- nique proposed by Foss (1981). The SWAW algorithm uses the sensor more parallel with the velocity vector to determine the angle (7), and the sensor which is more perpendicular to the velocity vector is used to obtain the speed (Q). An iterative algorithm is used which converges to the correct (Q,y) pair. The present technique is similar, although the iterative nature of the SWAW technique was replaced with a “curve intersection” method. The new algorithm is described as follows. First. the X-wire probe is calibrated using the spin—down technique (described in Section 2.2.1) in the range —36 < y< 36 degrees, in 6 degree intervals. This calibration results in 13 sets of (A,B,n) values for each sensor. The experimental data are then acquired as time series values of the voltage pair (E1,E2). From each of the recorded (EI,E2) pairs, equation 2.1 can be solved for Q1(E1,‘y,-), Q2(E2,y,-) where i=1,2,...13. That is, Q1 and Q2 represent predicted velocities which are calculated for each sensor using the wire voltages (E1, E2). and the calibration coefficients for i=1,2...13 representing the 13 angles used in the calibration. Figure 2.7 shows an example of the 13 predicted velocities that were calculated from a voltage pair. The “true” value of (Q,‘y) is now represented by the intersection of the two curves shown. This intersection point is found by fitting the neighboring points of each curve with a second order polyno- mial fit, and then solving for the intersection point. One method of testing the accuracy of the calibration/processin g algorithm is to calculate the speed and flow angle of an X-wire probe placed a known angle with respect to a steady flow. The calculated velocity agreed with the actual value to within 2% at angles less than 18 degrees, and within 4% at angles less than 36 degrees. Figure 2.8 shows the results of time series data of the predicted flow angle with a probe placed at +18, 0, and -30 degrees with respect to the jet of the calibration unit. The standard deviation of the error in pre- dicted angle was typically less than 0.5 degrees. and often as good as 0.2 degrees. These errors reflect the contribution of electronic noise to the processed time series. Errors due to calibration drift were typically less than 1 degree. 2.2.3 Multi-sensor Probe Configurations and Processing The previous sections have described the calibration and processing algorithms for both single wire probes and X-array probes. In the following three sub-sections, probes which utilize combinations of single and X-arrays to compute vorticity and velocity gradients are described. The use of multi-sensor probes leads to a number of complications regarding the uncertainty of the measurements. These can be dependent on the probe configuration, probe-scale to flow-scale effects, and the presence of electronic noise in the time series. Because the measurement techniques described were not developed as a part of this dis- sertation, they will be described only briefly along with the estimated uncertainty. Addi- tional information can be found from references to those communications in which the developed techniques are described. These references provide more comprehensive uncer- tainty evaluations. 2.2.3.1 Spanwise Vorticity Probe A schematic representation of the spanwise vorticity probe, or “Mitchell” Probe, is shown in Figure 2.9. This probe consists of an X-array with two parallel wires. These were cali- brated using the procedures described above. The time series output from the probe is the magnitude of velocity in the plane of the probe as determined by the two straight wires: (ql, qz), and the angle (7) determined from the X-array. This section will review the method by which the time series values of (q,, (12.7) were used to calculate the spanwise vorticity. The probe geometry and processing algorithm were originally developed by Foss and co-workers, and can be found in the references by Foss and Haw (1990), and Wallace and Foss (1995). The data processing algorithm can be summarized as follows. The average of the two recorded velocities (q,+q2)/2 is used as a convection velocity to create a parallelogram shaped micro-circulation domain using a local (in time) Taylor’s hypothesis; see Figure 2.10. The height (Ay) of the microcirculation is fixed by the physical separation distance between the two straight wire sensors. The length of the domain (As) was set to be roughly equal to Ay. Specifically, from the discrete time records of ql(tn), etc., the sum was com- puted for each tn as: Asn(m) = 2.2 l l'=""‘“(ql(t,)+r‘12(t,-))At 2m +1 2 ' i=n-m The value of m was increased, starting from m: 1, until As" > Ay. This defines M, the integer number of time records needed to create the one half of the micro-domain. The value of M for each calculation time (tn) would vary depending on the flow speed. Typical values of 2M were between 6 and 20 time steps. The mean value of a given quantity during the interval In _ M < t < tn + M is denoted by the brackets ( > . For example, the angle of the microcirculation parallelogram was the mean value of yduring the time period of the microcirculation: i=n+M Z 70,-) 2.3 i=n-M l 2M+l = Once the microcirculation domain has been “created“. the average vorticity of the domain is calculated from the velocity circulation around the domain. The magnitude of the veloc- ity can be projected onto the “top” and “bottom” of the micro-domain as q,-cos(y(ti)—(y>) and qz-cos(y(t,)—(Y>) . and then integrated along the length circulation domain. The normal velocity component is calculated from v = (qI + q2)( siny)/2 at both the upstream and downstream sides of the domain. For each point of the time series (n), the vorticity is calculated from: l §(-‘7d§)= VUH+M)—v(tn—M)_ = (AS)(A,v) As Ay 2'4 The approximation cos( (y(tl.)—(y>) z 1 ) was made in order to decrease the computational time required to process the time series. This assumption is justified based on the realiza- tion that the differences between the instantaneous values of y( t,) and (7) were less than 5 degrees. Since the cosine can be factored out from the q. and q2 terms, the approximation cos(lyj < 5) z 1 is reasonable. A detailed uncertainty analysis of the vorticity probe and processing algorithm has been given by Foss and Haw (1990). One of the primary sources of error was found to be due to the presence of electronic noise in the time series data. The contribution of electronic noise to the uncertainty is reported to be of order 20 ( l/sec’) in Foss and Haw (1990). This is well supported by the measurements of vorticity in the free stream of the shear layer where little or no vorticity is assumed to exist. Specifically, at x/0(,=384, values of (Bz < 25(1/sec) were measured at the transverse locations (n>3.8). 2.2.3.2 Streamwise Vorticity Probe A schematic of the streamwise vorticity probe is shown in Figure 2.1 1. This probe uses 8 senors configured as four X-arrays in a square pattern. From this, two of the opposing X- wires will provide time series of (VI, v2), separated by a distance of Az=5.6mm. The other two X-wire will then measure time series of (WI. wz). separated by Ay=5.6mm. The vor- ticity calculation is then straightforward: - Ay A2. The uncertainty of this probe is dictated by the ability of the X-wire processing to recover the true value of the lateral component of velocity. The errors will be due to electronic noise and calibration errors. Electronic noise was shown in Section 2.2.2 to contribute to a random error of typically 0.2 degrees. At a flow speed of, say. 7.0m/s, this would lead to a random error of roughly 24mm/s in the normal component of velocity. If these errors are Gaussian distributed and equal for the four X-wire probes. then the standard deviation of the sum (w2 — wl — v2 + vl) is twice that of the single measurement. That is the standard deviation for (it)x in an irrotational flow would be 2 . (0.024)/(0.005) = 9.8(1/sec) . The measurements support this conclusion. For example. at x/0 =384, values of () (I)Z < 8 (1/sec) were measured at the transverse locations (n>3.8). Additional details regarding the use and uncertainty of this probe configuration can be found in Kock am Brink and Foss (1993). 2.2.3.3 Multiple X-wire Configuration A four wire sensor was configured as two adjacent X-wire probes in a single probe body; see Figure 2.12. The same probe configuration was used in the measurements made by George and Hussein (1991), and by Lui (2001). This probe is similar to the streamwise vorticity probe discussed in the previous section, in that closely spaced X-wires are used to determine the lateral derivative of a velocity component which is normal to the stream- wise direction. In this case, the probe was used to estimate the mean squared value of dv/dz. This measurement was required for the dissipation estimate which uses the assumptions of axisymmetric turbulence; see Section 4.5. The four sensors shown were arranged such that wires 1 and 2 were processed as a single X-wire, with the output vl. Similarly. wires 3 and 4 are processed to provide v2. The desired velocity gradient is simply the difference between these, divided by the spatial separation, which was 3.0mm. With this probe separation, the electronic noise from the anemometers leads to an RMS of the velocity gradient of about 16 ( l/sec) in the free stream of the shear layer. The finite separation between the two X-wire probes can lead to an underestimate of the true velocity gradient. In the present flow field, the separation between the X-wires was of order 1511K, where “K represents the Kolmogorov length scale of the flow. Wallace and Foss (1995) review several references which discuss the attenuation due to finite probe scale. Figure 2 of Wallace and Foss (1995) indicates that there is reasonable agreement between different authors on this issue, and that the present configuration likely underesti— mates the true RMS of avflz by roughly 50%. However, the present data indicate that the performance of the probe is likely better than this estimate. Specifically, the variance of av/dz was found to be larger than the isotropic value, which is twice the variance of an; see Section 4.1. Given these already larger than expected values for the variance of £9sz , it is unlikely that significant attenuation has occurred in the measurements. SCFE’E’VWS \. __ entrance to test sectbn lllllll acoustt {fi’m-m J axml Fan a 35?: z‘ -. T_"-::i treatment, mgyxs .3331 \k \.. 3| ‘ a“.‘ T o s v .u .o. o- o .. II .0 .- a"; ._ u I. 1 I ,H .. .1 1: turnMg ,vanes =25 acoustk '5 treatment Figure 2.1 Primary flow delivery system 21 PRIMARY lNFLOW / 11111111 FLOW CONDITIONERS \" O 0 Z —-' :0 D O —{ 0 2 h 0: T 5.71m BOUNDARY L YER PLATE \ SEPARATION EDGE LI 111111111111111111111111111111111111 Q—- ENTRAINMENT H FANs ““r\~ mu 1 IKIXXIXIIIIIXXIIXXXXXIIIXXIIXIXXIIIXIXXXXIIXIIXIXIXIIIIXXIXIIIXXIIXIXIXIXXXIX Lxxxxxxxxxxx x x xxxxxxxxxxxxxxxxxxxxxxxxxulx x xxx xxxxxxxxxxxxxxxx 9.75M :— *‘ BYPASS HOLES\ I‘_ L i = 1 A; C: TUNNEL Exrr . Figure 2.2 Schematic of SSSL facility 22 / Copper Plating ilk/0951010 M/ // g.\ , Figure 2.3 Geometry of single hot-wire probe Cross—Pkkafiower stepper motor For probe angMacflfive \\ [fill/f ‘ / , ’//// motor For /\ C7 throtfle contro \\\\\ pflanar contractmn probe support Figure 2.4 Schematic of Calibration Facility 23 20 I j r 1 1 T Y I E2 18 r 1:.2=7.334+4.142v”44 j 16 L _ 14 C- — . 1 18 A l A l . l 1 l r .50 1.00 1.50 2.00 2.50 3.00 Vl'l Figure 2.5 Sample calibration and fit, 7:0 degrees. H0 ”11‘. 9:9:910XOXOXOX 1919:9192" { XOXO' 3:97.93 0 O f I 20202.202020939339323:X101917 Viva O 0 Figure 2.6 Geometry of the X-wire probe 24 30.0 3* . '5 25.0 ~ - g . § 20.0 — — 8 ~ . g 15.0 — J .D . 8 10.0 - a. . 5.0 — - A‘JAA‘gLLlAAAJLLA‘LAALLALLL 00 1 1 1 l 1 1 -40 -3O -20 -10 0 10 20 30 40 angle 00 Figure 2.7 Example of the locating algorithm of the intersection point for X-wire process- ing. 36 30 g- 24 an 12 cc 6 B 0 " q .2 -6 r j g -12 - _ A. -18 P _ -24 _— _ -3O mavzfievrwr; 1:»; A w T v A __ _ Aswww m ,1 __ -36 L 1 1 1 1 1 1 r 1 i 1.0 3.0 5.0 7.0 9.0 11,0 Velocity (m/s) Figure 2.8 Predicted values of angle (7) from re-processed calibration data at nominal val- ues of 18, 0, and -30 degrees. The standard deviation of these measurements was 0.18, 0.17, 0.21 degrees, respectively. 25 1+ rammii {03332222101033z~ {0101010103103 0 O 182:2:23232323232339;Wm» 0 O T 1 4 mm 13.032010202020203!0Y0Y0Y0Y0‘1'awe O O —-1- izgggggégiofo’l‘sSI!!!A A 1. .0103.) O 0 Figure 2.9 Schematic of spanwise vorticity probe tn-M t n-l t“ trwl JCrHM Figure 2.10 Schematic of micro-circulation used for the spanwise vorticity calculation. The velocities q, and q2 are known at the points represented by the dark circles on the top and bottom of the domain, respectively. The normal component of velocity, v, is known at the open circles at the left and right sides of the domain. 26 56mm ---1 lb— 1.0mm 4 O l“— 4.0mm '\ ,1 C C C C OKOZQ 29.9‘ ‘D .10. D V .0. O. A. 91.1. ' 02.1. z.. 3.3. v V O V v O. 1.91.. V v A A 6 v A 0'. '.V o. o A A A '01. Figure 2.11 Schematic of the streamwise vorticity probe. 27 Figure 2.12 Schematic of the multiple X-wire probe. 3.0 'lhrbulent Boundary Layer to Single Stream Shear Layer - The initial transition region. 3.1 Introduction A detailed investigation of the near separation region of the single stream shear layer was conducted. A schematic representation of the area of interest and coordinate system is shown in Figure 3.1. Measurements were acquired in the region 00) the approximation qzu is nearly exact. The time average and standard deviation of the velocity measurements will be denoted by a and 2'] respectively. The profiles 21(x= i 0.10”, y) are shown in Figure 3.2. Note that these data are normalized by the standard viscous units in log-linear coordinates. The wall shear stress was computed from the slope of a linear fit of the data points closest to the wall at x/00= -0. 1. The standard log-linear region can be observed through the range 405(0.25mm). The insert to the figure shows the deviation of the two traverses for y“<5 in linear coordinates since nega- tive y values were measured at the downstream location. The deviation between the two profiles indicates a strong streamwise gradient in velocity given that the x displacement of the sensor for the two traverses was 1.8mm. For example, Ail/Ax z 400( 1 /sec) at y=0. The integral statistics of interest, which were calculated from the mean profile, were the momentum thickness: 0:9.62mm, and the displacement thickness: 5*=12.6mm. The Rey- nolds number based on momentum thickness was Re9=4650. Additional parameters are 34 the friction coefficient and the shape factor, which are defined as C,- = 2(141/U”)2 and H=5*/ 0, respectively. These were determined to be C,~=0.00295 and H =1 .31. The 21(x= i 0.10", y) profiles in viscous units are shown in Figure 3.3. The observations that can be made from the profiles of 2'; are similar to those already mentioned for 2']. That is, differences between the two profiles can only be observed in the very near wall region. Note that the maximum value for the x/0=-0.1 data is (”l/uT = 2.74. Other researchers (for example Klewicki and Falco (1992)) have found a peak in the fluctuation intensity to be ~2.9 for similar Reynolds numbers. This discrepancy can be explained by the relatively large sensor length used in the present study. Specifically, the hot wire length in viscous units was 71+: 20v/uT, where Klewicki and Falco (1992) recommend a sensor length of less than lOv/uT in order to avoid spatial averaging of the velocity fluctuations. The spanwise vorticity in the boundary layer is also a variable of interest. The magnitude of the time averaged vorticity in viscous units is approximated by 111—: = ant/8y“ . The slope of the mean velocity profile shown in Figure 3.2 was computed; see Figure 3.4. The stan- dard deviations of the fluctuations in spanwise vorticity: 05:, were measured using the four-wire vorticity probe described in Chapter 2; these data are also shown on Figure 3.4. The measured values of r15Z are consistent with the data presented in Klewicki et a1. (1992) and references therein. These data show that both positive and negative values of vorticity were realized for y+>20. This can be compared with the near wall region wherein only one sign of vorticity that is consistent with that of the mean shear is observed. The vorticity 35 fluctuations normalized by the local mean value: (115/ jail have been plotted for the region 10800 because both the numerator and denominator are approaching zero. The data show that the relative fluctuation levels of vorticity increase approximately logarithmically (lin- ear in semi-log coordinates) throughout the “log region” (100. The visually apparent “growth” 37 of the shear layer is clearly at a length scale which is significantly smaller than the bound- ary layer thickness. Based on these observations, and equipped with the data to be presented in this and subse- quent sections, several distinct regions of the flow field have been identified. These are shown schematically in Figure 3.7. These regions and their defining characteristics will provide a conceptual framework and vocabulary which will be useful in describing the results to follow. The “canonical turbulent boundary layer” is the region which exhibits stochastic values of the velocity field that reflect the evolution of a boundary layer that is still attached to a flat plate. The “sub-shear layer” represents the turbulent region where the flow is measurably different from that which would be present if the bounding wall were present. The region marked “significant irrotarimialflzwtrmtions” represents the domain where measurements of the velocity fluctuations exhibit essentially no high fre— quency (small scale) motions. It is inferred that these motions are driven only by the unsteady pressure field created by the shear layer. Lastly, the transitions between sub- shear layer and the developing shear layer are not well defined. Hence no such boundaries are drawn in Figure 3.6. Point statistics were acquired using a single hot-wire probe which was traversed across the shear layer at 12 streamwise locations. The data were acquired at 5 kHz for 60 seconds. If the characteristic velocity is taken as U0/2, and if the shear layer width is taken to be 60(x), then the 60 seconds represents 500 to 2000 shear layer widths for the 12 locations. The sensor was aligned in the z direction to recover the time series of q as described in Section 3.1.2. Note that in the high speed free stream the sensor records q=u=U0, and in the entrainment stream the sensor records q=v=v(,. The time average velocity (21) and the 38 standard deviation of the fluctuations (6]) were then computed from the time series values. The mean velocity is shown in Figure 3.8 as an .r-_v plot, and in Figure 3.9 as a contour plot. The values of Z] are shown in Figures 3.10 and 3.1 1 in x-y and contour formats, respectively. A striking feature observed in these data is the streamwise invariance of the boundary layer statistics. As a specific example, the mean profiles are identical to the boundary layer profile for y/00>2 in the range 04 in the range 02 in the range 04 in the range 060 was found to be 0.035 as measured previously by Foss et al. (1977). Note that Hussain and Zaman (1985) found a linear growth rate of 0.032. The second variable derived from the point sta- tistics that will be of interest in subsequent sections is the maximum slope of the velocity profile. These data are shown in log-log format in Figure 3.14 for streamwise locations (1 .8120. This is inferred to be a result of the sub-shear layer phenomena where only the vorticity from the near wall region participates in an inflectional instability. Several statements can be made 42 to support this inference. The first observation is that the velocity gradients are very steep in the region x=y=0 which lead to length scales which are considerably smaller than the local integral scales. Since the length scale of the most amplified motions will scale with the velocity gradient at the point of inflection, it is reasonable that only the fluid in the y=0 region will participate in the instability. A second argument to support this idea is related to the vorticity field of the turbulent boundary layer upstream of separation. It is known that the filaments of vorticity in the very near wall region (say y+<1) are parallel to the wall, and are perpendicular to the wall shear stress vector. In contrast the outer region of the boundary layer does not have this constraint which leads to a vorticity field with a high level of disorganization. It can be inferred from these observations that the “organized" near wall vorticity participates in the first instability, whereas the “disorganized” outer motions do not. It is possible to calculate “how much ” of the time averaged vorticity participates in the first instability based on the above observations. Given the relatively small contribution of the y component of velocity to the vorticity in the near wall region. the mean streamwise velocity can be written as: L 17(y) = jwdy 3.3 0 The velocity scale of the first instability is observed to be 0.54U0 from Figure 3.14 near x=0. This corresponds to y+<42 at x=0 in terms of the mean velocity as observed from Figure 3.2. Evaluating equation (3.3) at y+<42 shows that 54% of the vorticity which sep- arates from the boundary layer participates in the first instability. The growth of the sub- 43 shear layer in the streamwise direction indicates that an increasing amount of vorticity from the boundary layer participates in the coherent motions as the flow evolves in the streamwise direction. This is also supported by the slow increase in UC to the nominal pre- dicted value of 0.5U(,.at large x/00 locations. 3.4.2 Spectral Properties of the Entrainment Stream The spectral properties of the near entrainment stream were used to investigate the coher- ent motions of the sub—shear layer region. The power spectral density (the Fourier trans- form of the autocorrelation) was computed for the 12 streamwise locations (identified in Section 3.2),. These data are shown in “log-log” format in Figure 3.17. The amplitude of each data set was shifted vertically in order to distinguish the data sets on one figure. These data show a distinct local maximum, or “bump” in the spectrum for all streamwise locations except x/0(,=0.73. For example, the hump at x/00=29.5 occurs at 10Hz. Nearly identical results for the range 371000, wherefjn~ l/xo'7| as measured in the present data, seems to correlate well with the H-2 data for x/0(,< l 000. It is possible that the present shear layer would also tend towards a 1/x dependence for larger x values given the agree— ment withe the H-Z data in the region where the data overlap. The non-dimensional frequency in the limit of small x values is also of interest, particu- larly with regards to the sub-shear layer described in the previous sections. In terms of equation 3.6, the second term dominates the magnitude off* for small x/00. This is because the frequency grows exponentially for small x, and the momentum thickness approaches the value at separation. A preferred scaling might be considered by relating the frequency to the local maximum velocity gradient: fjmlu) = fm/(du) . Note that (1 y m ax 46 this definition is equivalent to using the “vorticity thickness” scaling. These data are also shown on Figure 3.20. This definition was motivated by the fact that both variables follow a power law ~x'" as shown in Figures 3.14 and 3.19. In fact, the exponent of the maximum velocity gradient: n=0.79 closely matches the values of 0.71 measured for the frequency dependence. The gradient based definition provides a dimensionless frequency with con- siderably less streamwise dependence than f" for x/0(,<20. 3.4.3 The Initial Instability The frequency of the initial instability has been the focus of numerous studies (see, e.g., Ho and Huere (1984)). The understanding of how the boundary layer vorticity is re-dis— tributed downstream of separation is of considerable interest both fundamentally and tech- nologically. Manipulation and periodic forcing of the boundary layer at separation can take particular advantage of the first instability in order to control the unsteady dynamics of the shear layer. The near separation region of a laminar shear layer is known to be unstable to small ampli- tude perturbations due to Kelvin-Helmholtz instability. Linear stability theory has been used by many authors to understand and predict the frequency and growth rate of the first instability; (see, e.g., Ho and Huere (1984)). There are several important assumptions of the linear theory which often include parallel flow, laminar flow with infinitesimal distur- bances, and a specific velocity profile such as a hyperbolic-tan gent function. Although these assumptions are not physically realized in any shear flow, the theory predicts the fre- quency of the fundamental instability for various geometries with great success. The effi- cacy of this theory for shear layers and jets was reviewed by Ho and Huere (1984) and 47 Thomas (1991). The Strouhal number of the most amplified frequency is defined as St 0 = me / U 0. Note that some references use the average velocity in the case of two () stream shear layer. This would change the definition by a factor of two; the referenced val- ues have been divided by two where appropriate for comparison to single stream results. In contrast, the effects of laminar-turbulent boundary layer states and Reynolds number on the initial instability have not been previously been extensively investigated. Although the initial development region for laminar flow has been studied extensively, most shear layer studies which have a turbulent boundary layer do not consider the initial instability of the shear layer. Ho and Huere (1984) comment that the most amplified Strouhal number changes from 0.016 for a laminar flow to 0022-0024 for turbulent flow due to “presently unexplained reasons.” The remainder of this section will explain their observation in terms of the sub-shear layer phenomenon presented in Section 3.2. The single sensor hot-wire data described in the previous sections was used to observe the frequency of the initial instability. Using the spectral information shown in Figure 3.17 it can be observed that the time-averaged spectra of velocity shows a local maximum at the streamwise locations x/0021.87. Although these time series are not periodic, the local maximum is interpreted to represent the preferred frequency at that location. However, the lack of a local maximum in the spectra for x/0(,=0.73 does not imply that a preferred fre— quency does not exist. The flow visualization shown in Figure 3.5 indicates that the first instability does occur in this region. This has motivated a more detailed examination of the time series data at this location to identify if a preferred frequency exists at this location. 48 Several segments of time series data are shown in Figure 3.21. These data show that pseudo-periodic motions do exist, although the frequency, phase and amplitude vary con- siderably during a relatively short time period. This explains why the spectra failed to show a local maximum value at a particular frequency. This result is also physically rea- sonable given that the vorticity, which is convected from the viscous region of the bound- ary layer, is highly unsteady and highly perturbed by the turbulence in the “log region”. An alternative method of determining the frequency content of a signal is to look for the time delay of the first local maximum of the autocorrelation. The inverse of this time will indicate the frequency. The autocorrelation of the entire signal, along with several short time autocorrelations is shown in Figure 3.22. The time series data were divided into 0.1 and 0.23ec segments for analysis. The autocorrelation was then calculated for each of these time segments. A distinct local maximum was found in the autocorrelation function for 51% of the 0.1sec data and for 37% of the 0.2sec data. A histogram of the observed frequencies is shown in Figure 3.23. The mean frequency from the 0.1sec samples was 130Hz. The mean value from the 0.2sec samples was 146Hz. This value corresponds well with fm=124 which is calculated from equation 3.4. This agreement adds confidence that a preferred frequency does exist for the initial instability, and that its value can be accurately determined. The next step in the determination of the initial Strouhal number is to choose velocity and length scales to non-dimensionalize the indicated frequency. The standard choice of U0 and 00 would not be appropriate since it is only the inner most wall region that is partici- pating in this first instability. It seems illogical to use. for example, the boundary layer 49 momentum thickness (00) to scale a phenomenon which does not involve the entire boundary layer. The new scales of velocity and length will be U0” and Befito imply that it is the effective velocity scale and only the effective momentum thickness that lead to the first instability and roll-up. A logical choice for the velocity scale is Ucff =0.54U0. This is justified by the apparent convection velocity of the first instability discussed in Section 3.3.1. The effective momentum thickness of the sub—shear layer was determined by fitting the hyperbolic tan- gent profile to the mean flow measurements acquired at x/00=0.73. The data were fit to the equation: QO’) _ if; ( .v U0 — 2 [I '1' tanh Z—GJ-D] 3.7 in the range 02 is stochastically identical to the boundary layer. In order to directly contrast this large spanwise coherence with the boundary layer flow, the correlation coefficient for these four locations as well as one position within the boundary layer (at x=0, y=0.500) are shown in Figure 3.30. It is clear from these measure- ments that the larger spanwise coherence is developed downstream of the separation point, and within the sub-shear layer region. The length scale of the correlation measurements is clearly many times greater in the region near the sub-shear layer compared with the boundary layer, even at small x/00 values. The continued growth in coherence is also seen in Figures 3.28 and 3.29, where the irrota- tional fluid motion is beginning to appear nearly two dimensional. These measurements closely parallel those of Bowand and Trout (1980, 1985, collectively referred to as BT) where 12 wires were placed in the low speed side of a two stream shear layer facility. They also found the correlation field to increase in the streamwise direction. The short time real- izations of the velocity fluctuations shown in BT look very similar to those shown in Fig- ure 3.29. 3.5.2 Correlation Coefficient Data The correlation coefficient R(Az) was calculated (as shown in Figure 3.30) for each of the (x,y) locations (see Figure 3.9 for the grid of data points used). The data set can be repre- sented by the “volume” of points (x,y,Az). The visualization of these data was made possi- ble by “slicing” the volume in planes perpendicular to the x and 2 directions. First, the streamwise locations x/00=3.53, 12.0, 40.5, and 84.2 (i.e., the positions used in Figures 3.26-3.29) were selected. Figures 3.31 through 3.34 show contours of the correlation coef- ficient at these locations. The mean velocity profile is included with each of the figures for reference. Figures 3.35 through 3.39 show contours of R at specific Az values. The solid boundary surrounding the contour plot represents the spatial extent where the data were acquired. The dashed boundary represents the contour line where the fluctuation levels are 1% of the free stream velocity (see Figure 3.10). Correlation data are shown in the region where the fluctuations exceed 1% of U0. Several important features of the flow field can be observed from these measurements. First, the central region of the shear layer shows quite low spanwise correlation values. This is in contrast to the low speed irrotational flow measurements which exhibit a large correlation length with respect to the local length scales. Although it is apparent that the irrotational fluctuations are caused by the motions existing in the turbulent fluid, the corre- lation measurements in the central region of the shear layer are “washed out” by the high levels of small scale turbulence. A second feature observed from these data is the difference between the high and low speed sides of the shear layer. Specifically, the correlation length grows smoothly and con- tinuously from the separation point, whereas the high speed side shows a discontinuity in the data near x/0(,=55. These data are consistent with the concept of the sub-shear layer, and also provide insight as to how the sub-shear layer evolves into a fully developed shear layer. Specifically, the vortical fluid which occupied the outer region of the boundary layer does not participate in the instability which leads to the large scale spanwise motions. As the flow progresses in the streamwise direction. more of this vortical fluid is drawn in to “participate” with the coherent motions. The dramatic change in the correlation field near x/00=55 indicates that the motions have become strong enough to have a measurable effect on the high speed irrotational fluid. The time averaged convection velocity of the coherent motions at this location is approximately 82% of the final value of 0.5U0. 3.6 Summary and Conclusions A number of observations can be drawn from the collection of data described in the pre- ceding sections. These data have revealed information which does not currently exist in the fluid mechanics literature. As stated in the introduction, the unique features of the cur- rent facility that have allowed these discoveries are the high Reynolds number at separa— tion (4650) and large length scale (00:9.6mm) of the flow. For comparison, Hussain and Zaman (1985) studied a similar geometry, albeit the boundary layer Reynolds number at separation was 428, and the momentum thickness 00:0.5mm. Many of the features described in this chapter have been at scales of the order of 00 or smaller which would make them unnoticeable in a facility with a small value of 00. In the following, the main conclusions are given in a numbered format followed by a discussion and interpretation of the relevant data sets. 1. The stochastic properties of the boundary layer at separation (x=0) appear to be unaf- fected by the flow field downstream of separation. This conclusion is supported by several of the experiments. First, the zero cross-correla- tion magnitudes between velocities measured in the boundary layer and in the entrainment stream is a strong indication that a causal relationship does not exist between the two regions. The lack of correlation in the low-pass filtered time series further supports this result. The reported stochastic properties of the boundary layer show that the moments of the velocity and vorticity field are in good agreement with the existing literature on boundary layers without separation. If the shear layer did have an effect on the boundary layer at separation, it was not apparent in any of the point statistics or correlation measure- ments. The streamwise invariance of the statistics in the outer region of the boundary layer pro- vides additional evidence to support this conclusion. In other words, not only do the shear layer fluctuations not effect the boundary layer at x=0, there is no measurable effect on much of the y>0 region for a considerable length downstream. This will be discussed fur- ther as conclusion number (2) below. The fact that a high Reynolds number boundary layer which separates at a 90 degree edge is not affected by the separation is important for several reasons. If the separation region 56 was to be modeled using a computational method. the upstream boundary conditions could be completely specified at x=0 by the stochastic properties of the approach bound- ary layer. If in contrast, the separated flow field were to effect the stochastic properties upstream of separation, then the inlet conditions to the computation would have to be specified at some x<0, rather than at x=0. It is likely that the boundary layer to shear layer communication is quite dependent on the Reynolds number. Laminar and low Reynolds number turbulent flows at separation are likely to be more sensitive to the perturbations caused by the downstream motions. A future study of this topic may attempt to measure an objective indication of the communi- cation for a range of Reynolds numbers. 2. The outer region of the boundary layer is stochastically streamwise invariant for sev- eral integral lengths downstream of separation. This streamwise invariance of the boundary layer properties is quite evident in all of the data identified above. For example, the contour plots showing statistics in the x-y plane (Figures 3.9, 3.1 l, 3.35-3.39) all show invariant properties of the flow from in the region 00'07 on Figure 3.1 1. Many of the salient features that are associated with canonical shear layers have been identified in this region. It can also be observed from the u /U0>0.07 contour on Figure 3.1 1 that the “spread angle” of the sub- rms shear layer and of the fully developed shear layer are the same. An important feature of shear layers is the presence of coherent motions which develop from the inflectional instability mechanism (see Ho and Huere (1984 for a review). These motions are known to move in the streamwise direction with a well defined convection velocity. In fully developed single stream shear layers, this velocity is known to be approximately 2anv/Uo=l .0. The convection velocity in the sub-shear layer region is shown in Figure 3.16. These data indicate a dramatic rise in the convection velocity from 20.54 at separation to 1.0 near x/00=120. It is inferred from these data that only the near wall vorticity is “rolling up” or participating in the inflectional instability. A quantitative estimate of the fluid domain which participates in the initial roll-up has been identified as that below the q/ U0 2 0.54 isotach. Equivalently, this bound is given by y+<42 for the present boundary layer. As the sub-shear layer grows into the outer part of the boundary layer, more of the vorticity is pulled into the coherent motions thereby increasing both the width of the sub-shear layer and the convection velocity. The present understanding of the vorticity field within a turbulent boundary layer is con- sistent with the current findings. That is, the stochastic and instantaneous properties of the motions which define the inner and outer scales of the boundary layer are significantly dif- ferent. The near wall motions are described by, for example Adrian et al. (2000), and Pan- 58 ton ( 1999). In these references, the near wall region is understood to be comprised of groups of coherent motions which define localized shear layers. These result in “vortex packets” which evolve into the outer scales of motion. It is intuitively reasonable that the near wall region with a high level of organization of vorticty would be subject to inflec- tional instability immediately downstream of separation. In contrast, the motions farther from the wall are usually considered to be more “disorganized” and not closely correlated with the motions very near to the wall. The current measurements of spanwise vorticity fluctuations at x=0 support this viewpoint. Again, it is intuitively reasonable that these motions would not participate in the roll-up phenomenon. 4. The recognition of the sub-shear layer allows a physical explanation for why shear layers with a turbulent boundary layer at separation approach self-similarity at smaller x/00 values compared with shear layers with a laminar separation. The streamwise location where the flow field becomes self-similar has been found to be quite dependant on the state and Reynolds number of the boundary layer at separation. For example, Bradshaw (1966) recommends a value of x/0(,>1000 for self similarity given a laminar separation. Hussain and Zaman (1985), Bruns (1990), and the present study have found that x/00>100 is sufficient for the mean velocity to exhibit self—preservation when the separating boundary layer is turbulent. These differences can be explained in terms of the length scales of the turbulent vs. lami- nar boundary layers. The suggestion was given by Dimotakis and Brown (1976) that it is not the downstream location as defined by a number of original integral length scales (e.g., 00) that is important for the establishment of a self-similar condition. Rather, it is the num- 59 ber of “interactions” between the large scale motions of the flow. The number of interac- tions will depend on the ratio of the streamwise location to the spacing of the initial disturbance. Specifically, the number of interactions mlx) is estimated (equation 14 of Dimotakis and Brown) to be m(x) = log2(x/ln) 3.8 where 10 is the spacing of the initial disturbance. For a laminar boundary layer at separa- tion, the criteria x/0(,>1000 equivalent to a value of m==4. 1n the present study the stream- wise location x/0(,=100 corresponds to a value of m between 6 and 7, given that the spacing of the initial disturbance was [0:00. This small value of initial disturbance length compared with 00 is a direct result of the presence of the sub-shear layer as described in Section 3.3. 5. The dimensionless frequency of the initial instability agrees with linear stability theory when appropriate effective velocity and length scales are identified. The measurement and prediction of the dimensionless frequency of the initial stability has received considerable attention in the literature: see. for example, Ho and Huerre (1984) and Thomas (1991). This focus is a result of the universal nature of the initial instability in a variety of shear layer and jet flow fields. Additionally, linear theory has been quite suc- cessful in the prediction of this dimensionless frequency for many experiments with lami- nar boundary layers at separation. However, differences in the measured frequencies have been found in shear layers with turbulent boundary layers at separation. The present study has shown that the above noted differences are a result of the different length and velocity scales that exist in a turbulent boundary layer. Specifically, a laminar 60 boundary layer can be fully characterized by the free stream velocity and momentum thickness. In contrast, a turbulent boundary layer has outer scales (free stream and momentum thickness) and inner scales (uI and War). The data presented suggest that only the near wall vortical fluid participates in the initial instability at this Reynolds number. It is then quite reasonable that the outer scales would not be functionally related to the mea- sured frequencies. A hypothesis that can be inferred from the above statements is that the Strouhal number will properly collapse using inner scaling. From the present data, this suggests that Utflz l 514,, and 0,,fi~10v/ut. Additional measurements are required to test this hypothesis. 6. Preferred frequency and spanwise correlation measurements have shown that the char- acteristic large scale motions observed in self-similar single and two-stream shear lay- ers are present in the sub-shear layer region. Several investigators have studied the properties and importance of the largest scales of motion in shear layers. These motions appear to have a large spanwise coherence and a significant fraction of the total turbulent kinetic energy of the flow. Experimentalists have sought to measure these features in order to better understand their role and dynamic sig- nificance; see, for example, Hussain and Zaman (1985), Browand and Weidman (1976), and Browand and Trout ( 1980, 1985). The present experiments have found that a number of the features measured in both two stream and single stream shear layers in the self-similar regions have also been observed in the sub-shear layer region. For example, the preferred frequency exhibited by both the sub-shear layer and self-similar shear layer can be described by the relatively simple equa- 61 tion: f,,,~x’”'7’ (for x/0(,>l ) as shown in Figure 3.19. 11 can be inferred that the physical mechanisms which lead to this preferred frequency are present throughout the flow field downstream of separation. Similar results were found from streamwise dependence of the maximum slope of the mean velocity profile; see Figure 3.14. Note that the y location of the maximum slope was found in the center of the sub-shear layer. A third feature, observed in the sub-shear layer region. was the large spanwise correlation length; see Figure 3.31. These data clearly show that large scale motions develop at small x/0o values. The spanwise correlation of the velocities in the entrainment stream increases as these motions convect downstream. These observations are consistent with observa- tions made by Browand and Trout (1980, 1985) in the self-similar region of a two stream shearlayen The above noted observations justify the terminology “sub-shear layer.” Specifically, the region exhibits characteristics which are associated with self-similar shear layers, while the prefix “sub” is meant to describe the fact that the shear layer is growing adjacent to what appears to be the outer region of a flat plate boundary layer. One distinct difference noted between the sub-shear layer and the self similar shear layer is the peak value of a which increases smoothly from 0.12U0 in the boundary layer to a constant value of nomi- nally 0.16U0 for x/0(,>150. 62 , I . I . 41/. \ \\ 3\\ \ \ \ k\ \\i U,;,=7.|M/S T 5=|OOMM y “I: / 1 . g——‘x e=9.6MM FT? zx/uT 20.05MM ' arrrrttrrrrtrr VE(Y=-°<9=O.055UO H—9.75M——1/——-— Figure 3.1 Schematic representation of near separation region. 30 4+ ' n00: +0.1 25 — g r . #90: -0.1 20 e l u+=2.31n(y +)+6.7 '5 ” u+=2.51n(y +)+5.5 / \ I, \ / + “ \‘\ + U=y Figure 3.2 Boundary Layer Mean Velocity. The y+=0 point (x) on the inset figure was added based upon the no-slip condition for clarity. Note that U0/u1=25.51 /, 5.0 (1.1) r 1 * ‘ -l(l.(l -5.0 101) 15.0 20.0 25.0 30.0 0.0 1 la) +I 071 1 V“ + 1 1 L A._._A A ' L A 1 A l 10 100 1000 y+ Figure 3.3 Boundary Layer Velocity RMS 0.7 "'0— Mean value from slt)pc of sin gle sensor traverse 0.6 ~ "Error Bars" denote +/- 1 standard deviation of vorticity calculated from the four wire vorticity probe time series. 0.5 ’ 04 . Vorticity (RMS/Mean) 0.3 » //A'""" 0.2 ~ -1» " .11 g 1 . '0” 2109 ().l * 0.0 r -01 e A 1+- 1 A A . A .1 10 100 1000 y+ Figure 3.4 Mean and RMS of vorticity at separation. 64 2 . (pman/un 100 3 Model Spectrum _______ Measured Spectrum ' 00 1.0 8.0 3.0 4,0 50 Figure 3.6 Sample flow visualization of the near separation region. 65 FFEESTREAM A... ———> . « ' ‘ W'— ‘ A, u I c l I 1. n..." .W_»’,.. I T LAYER M" DEVBJPING SEAR | AYE? a «0 . -.. a . ”“1541. Van‘- . 1.11.1.V1w (4.7.7.373 :: 7:71.71 77-513.; 2 1,4. 5.-.: .3234}? - l : \\ . I; ”Ir-u NW1‘1 4 ““32:- ’i I .1 ..... ,_ SIGNIFICANT l ‘ ‘k . E 1 51' S‘IREIIIWN”‘--- lmorAmNAL \. Figure 3.7 Figure of flow regions; not to scale. 1.0 ”90 ——<>— 101 (Two —A— 114.4 0.8 1 “—0— 63-3 —g»— 54.2 —>(—- 40.6 0‘6 i ——+_— 29.6 ——+-— 19.3 0.4 + ”‘0 + 7.19 + 3.54 0.2» + 1.88 + 0.13 —-.—— 0.1 0.025 Figure 3.8 Mean velocity profiles for 0 F l 1 L 0.0114111111211112 0 0 20 40 ()0 80 100 x/OO Figure 3.20 Non-dimensional frequency vs. x/GO 72 Fig We. _2 1 l 1 l 1 l 1 1 0.00 0.02 0.04 0.06 0.08 . 0.l0 “rm Figure 3.21 Time series samples from x/80=O.73. 1.01 . . 2 0'8 " Time averaged _ 0.4 L 0.2 1— 0.0 b -0.2 1 . 1 1 1 1 0.000 0.005 0.010 0.015 0.020 time delay: I (sec) Figure 3.22 Autocorrelations of the time series at x/90=0.73. The long time averaged as well as 0.1 second averages from the time series in Figure 3.21 are shown. 73 0.15 E 0.1secsampks I 0.2 sec samples 0.10 1~ ’fi \‘V‘l'f' ;ra Q A . - .11. -.- ,‘ .. . _. . ' _ .-4 13"“ ‘Hz‘. ‘ '.- . . ~ ' " a -2 0.05 ~ fraction of realizations I IICICCIIIICUCI - -.--.---. . ‘I 'I ‘ ‘- ..|&‘ Q" IIIIIIIIIIIIIIIIIIIIIIIIII! ______--______ r.n_.l .“v.h .~ IlllllIlIIIlIIIIIIIIl 0.00 * 60 80 I t—ul III-IIII-I-I-I-I-II-I-C. 00 120 140 160 180 200 220 frequency of local maximum Figure 3.23 Histogram of observed frequencies from short time autocorrelations. 0.7 . _ 0.6 E " q/Uo T 0.5 1— 0.4 U .. ; K - ‘“ [1+tanh(y/2 Gem] 0.3 .‘." 2 0.2 Uc,1,1/U0=0.56 0.1 ; Gen/90:0.OSZ 00 h. II 1 1 1 1 l 1 l 1 -().6 -04 -02 0.0 l 0.2 0.4 0.0 0.8 1.0 Figure 3.24 Comparison between tanh function shown and the velocity data at x/00=0.73. 74 :21 :: L_J :2 1:1 1 IACM :1 :2 3:1 Figure 3.25 Schematic of 8 wire rake in shear layer. 75 3 2190:205 ec=0.02 W “ 2190:1221 Figure 3.26 Velocity fluctuations at x/60=3.53, y/90=-3.7. Contour plot shows magnitude of velocity from 8 sensors normalized by the standard deviation. X-Y plot shows time traces taken from wires located at z/Bo=2.05 and 12.4 for the same time period. Note that the correlation coefficient indicated in the top figure by ‘cc’ represents the long time aver- age and not that for the time segment shown. 76 cc=0.23 \ 1} J V \ \ \7190=12.4 o-Arouammumco Figure 3.27 Velocity fluctuations at x/00=12.0, y/60=-5.37. See Figure 3.26 for descrip- tion. 77 1: up 1 - .31.“ .11 o 8 7100 NQJIUImNQID .1 O Figure 3.28 Velocity fluctuations at x/00=40.5, y/00=-10.9. See Figure 3.26 for descrip- tion. 78 7100:105 ’2/90=12,4 CC=O.89 -- Ga) 4.50 0,00 150 3.0) 2/80 O-‘NO’&UIO’)\IQO I Figure 3.29 Velocity fluctuations at x/00=84.2, y/60=- 19.6. See Figure 3.26 for descrip- tion. 79 .\: 0.8 » ' - (1090:1142, y002-19.6l 0.6 »» - . (x190=4().5, y00=-1().9l 0.4 L ' ' . /(Xflo=l2.0, 300:5.7) 0.2 ~ ' . (xz‘Bo=3.53. )0(,=-3.7) ' . 00 L N‘F— ' -_ 1 \ (x=0. y10(,=().5) -02 1 g 1 - . 1 1 O 5 1() 15 20 Az/GO Figure 3.30 Spanwise correlation coefficient of the data shown in Figures 3.26 through 3.29. 80 al./U0 1.00 . 0.50 0.00 ' _L —L —L —L .15 O -‘ N (A) «5 f f 1 r 041003-0103me Figure 3.31 Spanwise correlation field at x/00=3.53. 81 1.0 q/Uo 0.5 l 0.0 ‘ 1 1 1 . 1 1 4.1.14.1 O-‘NCO-h O-‘NOD-hO‘lODVCDCD Figure 3.32 Spanwise correlation field at x/00=12.0 82 1.00 r q/Uo 0.50 r 1 141 M6013 12— 11 L 10» 91— 8» \ 61 5- \ 41 31 ¥JTXX 2 ~ _; 1 _ O111111111111111111111111111 -1O -5 0 1O 15 5 y/Oo Figure 3.33 Spanwise correlation field at x/90=40.6 83 9~ \‘ 8. B \007 7c 6_ 4. 3’ ~~-'~—-———-0.33 2» --~047 Mom 1 ' 073 Ollllll1llllllllllllilll14111111111 -20 -15 -10 -5 O 5 10 V90 Figure 3.34 Spanwise correlation field at x/00=87.2 84 y/Oo Figure 3.35 Correlation coefficient at A2190=2.08. Note that the solid boundary repre- sents the spatial extent of the measurements (shown by the grid). The dashed boundary represents the region where urmslU0>0.Ol. A schematic of the boundary layer mean pro- file near x=y=0 is shown for reference. y/Bo X/00 85 Figure 3.36 Correlation coefficient at Az/60=4.l6. See Figure 3.31 for description. ___- 20 - fr --- 1 —>--—---"""‘""" ,x'" o.95-—-« aw --- ----....._.-§-__ """" W032 W006 \ A 0.19 X /60 Figure 3.37 Correlation coefficient at Az/60=6.24. See Figure 3.31 for description. y/Go '20 ~ 0.92 X/Oo Figure 3.38 Correlation coefficient at Az/9(,=8.32. See Figure 3.31 for description. 86 y/GO -20 x /00 Figure 3.39 Correlation coefficient at Az/9(,=l4.5. See Figure 3.31 for description. 87 4.0 The Developed Single Stream Shear Layer 4.1 Introduction This chapter presents the results from an investigation of the developed region of the shear layer. The word “developed” is used to indicate that the time averaged flow variables can be represented in a self-similar form. Specifically, any dimensionless stochastic quantity, say, a(x,y), can be written as a function of the dimensionless transverse coordinate: a((y- yC)/l), where yC represents the “center” of the shear layer, and [(x) is a local integral length scale of the shear layer. In the present measurements, the momentum thickness: 0(x) was used as the length scale of choice. In shear layers, this length scale grows linearly. That is, 0~x. This was shown to be valid in the measurements presented in Chapter 3 for x/00>6O. These data also indicated that the profile of the fluctuation intensity (see Figure 3.15) is approximately self-similar at streamwise locations x/9(,>300. This is consistent with the results of Hussain and Zedan (1978) and Bruns et al. (1991). The data presented in this chapter were acquired at downstream locations 384 O) of the entrainment fluid. Also, gm) < 0 is observed for l()() n > —0.38 . This indicates a net flow of fluid, which originated on the high speed side, towards the center of the shear layer. A consistent but different view of the flow field can be understood from the tilted coordi— nates. An interesting feature of the g(fi) profile is the positive lateral mean velocity com- ponent observed for all f] . Thus, to an observer aligned with the centroid of vorticity, fluid is observed to move from the low speed side to the high speed side of the shear layer. In this view, the shear layer entrains and “extrains” fluid at the same rate, as already observed from the boundary conditions (equation 4.22). Although the ‘tilting’ of shear layers with respect to the free stream flow has been recog- nized throughout the literature, the failure to recognize the non—zero value of or in the for- mulation of the self-similar equations has resulted in incorrect statements and conclusions in the literature. For example, Holmes et al. (1996) derive the self-similar equations for a single stream shear layer. Because the movement of the shear layer into the low speed side is not explicitly accounted for, the equations imply (1:0. That is, the equations are derived in a tilted coordinate system. A contradiction is then encountered (see page 54 of Berkooz et al.) in which the statement is made that g(fi= 00) = 0. The argument is made that “extrainment is never physically observed.” The resulting analysis (e.g., equation 2.23 of Holmes et al., 1996) is therefore incorrect. Note that geometrical reasoning (based on equation 4.9) argues that the isotach aligned with the streamwise direction occurs at f] = l . Note however, that this istotach is not a streamline. In the untilted coordinates, a small negative value of \7 is observed at 11:1, indicating the movement of fluid towards the low speed side. The direct experimental 101 measurements in room coordinates show that the mean velocities which are acquired at constant y=0 location result in 51/ U0 = 0.685 . corresponding to n=0.90. This small dif- ference between the predicted value of n=1 and the determined value of n=0.9 is likely a result of the growth of the displacement thickness in the boundary layer of the outer (y=lm) wind tunnel wall. 4.2.2.2 Profiles of the Fluctuating Velocity The probability distribution function (pdf) of the unsteady fluctuations of the three com- ponents of velocity can be characterized by profiles of the moments of the pdf. In this sec- tion the root mean square (RMS) of the fluctuations are considered. Higher order moments of the velocity field in a single stream shear layer are examined in detail in several refer- ences. See, for example, Wygnanski and Fielder (1970) or Champagne et al. (1976). The spatio-temporal characteristics of the velocity field will be considered in Section 4.6. The RMS of the streamwise velocity fluctuations (17) was calculated from the time series data acquired at x/00=384 and 675. These data are shown in Figure 4.6. Similarly, the RMS of the lateral (v) and spanwise (w) fluctuating components of velocity are shown in Figures 4.7 and 4.8 respectively. These measurements agree well with the measurements published in the literature. For example, the maximum value of fl/Un was reported to be 0.168 and 0.176, and 0.165 by Champagne et al. ( 1976), Wygnanski and Fiedler (1970), and Bruns et al. (1991) respectively. The present data indicate a value of 17/ U0 = 0.17 at both locations: x/00= 384 and 675. Similarly, the maximum value of \7/U0 was found to be 0.12 by Wygnanski and Fiedler (1970), and 0. 135 by both Champagne et al. (1976) and Bruns et a1. (1991). The present data indicate a value of 0.12 at x/0(,=384, and 0.13 at x/ 102 00:675. The general agreement of these statistics supports the inference that the facility provides a canonical single stream shear layer as studied in these other references, and that a self-similar state of the flow field was achieved. 4.2.3 Profiles of the velocity gradient variance The determination of the velocity gradient moments: (811/ 8x1.)2 is of interest. For exam- ple, the description of the flow field in terms of the vorticity field (Section 4.2) or the dis- sipation of kinetic energy (Section 4.4) will make explicit use of the measured values and their relations. Secondly, the velocity gradients represent information about the smallest scales of motion. The relationships between the various derivative moments can be used to validate assumptions about the small scale motions. A complete description of these moments requires the measurement of nine independent gradients, written in matrix form as: Bu Bu Bu 5? 5 a7. 01,. = £1" 91 ‘1’ 4.23 9") Bx Ely Bz a_w 21» a_w -ax By 821 In the present work, only six of these, highlighted in bold face in equation 4.23, were mea- sured. This choice of measured quantities is partially due to experimental limitations, and partly as a result of interest in the theories of locally isotropic and locally axisymmetric turbulence. These hypotheses require specific relationships among these terms, which can be validated with these measurements. 103 The measurements were obtained as follows. The streamwise derivatives were calculated using the Taylor microscale formulation (see, e. g., Pope (2000)). For example: ~2 (Bu 2 u _ = . .24 91- 2 712 4 where 2» is the Taylor microscale calculated from the autocorrelation function. The stream- wise derivatives of v and w were calculated similarly. These approximations do, of coarse make use of the Taylor’s “frozen turbulence” assumption. In regions where the fluctuation levels are not small compared to the mean velocity, this assumption can lead to significant errors. Lumley (1965). and later Wyn gaard and Clifford (1977) derived corrections for the longitudinal and transverse derivative variances: 2 m 1’ a s 11 red 2 2 2 1 .' +2' +2 ' 11 _ I,” V W l: c 4.25 [(a '/a )zlcorrected {(2 u u x and m ca 8' 11 red _.—2 T —1_ [(avi/ax)2] - U ‘ = I+(H + V2/2-l-W2) : C 4.26 [(av'/ax)2](fim I ('r (a! (:22 v These correction factors are plotted across the shear layer at the two streamwise locations x/00=384 and 675 in Figure 4.9. Since the individual terms on the right hand side of equa- tions 4.25 and 4.26 have been shown to follow self-similar scaling, it is expected that these dimensionless corrections must also. This is found to be approximately the case. However, the small differences in the measurements on the low speed side are squared and added, leading to more noticeable deviations between the two streamwise locations. It is of inter- est to note that the magnitude of the correction in the low speed side is considerable, given the high levels of turbulence intensity in that region. 104 It is expected that the velocity gradient moments, along with the statistics previously described, should follow self-similar scaling. However. because the velocity gradients are related to the small scale motions of the flow, care is needed in the way in which the veloc- ity gradients are normalized. The dimensions of (Bu 1/ (it!)2 are (1/time2). Thus, a time scale of the flow must be used to normalize these statistics. The integral time scale 150:0(x)/U0, will not properly scale the gradient statistics because to represents the largest time scales of the flow. The proper scaling is obtained from the Kolmogorov time scale since the spatial gradient of velocity is related to the small scale motions: T In = f—K L2 = ” 4.27 ) where 8 is the rate of dissipation of kinetic energy. The use of dissipation to scale the velocity gradient statistics becomes obvious given the definition of dissipation in a homo- geneous flow: Sal-Bu,- 8 : V —— , 4.28 Bari-ax,- Equation 4.27 relates the dissipation to the mean flow variables from the relation 3 . . . . 8 ~ U0/0. This scaling can be argued based on the d1mens1onless form of the turbulent kinetic energy budget; see Section 4.5. Note that the ratio of the integral to dissipation time scale is proportional to the square root of the local Reynolds number. The velocity gradient statistics will be normalized by multiplying the measured values by 1.12; this will be referred to as dissipation scaling. The measured values of (ad/ax)2 for at the locations x/00=384 and 675 are shown in Figure 4.10 normalized by dissipation scaling. The corrected and uncorrected symbols 105 refer to the use of equation 4.25. Similarly, the terms (OW/8202 and (SW/Bx)2 were cal- culated from their respective autocorrelation functions, and are shown in Figures 4.11 and 4.12 respectively. The data indicate that the profiles of the velocity gradients are similar in shape to the fluctuating velocity components, although the data for n<-l do not exhibit a monotonic decrease towards the low speed side. It is assumed that these measurements could have significant error due to Taylor’s hypothesis that are not easily corrected by equations 4.25 and 4.26. It is also clear that the use of dissipation scaling properly scales the data (see n>1) from the two streamwise locations into self—similar form. The terms: (art/By)2 and (aw/am were obtained using the parallel sensors of the Mitchell probe with 1.4mm spacing; see Figures 4.13 and 4.14. Lastly, the term (avVaz)2 was obtained using the double X wire probe described in Section 2.2.3.3. These data are shown in Figure 4.15. Because these measurements do not require the use of Taylor’s hypothesis, the measurements show a more symmetric shape to the profile. In addition to providing information about the vorticity and dissipation fields, as will be discussed in later sections, the variances of the velocity gradients can also be used to test the symmetry of the turbulence properties. For example, if local isotropy is assumed, then only the term (Bu/8x)2 is independent and need be measured. This theory is reviewed in detail by, for example, Monin and Yaglom (1975). The other eight terms are related by: (8.1-0: 1%)2 = (3502’ 4-29 1—0 1—11-11—11—11—t-—- 2 106 The assumption of axisymmetric turbulence is less restrictive than the isotropic assump- tion, as originally put forth by Batchelor (1946). The set of relations between the deriva- tives are given by: Ix) (3‘92 = (851) , 4.31 19.-3”: 2 = @153"... 1... 1%,”: 2 = £893.01. 4-35 These relationships were derived and discussed by George and Hussein (1991). In order to test the validity of these assumptions, Browne et al. ( 1987) used the ratios of the deriva- tives: (8.4.22 1%)2 1%- (2+5 ——,K2 = 2 ,K, = ——,K4 = 4.36 (3.32 (9.02 i (9.02 1%le If the flow field is truly isotropic, then these ratios should be equal to unity. The axisym- metric assumption is less restrictive in that no prescribed values are required for the con- stants. The axisymmetric turbulence condition requires Klsz, and K3=K4. The ratios of equation 4.36 are plotted in shear layer coordinates for the two streamwise locations in Figures 4.16 and 4.17. Note that the values with the Taylor correction factors were used. These ratios do not clearly support either the isotropic or axisymmetric assumptions. However, it could be argued that the values of K 3 and K4 are closer to being equal than 107 they are to being unity. The same is true for K, and K2 over much of the shear layer. The consequence of these observations is that the additional information obtained by measur- ing as many terms as possible will lead to improved accuracy in the predictions of moments of the vorticity fluctuations and the measurement of the dissipation. l()8 4.3 Statistics of the Vorticity Field 4.3.] Examples of Vorticity Time Series Prior to the presentation and discussion of the statistical properties of the vorticity field, it is instructive to first view segments of the time series data. This will allow several of the qualitative aspects of the vorticity field to be observed. A sample of the spanwise vorticity (001(0) time series is shown in Figure 4.18. These data were acquired in the center of the shear layer: 11:0 at the streamwise location x//80 = 675. The mean vorticity at this location was found to be —8L7/8v = —9( l/sec) . Several features can be observed from this time series segment. First, there exist motions which appear very small in scale (high fre- quency), with very large vorticity magnitude compared to the mean vorticity. The root mean square of vorticity from the full time record was (I)z = l76( l/sec) . Although the largest excursions in magnitude appear to be of relatively short time scale, there also appear to be large regions of fluid whose vorticity is highly correlated. An example of this can be observed near the one second mark in Figure 4. l 8. The time duration of this large motion is approximately 0.24 seconds. Using Taylor’s hypothesis, this corresponds to a length scale of 0.78m, or 3.59. This is a remarkably large scale vortical motion, especially since the sign of the vorticity is positive. That is, opposite that of the mean vorticity. A fur- ther discussion of the length scales of the vorticity field will be given in Section 4.5. Another notable feature of the time series data shown in Figure 4.18 is the presence of regions of fluid with near zero vorticity. It is possible that this results from large scale motions which transport fluid from either the high speed or low speed streams across the shear layer. In Figure 4.19, a second example of the vorticity time series is shown from the l ()9 same data set. These data show an extended period of time in which the fluid is rotational (lwzl > 200( l/sec)) , although there is little or no high frequencies observed. This behav- ior was first described by Haw et al. (1989), and considered in detail by Foss et al. (1995) in terms of the difference between “intermittency" and "activity intermittency”. A third example of vorticity time series is shown in Figure 4.20. The location of this mea— surement was x/90=675, n=3.9. Here, on the high speed side of the shear layer the mean velocity is not measurably different from the free stream velocity. The vorticity is, how- ever, still quite intermittent. One can observe from these data that small regions of highly turbulent fluid convect past the probe. 4.3.2 Histograms of Vorticity The population of vorticity values that occur at a given spatial location can be character- ized by the probability density function (pdf). This provides information regarding the scaling properties of the vorticity fluctuations. Although the shape of the pdf can be com- pletely described by the moments of the distribution, it is first instructive to observe the pdf at various conditions. The present experimental facility provides an opportunity to observe the vorticity histogram over a range of Reynolds numbers and intermittency con- ditions. The true (continuous) pdf is approximated experimentally by the histogram of the time series measurements. In order to “fill out” the entire histogram to best approximate the pdf, very long time series are required to converge the statistics of the vortical motions which occur relatively infrequently. Therefore, four time series were acquired in addition to the traverses described above. These data were acquired in the center of the shear layer, 110 at two streamwise locations: x/00=384 and 675. and two free stream velocities: UO=3.5 and 7.1 m/s. Time series of (i)z were acquired at a rate of 4000Hz for a duration of 30minutes. A summary of the four conditions, including several of the characteristics of the time series data, are shown in Table 4.1. The histogram of the vorticity acquired at these points is shown in Figure 4.21. These data Table 4.1: Measurement conditions data shown in Figure 4.21 x/GO (Iii/Os) UL.“ U1 (33;) (I): 811:: Rel ‘ 0 0 ((02) A 384 7.1 0.47 0.15 1 1.3 227 0.23 835 B 384 3.6 0.52 0.17 8.6 136 0.30 411 C 707 3.6 0.55 0.18 10.9 124 0.26 551 D 707 7.1 0.50 0.15 17.4 188 0.32 1210 are shown in semilog format, with each histogram shifted by an order of magnitude so that they can be distinguished. The first notable observation is that the shape of the histograms is similar in all four cases. This is quite remarkable given the range of Reynolds numbers between the measurements. The shape can be characterized by the exponential tails for large ‘sz and a lack of symmetry in the population of measurements with small |w2| . This characteristic shape is nearly identical to the histogram of vorticity measured by Antonia et a1. (1988) in the wake of a cylinder. In that study, the measurements were taken in a turbulent region where the mean vorticity was non-zero, and with a turbulence Rey- nolds number of approximately Rel: aka/v=60. The “cusp” near the (02:0 location is 111 distinctive in both the shear layer and the cylinder wake, and clearly indicates the asym— metry of the population of measured (it)Z values. The insensitivity of the exponential tails of the pdf to the Reynolds number is further sup- ported by the direct numerical simulation (DNS) results of Cao et a1. (1996). These com- putations were conducted using a cubic domain with periodic boundary conditions. Forcing at low wave numbers was used to balance the viscous dissipation of kinetic energy. The circulation of the flow field was calculated around small rectangular domains. In the limit of a very small circulation domain, this calculation closely resembles the micro-circulation algorithm used in the present experimental method (see Section 2.2.3. 1). The population of circulation values was used to create a histogram of the realizations. It was found that large circulation areas led to histograms which appeared nearly Gaussian. However, as the circulation area decreased, thereby approximating a true vorticity mea- surement, the tails of the pdf became more exponential in character, with a distinct cusp at center of the histogram where (0:0; see Figure 3 of Cao et al. ( 1988). Naturally, the cusp is symmetric in the computational work because of the symmetry of the boundary condi- tions. Time series of spanwise vorticity in a high Reynolds number atmospheric boundary layer have also been acquired. The histogram of these data are provided here for comparison with the shear layer results; see Figure 4.23. The flow conditions are characterized as a flat wall boundary layer with Re9~106. The turbulence Reynolds number was determined to be Rex=2500 at the probe location y+=3800. The ratio of the probe scale to the Kolmog- orov length scale in this flow was 2.5. The features which are of interest in Figure 4.23 112 include the Gaussian-like region near the small values of [(1):] and the “tails” of the pdf which have an obvious positive curvature. Further discussion of these data is provided in the conclusions (Section 4.8). The data and references described above vary in both geometry and Reynolds number, but do not account for varied intermittency levels. The histograms of (it)Z at the center of the shear layer (11:0) where the fluid is almost always rotational, and at the high speed edge (n=3.9) where the fluid is intermittently turbulent, are shown in Figure 4.22. It is observed that the n=3.9 histogram exhibits exponential tails similar to the 11:0 data. Note also that the slope and magnitude of the exponential portion are not symmetric for positive and negative realizations. The central region (Ito'I/(I) < l of the histogram at 11:39 is charac- terized by a Gaussian like histogram (note that a Gaussian distribution is represented by an inverted parabola in semilog coordinates). 4.3.3 Profiles of the Vorticity RMS The unsteadiness of the fluctuating vorticity field can be partially characterized by the RMS level of each component of vorticity. These values were obtained directly using the time series from the vorticity probes, and indirectly using the velocity gradient informa- tion described in Section 4.1.3. If the assumption of homogeneous and isotropic turbu- lence is used, then the mean square of each component of vorticity is equal, and related to the streamwise velocity gradient by: (3,? = $63!)" 4.37 113 The assumption of axisymmetric turbulence can also be used to relate the vorticity to the velocity gradients. In this case, only the spanwise and lateral components are equal, and are given by George and Hussein (1991): (If, = a: = (g—DZ +(3—‘j +(%—? 4.38 The estimates of the vorticity variances were calculated from the velocity gradient statis- (to MN tics presented in Section 4.2.3 based on equations 4.37 and 4.38. These data are shown along with the directly measured values of (i): and (I), in Figures 4.24 and 4.25. The data were sealed with both the integral time scales (ti/U”) and dissipation time scales ( VG/ U5 ). The comparison of the two streamwise location indicate that the integral time scale leads to values of (1)20/ U0 which increase substantially in the x direction. For exam- ple, the maximum value of (Ike/U” increased from 4.8 at x/00=384 to a value of 7.25 at x/00=675. The dissipation time scale leads to values of (I): lvO/U; which increased only slightly from 0.0182 to 0.0205 for the two streamwise locations. This relative agreement using the dissipation time scaling could be used to argue that the vorticity fluctuations scale self similarly. However, this increase is considerably greater than the estimated uncertainty of the measurement (roughly 3x10'4 based on sample size and probe uncer— tainties, see Section 2.2.3.1.) This increase in vorticity fluctuation levels when using the dissipation scales will be discussed with the conclusions of this chapter (Section 4.8), since the explanation of this result will require the calculation of the one dimensional vor- ticity spectral density function, which is described in Section 4.6. 114 The agreement between all the directly measured values of (by and ml with the isotropic relation (4.37) is quite surprising, especially given the lack of isotropy shown in the vari- ous velocity gradient statistics in Figures 4.16 and 4.17. The agreement with the axisym- metric prediction (4.38) is also encouraging, although not surprising given the support for the axisymmetric assumptions shown previously, as well as the observation that (byza)Z throughout most of the shear layer. Measurements of the streamwise component of vorticity were also acquired. Because the measurement area of this probe is 5.6mm x 5.6 mm, it is explicit that the spatial smoothing of these measurements will severely under-predict the actual vorticity fluctuations. For comparison, the measured values were scaled by an arbitrary factor of 2.7, and plotted along with the axisymmetric calculation by Hussain and George ( 1991): —2 _ ACE gag) (bx — 3 3 +3 82: . 4.39 These data are shown for both measurement locations in Figure 4.26. The arbitrary multi- plier of 2.7 was only used to show that the shape of the directly measured and predicted profiles agree well. Assuming the values from the axisymmetric calculation (eqn. 4.39) are correct, the ratio (was was found to be 1.28 at the shear layer center, compared to the value (by/('1)z z 1.35 found by the direct measurements of Balint and Wallace (1989) in a two stream shear layer. Note that the nearly equal values of the spanwise and lateral components, and the larger values of the streamwise component of vorticity, supports the assumptions of the theory of axisymmetric turbulence over that of isotropic turbulence. 115 4.4 Analysis of the Reynolds Stress The Reynolds averaged Navier-Stokes equations can be written as: _aa,. ' a( P5 2 :9 ‘fi 440 lL— = —— - i'+ ll U—puiu- . where 3;]- represents the time averaged rate of strain tensor. The modeling of the Reynolds stress u'iu'j is the subject of considerable research: see. e.g., Pope (2000) for a review. The physical mechanisms which lead to the Reynolds stresses are still a topic of research, since no universal model has been found that relates these terms to the mean flow variables. The single stream shear layer provides an interesting environment in which to study the Reynolds stresses. One reason for selecting this flow field for study is because the only off-diagonal component of the Reynolds stress tensor which is non-zero is u'v' . Assuming that the mean viscous stress and pressure gradient are negligible, the momentum equation for the streamwise direction (i=1) becomes (see, e.g., Pope (2000) page 1 l3): aBj+VB_& _ _au'_v'_a(u'2—v'2) Bx By _ By Bx The final term of equation 4.41 is usually neglected in thin shear flows, which results in 4.41 what is usually termed the boundary layer form of the momentum equation: [la—E.‘+\7QE — _QL——V: Bx By_ By ° Given that the left hand side of this equation is composed of variables which have been 4.42 found to be self similar, the Reynolds stress is assumed to be of the form: 0 ' uv hm) = —2. 4.43 0U 0 1.16 The use of o in the denominator of equation 4.43 is to make h(n)=O( 1). By combining equations 4.12, 4.19, 4.42, and 4.43, the dimensionless stress gradient can be written as: 71 Mn) =f'{—1+J' f(§)d§}. 4.44 This equation was integrated, and plotted with the experimentally measured values in Fig- ure 4.27.The agreement is satisfactory, and the values are similar to those measured by Bruns et al. (1991) and Champagne et al. (1976). Because the gradient of the stress appears in the mean momentum equation (4.42), it is instructive to look at the stress deriv— ative as well. The profile of h'm) was determined by fitting the data shown in Figure 4.27 with a polynomial, and differentiating; see Figure 4.28. One feature of the Reynolds stress derivative is the zero crossing at nz0.6. Consider the terms of equation 4.42 in the untilted (x,y) coordinate system at this location. In Section 4.2.2.1 it was observed that for T]<10_2 was found to be 1.9 (1113/53). The assumption of isotropy also leads to a unique relation for the third order longintudinal structure function (see, e.g., Landau and Lifshitz. 1987): . . 3 4 D|H(rl) = [u (x+rl)-u (1')] = —§€,r| 4.67 The interesting feature of this equation is that there is no arbitrary constant. The measure- ment of dissipation is then only a matter of achieving a high enough Reynolds number such that the compensated structure function D, I I is constant over a wide enough range of r/n. The compensated structure function for the present velocity field shown in Figure 4.31. This exhibits a relatively constant value near 81:1.30 over the range 10-l.8 due to limitations with the traversing mechanism. These data were acquired for 300 seconds at a rate of 3000Hz. The spanwise correlation of velocity is shown in Figures 4.42 and 4.43. The dimensionless data show excellent agreement between the two streamwise locations indicating that this statistic scales in self-similar coordinates. Several observations can be made from these figures. First, the two dimensionality of the flow in both the high and low speed regions is evident from the high correlation values, in agreement with the Browand and Trout (1980, 1985) results. For example, for n>3.5 the correlation is 0.9 or larger for O 0.01 . This result is reasonable, and suggests that correlations between u and v are significant only at low wave number, and that local isotropy is supported for higher wave number. 1 40 Note that the coherence is more sensitive to anisotropy than is the auto-spectral density, since the auto-spectra BW and Eww are both reasonably well approximated by the isotropic relation 4.78 over the entire range of the spectrum. In addition to the auto and cross spectrum of velocity components, the vorticity spectra can also be computed from the experimental measurements. Antonia et al. (1987) derived the isotropic relationship between the one dimensional vorticity spectra and the energy spectra: K1 °°E(K) l °°E(K) 2 2 ¢(,,z(l(l) = 3].“ —K—dK + EL. —K-[K + K1 ]d1<. 4.80 The model spectrum (equation 4.74) was used to calculate what the vorticity spectra would be in an isotropic flow. This, along with the experimentally determined spectra of (it)z are shown in Figure 4.53. These data are remarkable in that they are not similar to the isotropic prediction in any way. There is considerably less energy in the high wave num- ber range, and considerably more energy at low wave numbers than that predicted by equation 4.80. This result is especially interesting given the excellent agreement between the measured and isotropic vorticity spectra shown in the boundary layer of the present facility shown in Chapter 3. Antonia et a1. (1998) also found very good agreement between equation 4.80 and the directly measured vorticity spectra for grid turbulence, and Antonia et al. (1987 and 1996) found the similar results in wake flow. Ong and Wallace (1995) found the same in their boundary layer measurements. A notable difference between the boundary layer and wake flow studies and the present measurements in the free shear layer which created the spectra shown in Figure 4.53 is the existence of the organized large scale motions of the flow. The relative contribution of the 141 different scales of motion to the vorticity variance can be observed by plotting the com- pensated spectra: 19%, in semilog format: . 2 00 00 ((1):) =1 $1.10?!) (1K1 = j K, "1.011(1) d(ln(K1)) 4.81 0 ‘ 0 ~ The integrand of the second right hand side (Kl- ¢(‘)(Kl )) is plotted in Figure 4.54. It can be observed from this figure that a large portion of the integral contribution is from the low wave number range of the spectra. 142 4.7 Circulation Density A distinctive feature of the shear layer flow field is the existence of the large scale motions. Many previous research efforts have attempted to determine the nature of these motions, with the hope that an improved understanding of orderly structure may lead to improved turbulence modeling and control. The previous sections have described the results of single and multi—point measurements in the shear layer, which support the exist- ence of these large scale motions. In particular, the scales of the velocity correlations (Fig- ure 4.34), the temporal and spanwise coherence of the vorticity (Figure 4.47), and the velocity-vorticity correlation shown in Figure 4.48 all indicate that motions of relatively large length scales are an important part of the shear layer dynamics. The large scale motions of a shear layer are often referred to as vortical or vortex motions, or coherent structures in the literature; see, for example, the review article by Thomas (1991). However, information about these motions is generally limited to information derived from two sources: flow visualization, and measurements in the irrotational fluid near the active shear layer. For example, the studies by Dimotakis and Brown (1976), Winant and Browand (1974), and Hileman and Samimy (2001) use results of flow visual- ization images to obtain information about the structure of the shear layer. The measured irrotational velocity fluctuations were used by Browand and Trout (1980, 1985). Both of these techniques were also used in the present observations which described the large scale motions of the near separation region in Chapter 3. However, these measurements do not easily quantify the importance of the large scale motions, and, could lead to significant misinterpretations about the flow field. As an example of the possible misinterpretations, Hama (1962) has shown that even an infinitesimal disturbance to an otherwise uniform 143 thin sheet of vorticity will produce streaklines that “roll-up” into what appear to be large scale vortices. Hama concludes with a comment about flow visualization data in which images of streakline patterns are recorded. He states that “practically no truth can be obtained as to the nature of time-dependent phenomena and that images, which one might receive from such observations, can be entirely misleading”. A technique for extracting information about the coherent motions of the shear layer not mentioned above is conditional averaging; see, e.g., Browand and Weidman (1976), and Hussain and Zaman (1985). Its application by these authors has made use of a trigger sig- nal from the irrotational flow outside the active shear layer along with measurements of u and v in the shear layer. The velocity field is then averaged using points which satisfy a certain condition for the trigger signal. This technique is instructive, and supports the existence of large scale motions. Since only a fraction of the total time series of velocity (which corresponds to a specific event in the trigger signal) is used in the averaging pro- cess, conditional averages do not provide a complete description of the flow field at all times. This preamble is to establish the author’s motivation to obtain a measure of the “coherent structures” of the shear layer which is both quantitative, useful, and representative of the characteristics of the flow. Specifically, it is recognized that they to not simply seek to characterize a “conditional event.” The circulation density (7) of the shear layer is pro- posed as an alternative response to this motivation. The circulation density is defined as the integral of the spanwise vorticity along a line perpendicular to the streamwise direc- tion made dimensionless with U0: 144 0° a(v/UO) J ———dy 4.82 If :— (l)d’:— Y U0 2) 1+ 8x —m —m That is, 'y represents the “circulation per unit length.” In general, 7 can be a function of x, z, and t, since only the y dependence of vorticity has been integrated. However, note that a constant free stream velocity requires that the time average value of the integral on the right side of equation 4.82 at a given (x,z) location must be essentially zero, given that a(V/U,,)/ax « 1 .That is, 7 = —1. In order to gain a better physical interpretation of the circulation density, schematic exam- ples of y in several stages of a temporal (y-t) shear layer are shown in Figure 4.55. In the first example, a uniform sheet of vorticity is shown in gray, with a corresponding constant value of y=- l. The pdf of an experimentally determined population of 7 values might look like a very narrow Gaussian distribution, with a variance corresponding to the uncertainty of the measurement. The second example shows the case of a laminar shear layer undergoing a two-dimen- sional linear instability. The vorticity field during the linear and nonlinear stages of growth have been studied in detail both numerically (see, e.g., Pozrikidis and Higdon (1985)) and analytically (see, e.g., Michalke (1965)). During the linear stages of growth, 'yhas a sinu- soidal shape with increasing amplitude. A measured pdf of y for this case would be “dou- ble humped.” The third example represents the condition in which the coherent motions roll up into dis- crete vortex motions with a Gaussian core of vorticity. In this case, regions of nearly irro- 145 tational motion will exist between these motions as shown. A measured pdf of 7 would have a local maximum near 7:0, and a second peak located at y<~ 1. These idealized examples illustrate that information about the circulation density in the turbulent shear layer will be instructive with regards to the large scale dynamics. Ideally, time series of y at multiple (x,z) locations would permit a full characterization of the circu- lation field. However, experimental limitations do not permit such difficult measurements. In the present experiments, the pdf of yat a single (x,z) location was acquired. Although far from complete, this does provide a significant amount of previously unavailable infor- mation. For example, if the large scale motions of the flow are comprised of regions of rotating fluid which exist between regions of relatively small circulation, then the pdf of 7 would look something like that which is shown in Figure 4.55c. Note that this result assumes that each coherent motion contains the same net circulation. If coherent motions were to exist with a wide range of circulation values, then the histogram would be consid- erably wider, which could lead to a histogram of 7 values that is no longer double peaked. As a final example, if no large scale motions exist and the turbulent fluctuations distribute the vorticity more evenly, then the pdf would show few realizations far from the mean value of -l. The only reference to the circulation density of a turbulent shear layer in the literature is found in the direct numerical simulation (DNS) of a (y-t) shear layer by Rogers and Moser (1994). A similar definition as equation 4.82 is given, although they ignore the mean value, such that only the fluctuations: Y are considered. However, no information about Y is provided except to say “variations [in Y] are still dominated by small scales.” They then apply an arbitrary two dimensional spatial filter in the streamwise and spanwise directions 146 “to extra that larg HOWCW exampl 4.7.113 llie Cir nents 0 4.56. T size as; The ve (PIV) EM .1 0f the Streai The s was ' liOn , ure 4 Henc been “to extract the large-scale behavior.” This “roller parameter” (9?) is successful in the sense that large scale quasi-two dimensional regions with 9i<0 are visible in the computed data. However, significant spatial smoothing is required in order to observe this structure. For example, it would be difficult, and of limited use, to derive a transport equation for (58). 4.7.1 Experimental Configuration The circulation density was experimentally determined by acquiring the u and v compo- nents of velocity along a narrow rectangular region. This is shown schematically in Figure 4.56. The integration defined in equation 4.82 is modified to account for the finite domain size as: y = ii = {CIA—in?- at) 4.83 The velocity field was measured using a two camera digital particle image velocemetry (PIV) system. The equipment was a Dantec 2000 processor, with lk x 1k Kodak Megaplus EA] .0 cameras, and a New Wave Research Minilase 3, NDzYag laser. The configuration of the equipment is shown schematically in Figure 4.57. The primary and entrainment streams were both seeded using two Roskco 1600 smoke generators. The streamwise location of the measurement was x/90=170. The momentum thickness was 71mm at this location. The 313mm extent of the integration area in the lateral direc- tion was centered such that the range -2.21 0 for all n in the rotated coordinates as introduced in Section 4.2. That is, a positive mass flux exists across any surface parallel to the 3r axis, which supports the second statement of the conclusion given above. Note that the surface representing the centroid of vorticity is given by n =0, where g(fi) = 0.375 . It is inter- 150 esting to contrast this result with the observation that mass cannot cross a time steady sheet of vorticity. That is, the unsteady character of the vorticity field must be taken into account in order to explain this conclusion about the mean flow variables. It would also be of interest to extend this result to two stream shear layers. For example, the flow visualization results of Brown and Roshko ( 1974) clearly indicate the movement of the shear layer towards the low speed side. This movement, as well as the angle of the bounding wall, which defines the low speed entrainment rate in a two stream shear layer, could then be correlated to the shear layer growth rate. 2. The flow field is three dimensional in nature, as illustrated by the spatio/temporal velocity and vorticity correlation fields. However. a volume integral of the vorticity can be written which is two dimensional. The length scales of the shear layer can be divided into three general categories for the purpose of the following discussion. In the turbulence literature these are usually termed the energy containing, inertial, and dissipation length scales. Conclusions regarding the nature of the inertial and the dissipative scales will be given later in this section. The large scales of motion are of primary interest because they contain a large fraction of the turbu- lent kinetic energy. They are responsible for the majority of the material and momentum transport. In contrast to the small scales, it is considerably more difficult to develop strong conclusions about the nature of the large scale of the fluid motions. The present conclusion and discussion examines the large scale motions through the spa- tial correlation of velocity and vorticity. An interesting summary of these correlations is obtained by observing the spatial separation required for the first zero crossing of the cor- 151 relations acquired at 11:0. These results are tabulated in Table 4.3. These data show some relative agreement that indicates that the large scale motions are correlated over a length scale of order 2.59 both in the streamwise and spanwise directions. The exception to this result is the autocorrelation of the streamwise velocity (Ruu(t*)) which exhibits longer correlation times. It is thought that this is due to the use of the time delay variable rather than a true spatial correlation. That is, the large scale motions have a temporal correlation which is quite long (sometimes referred to as “quasi-periodic”) which leads to the long correlation time of the streamwise component of velocity. Further support for this observation can be drawn from the comparison of (Ruu(t*)) at the loca- tions n=0 with those acquired in the high and low speed irrotational flow; see Figure 4.34. Specifically, the correlations decrease towards ze(o\)approximately the same rate (e.g., |Ruu(t*)| < 0.05 for t*>22). A distinct point of contrast can be made when these results are compared to the spanwise correlation of velocity obtained in the irrotational flow in both the high and low speed sides of the shear layer; see Figures 4.42 and 4.43. These data indicate high (>0.9) correla- tion values over length scales of order of the shear layer thickness. This is in close agree- ment with the results of Browand and Trout (1980, 1985) in which correlation lengths of order 209 were observed in a lower Reynolds number two stream shear layer. The inter- pretation of the large scale motions as “two-dimensional” structures can be explored from these results. The term two-dimensional implies 9( )/az = 0 as a necessary condition. However, the measurements of velocity gradients shown in Section 4.2 as well as the vor- 152 ticity measurements in Section 4.3 have shown that the smallest scales of motion exhibit many of the properties of homogeneous and isotropic turbulence. It is therefore instructive to examine what is two-dimensional in the shear layer. The approximation 912(0an > 3_2)/9z z 0 has been shown from the correlation measurements shown in Figures 4.42 and 4.43. This velocity can be written in terms of the vorticity in the shear layer as: . _ 1 (“B x 7) 7 a(n > 3.2) — HUN-K r3 (dx)(d_v)(d‘.) 4.84 where the limits of integration are given by the bounding walls of the facility. The deriva- tive of this expression can be take with respect to the z direction to give: 3% = agitHK‘Br—j—r)(ill-)(dyxdz) = 0. 4.85 That is, the volume integral of the vorticity field expressed in equation 4.84 is two dimen- sional, even though the vorticity itself is highly three dimensional. That is, it is the integral properties of the vorticity field which lead to the observations of two dimensional motions. 153 Table 4.3: Summary of zero crossing points of correlation functions acquired at 11$. Correlation Location of first R=0 condition Ruu(t*) 4.89 thz") 2'97 R,,,,,.(l*) 2.84 Ram} t*) 2.35 Rm‘wfl.) 2.40 RwrmfAz/O) 2-0 RW(Az/9) 1.56 154 3. The histogram of circulation density values indicates that the description of the large scale motions of the flow may be more accurately stated as fluctuations in the circula- tion density, rather than “structures” or “vortices”. Considerable research has been devoted to the characterization and understanding of the large scale motions of shear layers. Observations from both flow visualization and veloc- ity measurements in the irrotational flow adjacent to the shear layer have been used to infer the properties of the large scale motions. For example Brown and Roshko (1971) used a shadowgraph technique to investigate the large scale structure of a two stream shear layer. Since then, significant research efforts have focused on the characterization, predic- tion, and control of these motions (see Thomas (1991) for a review). These results often lead to an idealized conceptual picture of the velocity/vorticity fields which exist in high Reynolds number shear layers. The circulation density measurements provide informatino about how the vorticity is dis- tributed within large scale motions of the flow. For example, the pdf of yis nearly Gauss- ian, and not double peaked. The relatively large standard deviation of 7 values measured indicates that considerable variability exists in the motions which convect past a given streamwise location. Also, the occurrence of y==0 is observed to be a fairly rare occurrence. These results are not inconsistent with the observations of the large correlation lengths observed in the velocity and vorticity data. These latter observations suggest that the fluc- tuations in the circulation density also have a significant correlation in both space and time. The suggested spatial correlation of the circulation density is consistent with the flow visualization results of Brown and Roshko ( 1971), and Dimatokis and Brown (1976). 155 That is, spatially correlated fluctuations in 7 could lead to the observed “roll-up”. Future research would be needed to confirm this statement. Conclusion 3 above states that this perspective of the large scale motions may be more useful than the generalization of the flow field as containing coherent structures or vortices. 4. The energy spectra have been found to be well represented by a model spectrum, and the assumptions of local isotropy. In contrast, the vorticity spectra do not appear to be even qualitatively similar to the predicted spectra using the same assumptions. The conclusion regarding the velocity spectra is apparent from Figures 4.49 through 4.52. The model spectrum provides a close representation of the El 1(k!) data. Assuming both homogeneous and isotropic turbulence, the model spectrum also represents both E22 and E33 quite well. This is especially notable at the lowest wave numbers where the fluid motions are clearly non-isotropic. The H12 data (see Figure 4.52) however, are more dis- criminating in that km>.01 is required for local isotropy to be approximately valid. Given these results, it is quite striking that the vorticity spectra cannot be predicted by local isotrOpy predictions as shown in Figure 4.53. Note that the one dimensional spec- trum of vorticity in an isotropic flow does not have a log-linear region as does the energy spectrum. Rather, the one dimensional spectra “flattens out” as shown, leading to decreased contribution from the small wave numbers. In isotropic turbulence the enstro- phy spectra also represents the dissipation spectra. It is consistent with K41 theory that small wave numbers would have minimal contribution to the enstrophy. However, the anisotropy of the large scale motions of the flow have led to vorticity correlations which 156 are non-zero over a large spatial extent indicating that the vorticity spectra contains signif- icant contributions from the low wave number range. The fact that the vorticity spectra does have a log—linear region with slope of k'1 is also of interest. This is because the compensated spectra (ki¢m) shown in Figure 4.54 is relatively constant over a wide rage of wave numbers. This implies that the relatively large scales also cause a significant contribution to the vorticity fluctuations. This is in contrast to the homogeneous-isotropic result where the vorticity and dissipation spectra are related only by viscosity, and the dissipation spectra obtains the majority of the contributions from the high wave number range. 5. The stochastic values of the velocity, velocity fluctuations, and velocity gradient fluc- tuations scale in a self similar manner, whereas the vorticity fluctuations increase with increasing Reynolds number. It is recognized that exact self-similarity cannot be achieved in a shear layer. This is because the ratio of the smallest length scales to the largest length scales is a function of Reynolds number. By scaling the dissipation with U3/9, the ratio of the integral to Kol- . . 3/4 . . . . . mogorov scales 1S found to be 9/1] K ~ Re . Stochastic quantities which scale only With the integral scales of motion (that is, have no dependence on nK) will exhibit self-similar scaling. Similarly, variables which depend only on the Kolmogorov length scales, with no explicit dependence on the integral scales of motions will also scale in a self-similar way. It will be shown in the following discussion that the vorticity fluctuations depend both on 9 and on 11K, and that this leads to a direct dependence on the Reynolds number. 157 The spectral representation of the vorticity field exhibits a log-linear region with a slope (pm ~ [(1 . This can be observed by the roughly constant value observed in the compen- sated spectra shown in Figure 4.54. The dimensionless variance of the vorticity fluctua- tions can be expressed as the integral of the spectral density: 00 kan:l ((092: j(¢,,)d(k.n,,) = j (k.,,)dtlog(km,.)r 4.86 0 truism/9 In the right hand expression, the limits of integration are modified to account for the observation that there is little or no contribution to the integral at scales larger than the integral scales of the f low, nor is there a substantial contribution from scales smaller than the Kolmogorov length scale. The Reynolds number dependence of the vorticity variance can now be concluded from the observation that the ratio of largest to smallest length scales varies as Rem; see, for example, Pope (2000). Therefore the integrand “1%) is roughly constant, and the lower limit of integration decreases with increasing Reynolds number as observed in Figure 4.54. Therefore, the value of the integral increases with increasing Reynolds number. The maximum values of the vorticity RMS measured at 11:0, x/90=384 and 675 support this conclusion; see Figures 4.24 and 4.25. Additional support has been found by compil- ing the vorticity measurements of several authors in other shear layer facilities. These include Bruns (1990), Lang (1985), Loucks (1998), and Balint and Wallace (1988). These data are summarized in Figure 4.59. Each of the data sets reported (where more than one streamwise location was measured) indicates an increase in the dimensionless vorticity RMS with increasing Reynolds number. The compilation of all these data also support the 158 conclusion. The scatter observed between the different data sets is larger than that which would be expected from the uncertainty calculations of the individual measurements. This most likely represents the effects of the different boundary conditions of these studies, such as velocity ratio and the state of the boundary layer at separation. 6. The pdf of vorticity is highly insensitive to Reynolds number, but quite sensitive to the level of intermittency. This conclusion was drawn from the comparison of the histogram of observed vorticity values obtained for the following flow conditions: a cylinder wake at Rel=60 (Antonia et al. (1988)), DNS calculations of isotropic turbulence at Rek=216 (Cao et al. (1996)), shear layer measurements in the range 41 1 B B y ——> —> x —> 7 4 ,, . L ltlltlltllitlllllliilll V9 e L a l\ /1 Figure 4.1 Schematic of control volume and coordinate system used in Section 4.1. Note that the growth rate is not shown to scale. )— lllllllilllllllllllllll Figure 4.2 Illustration of the physical interpretation of 5* in the boundary and shear layer 161 K< ‘<> / S $4 ////////////\Ln:o My]; \ Figure 4.3 Schematic of the un-tilted coordinates (left and the tilted coordinates (right) for use in similarity analysis 1.2 p E 10 L Brunstl990) TaFU—‘A‘ U 0 x/90=370 ,K’f" o 0.8 . 1:1 x/90=650 44% f” 0.6 "/4. / . ,v’ 0.4 . «1’ 1 ~’" .21” 0.2 'I’i A .. «9’ 0.0 . ‘ . 7 _4 -3 -2 -1 o 1 2 3 4 11 Figure 4.4 Streamwise mean velocity f(n) 162 Cllct <2) -2 11 1 2 3 Figure 4.5 Lateral mean velocity: gm) in tilted and untilted coordinates. T I l I l l I l T T I T l l l l l l l l a 6‘ 3 r. 0 x/90=370 r-— I ——1 .I o I. I x/90=650 1- . ' — .__. "II ____ L. O a I 4 y f— _ .— . O . .a O T T 1 l 1 l l 1 l l l 1 #1 I 1 l 1 1 1 I -22 - l (l l 22 El ‘1 -0.2() -040 -4 1.22() 1 .()() ().13() ().(3() ().41() ().IZ() ().()() ().IZ() (1.1 (1 (1.1 2! (1.1115 ().()<1 ().()() Figure 4.6 Root mean square of the streamwise velocity fluctuations 1(331 ~ (1.14 1 1 1 1 1 1 1 I 1 1 1 I 1 1 1 I 1 1 1 I 1 1 I I 1 1 1 v e . . - _ (1.12 >—— . x/O,,=37() . : I I 0 fl 4 I U0 f I x/B..=65() o" ' ' ‘ ().l() '—‘ .74 - . ~ *— 0 I I . ‘1 ( . 1 ~— _, 1( 8 . . . ._ ‘ . _ (1.06 r——- '. I _ A ‘. *— —1 . . (1.114 — ' ' _. . h— . —< . (1.02 —- .— (1.110 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 L 1 1 -3 -2 -1 (1 1 2 3 4 Figure 4.7 Root mean square of the lateral velocity fluctuations ~ (1.14 1 1 1 I f 1 1 I 1 r 1 I 1 1 1 I 1 1 1 I 1 1 1 I 1 1 1 w h . , ‘ U. (1.12 1-— . V9,,z370 ‘ I ' I. '- 0 F I V9..=650 "' ' ° — (1.1(1 ~—- .. I —‘ r—— . —‘ O I I (1.118 -— . .— _ . I I - (1.06 >—— I o — I r— . I. - 0.04 h— . I ~ 9 ' I (1.02 >— . __ I 000 l 1 1 1 L 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 l 1 1 1 1 1 Figure 4.8 Root mean square of the spanwise velocity fluctuations 164 1 . . . , . . . . . x/90: 384 (175 200 _ + _g_ correction for (du/dx)2 (Cu) d + {1— correction for (dv/dx)2 (CV) T Y T T T Y r T I I ‘ 1.50— 1.00 0.501111.11111111111..111L41114 -3 -2 -l () l 2 3 4 11 Figure 4.9 Profile of correction factor for streamwise and transverse velocity gradients (see equations 4.25 and 4.26) 165 1 .4E—4 1.2E-4 l 1 .0E-4 1 8013-5 1 6013—5 1 4013-5 r 2.0E-5 *- 0.0 T I T T T I T T T I T T T I T T x/o0 = 384 675 T I Tr r T I T T T . O uncorrected Figure 4.10 Profile of velocity gradient fluctuations: du/dx 1 1.25-4 .,...,.../_,,,,_..,..., .., 2 j ‘ (dv/dx)“ / 3 1.0134 — - _ , . (U0 /V9) 6 ' 8.0E—5 ~ ’ ' . 1 6.0E-5 — 4.(1E-5 1— X/GOZ 384 675 -.\\\ —g— —0— uncorrected at, 2.0E~5 — I C] corrected 1 \- ().O ‘ ‘ 44 1 r l 1 l 1 1 -3 -2 -1 (1 1 2 11 Figure 4.11 Profile of velocity gradient fluctuations: dv/dx 166 1.5E-4 . r 1 T TTr T T T (aw/(1x)2 (U03/v9) 1 013-4 1 5.0E-5 - x/eo: 384 675 _._ —O— uncorrected I [3 corrected 0.0 J A A l L A A l A A A l A A A L A A A Figure 4.12 Profile of velocity gradient fluctuations: dw/dx 4.0E‘4 T T T I T T T I T T T T T V I T T T I T T T I T T T 2 ., (du/dy 1' 3.5134 — s 3 . 1 (U0 /V9) 1034 x/9U: 384 675 ' —o— —o— l 1 2.5E-4 l 1 2.0E-4 T 1 1.5E-4 l 1 1 013-4 1 1 5 015—5 - — 0.0 A A A 1 A A A l A L A l A A A J A A L Figure 4.13 Profile of velocity gradient fluctuations: du/dy 167 2,5E'4 I I 1 I r T r I 1 1 1 I r r 1 I r r t I T T I I 1 r r 2 (du/dz)‘ 3 2013-4 — _ (UO /v9) 1,513-4 _ _ 1015—4 - _ 1 x/9(,= 384 675 5.05.5 — —o— —o— - ().O A A A A A A A A A A A A A A A A A A A A A A A i A A A Figure 4.14 Profile of velocity gradient fluctuations: du/dz 3.0E'4 1 1 1 I 1 1 1 I 1 r 1 I 1 1 fl f1 1 r T , {—ldV/dz 2 * moo: '384 (U03/v9) 2513-4 + 1 2.0E-4 A _ 1.5E-4 1- - 1 013-4 1 1 l 1 5.0E-5 0.0 A A A A A A A A A A A A A A A A A A A L444 A A A A Figure 4.15 Profile of velocity gradient fluctuations: dv/dz 168 3.00-1vrrvvitwvivnl.fir,...,.. L '1 2.50- 2.00- 1.50- _D_kl 1 ..Q.k2 1.00 —O—k3 0.50- k4 0.00 ‘ -3 Figure 4.16 Ratio of velocity gradient variances for x/9(,=384. 2.50 T T T I T I T I r T T I T T T I T T T I’ t l I I f T T 2.00 ~— 1.50 - 0.50 r 0.00 Figure 4.17 Ratio of velocity gradient variances for x/9,,=675. 169 1500 ‘ l I l I l 1(100 (02(t) (l/sec) 500 ll 1 , . ' l 0 '1" . - . ‘1 . i l. 1M1 1‘11 ‘ ‘m.n' __‘ 1' _ L It ' 1 ”WIWIIV I 1 1 II 11 w 1 1 I -l(100 -15(10 1 1 1 1 1 (1.0 0.5 l .(1 1.; time(sec) Figure 4.18 Example of vorticity time series at x/9(,=707, nzO. Periods of irrotational fluid are observed. 200 _ f . . T - T . . . . . . - , 100 :— h I 1 0 : 1 .1 1A A 11 A A 1\ M: (1) z(t) (“8%) V 1“ -100 .2 -200 1 l 1 _ 1 1‘11 1 : {too 1 U 1 400 5 A_ J A AA A l A A . l A I . 3.0 3.1 3.2 3.3 time (see) Figure 4.19 Example of vorticity time series at x/9(,=707, n=0. Note the extended period of rotational fluid with little high frequency content near the beginning of the time record 11100 . , . , - , . , - (1) z (l/sec) 5111) -5()() -l(1()(1 ‘ ‘ ‘ ‘ ‘ (1.00 (1.50 1.00 1.50 2.00 2.50 time (see) Figure 4.20 Example of vorticity time series at x/90=707, 11:39. The fluid is mostly irro- tatoinal with intermittent bursts of highly fluctuating vorticity. 1 7(1 Figure 4.21 Histogram of vorticity from conditions A,B,C, and D; see Table 4.1. f V I V V Y I T T f r ‘1' Y Y I 1 f V I Y -1» u b 4 IO» ~ .4 4 y y y Figure 4.22 Comparison of vorticity histogram between points of high and low intermit- tency. I71 ,.- 0’. 'a’ f ‘4’ 0", ’ .O’.O. A l l l -l0.0 -5.0 0.0 (Di/(Dz Figure 4.23 Histogram or spanwise vorticity in the atmospheric boundary layer at Rex=2500. Solid line represents the curve fit given by equation 4.87. I72 0.020 e 8 CD (003/ve)(”2’ 0.015 I 0.010 - 1 0.005 0.000 -3 I (Dz ____ Axisymmetric . coy _ Isotropic Figure 4.24 The RMS of the spanwise and lateral vorticity fluctuations at x/90=384 (Re9=7. 1x104). The isotropic and axisymmetric results were calculated from the measure- ments of the velocity gradient variances. y 8 1 ~ ~ 0020 ~ ‘ 7 w 36 52.9 3 (1/2) 1 U /v0 4 ( 0 ) 0015 ~ :5 U0 ~j 4 0.010 - : 1 3 0.005 - -: 2 f l 0.000Hauuu-n‘-~:u-_#-n.-fi0 -3 -2 -| 0 l 2 3 4 I (Dz [I'4Xl'4mm] ___- Axisymmetric . (Dy [1 .4x 1 .4mm] ISOtI‘OpiC Figure 4.25 The RMS of the spanwise and lateral vorticity fluctuations at x/90=675 (Re9=1.2x105). The isotropic and axisymmetric results were calculated from measure- ments of the velocity gradient variances. I73 0.030 N xx x/(-) =675 /’ ‘s O (O 0.025 » ,1, ° °\\ / 6 ,- U 3N0 (1/2) , I]. ’. . ‘\\ O \\\ ( o ) 0.020 - , u ° \ 4’," / 3‘ o\ 'I’ 0.015 - ,xfo/ \\ ;\ 0.010 - ./ \°\ A}: 0005 - ‘9 . O \\ ' i 0000 l 1 l l 4 l . l . -3 -2 -1 0 1 2 3 4 T1 2.7%)x [5.6 x 5.6 mm] Axisymmetric Figure 4.26 RMS of streamwise vorticity. Broken lines represent the predicted values from the measured velocity gradient terms assuming axysymmetric turbulence. The symbols are directly measured values from the 8—wire probe, that have been multiplied by the arbi- trary factor of 2.7. I74 0.25 7 r T I 7 I Y I I I Y T f I I I I I r I mi! 1 I 1 1 v b . x/00=384 -h( 77) 0.20 - I x/00=675 0.15 — 0.10 — 0.05 — 0.00 ' . -3 -2 -1 0 1 2 3 Figure 4.27 Measured and calculated Reynolds stress. 0.100 0.050 ~ -h '(77) 0.000 - I —0.050 I -0. 100 0150.11L144141.1.11LL1111111111 3 Figure 4.28 Gradient of the Reynolds stress. I75 0.300 7 T T I F r I i F F I I r I I I 1 1 l I I I I . . 0200 " +va" ‘ x/90=384 0.100 - + 5 0.000 — - 0.300.......,...,.-,,.fifl 0.200 r- - x"’o=675 0.100 - _ 0.000 - .. 41100-2 . . . -11 . 1 . (1) . . 1 i . . . 12 s L . 3 4 Figure 4.29 Measured velocity-vorticity correlations. 2'0 (Q /\\VI‘:11”IIII.LIEAI I III ‘ . _ U V‘JUV IIIHI 1,1: * * (5/31)(3/2) 1.5 - d (2 E110(1) k1 1.0 — _ 0.5 F" / ' - / 0.0 l l llLllll 1 L Lllllll 1 1 lllllll l l l -_ 10‘4 10'3 10'2 10“ Figure 4.30 Measured dissipation from compensated energy spectra: E1 1(kl) I76 I.5() II I Y Y YTTYTT T T YTYYIII -(5/4)*D 111/r1 l.()() ().fi() ().()() r1/T1K Figure 4.31 Measured dissipation from compensated third order structure function: D, 11 0.0025 1 U I I Y Y Y I I T V I 1 I I I . ' T'C. 89 0.0020 I — 0.0015 I 0.0010 T 0.0005 T 0.0000 -2 Figure 4.32 Comparison of the profiles of several dissipation estimates at x/00=675. The . TC . . . . . estimate 83 assumes isotroplc turbulence, whereas the others assume vanous sem1—1so- tropic properties. I 7777 0.0015 0.0010 P 0.0005 0.0000 -0.0005 P -0.0010 5 -0.0015 - -0.0020 F -0.0025 1000 = 384 675 . O Convection _. _ —-{‘] Dissipation ‘ A. Diffusion * 0 Production I Pressure transport Figure 4.33 The measured budget of turbulent kinetic energy. The ordinate represents the magnitude of the bracketed terms in equation 4.68 scaled by 0/ U:. 178 1.0 x/eO ___. 384 =-l.5 Ruu(t*) 0.5 - \ ‘1 675 \ \‘l \ \ \\ ’ \\‘ _ ‘T'b‘ 00 . g , 7 ,aww \ x’ T]: $ _ 11:3 1 _O.5 | 1 1 I 1 1 1 I 1 1 1 l 1 1 1 L 1 1 1 0.0 5.0 10.0 15.0 20.0 25.0 1.0 Ruu(t*) 0.5 . 0.0 -05 . 10" 10 Figure 4.35 The autocorrelations of Figure 4.34show in semilog format 179 1.00 0.80 0.60 0.40 0.20 0.00 -0.20 0.0 Figure 4.36 Autocorrelations of u,v, and w at x/0(,:675, 11:0. Ruaub Figure 4.37 Cross Correlation between probe (a) located at n=4.53, and probe (b) located at n=1.26. I80 0.15 Ruaub 0.10 0.05 0.00 -0.05 . -0.101 _0.15 1 I 1 I 1 I 1 l 1 1 1 I 1 I 1 ~40 -30 -20 -10 0 10 20 30 40 Figure 4.38 Cross Correlation between probe (a) located at 11:308. and probe (b) located at n=-0.194. 0.10 Ruaub 0.05 0.00 -0.05 _010 1 I 1 I J_ I 1 I1 1 I 1 I 1 I 1 -40 -30 -20 -IO 0 10 20 30 40 t* Figure 4.39 Cross Correlation between probe (a) located at n=l .76, and probe (b) located at n=-1.52. 18I 0.10 Ru aub 0.05 0.00 -().05 e ' _010 1 I 1 l 1 I 1 I 1 L 1 I 1 I 1 -4O -30 -20 -10 O 10 20 30 40 Figure 4.40 Cross Correlation between probe (a) located at n=.434, and probe (b) located at n=-2.84. 0.3 RWNb '—40 -30 —20 -10 0 10 20 30 40 Figure 4.41 Cross Correlation between probe (a) located at n=—O.888, and probe (b) located at n=-4. 16. HQ Az/B \. 01L11l111111111111111111111111olw111111111l11114 -5 -4 —3 -2 -1 O 1 2 3 4 71 Figure 4.42 Spanwise correlation of velocity at x/80=101. 4— Az/O 1 — ‘r—-~~, “x_ _fif/L—I \__\ ,\ r/ M\ - mo 60 -\ _0.80 0.70k lLIIALJlLAllll- Jl llJL lT—ll O -5 -4 -3 -2 -1 o 1 Figure 4.43 Spanwise correlation of velocity at x/0(,=650. I83 1.00 szwz 0.80 —‘ I. , 0.60 ~ i Velocity(u) 0.40 0.20 v 0.00 -O.20 1 I 1 I 1 I 1 I L I l O l 2 3 4 5 6 t* Figure 4.44 Autocorrelation of spanwise and lateral components of vorticity l.()() \ ' Velocity (u) ().80 ().()() t— -_-. 384 __ 675 ().40 l ().2() ’- ().()() _().20 1 1 114111I 1 1 111111I 1 1 111111l 1 1 1 1 -3 -2 - l() 10 H) 10 Figure 4.45 Autocorreltion of spanwise vorticity in semilog format I84 ().()50 ().()30 R (1.1110 »~ (”sz -().()I() I -().()3() I -().()5() 0.0 Figure 4.46 Close view of the 2.0 4.0 6.0 8.0 t* zero crossing region of the (DZ autocorrelation. I-00 I I CT] I I 7 I I I I I I I Ruu(t*) Ruu(Az/6) 0.50 ”—' ‘1’}! R11 .(t*) ’ (1)/, \S \ .. \9 0.00 C -w a." Rmszmz/e) _050 1111l 1111I 1111I 10'2 10" 10“ 10 t*, AZ/G I0.0 Figure 4.47 Spanwise correlation of u and (oral 1’1/UU = 0.5 Autocorrelations of (11)Z and u are included for comparison. 185 ().l()() Raglan 0.050 I (1.11110 ~ -().()5() » 4 -().|()(I -4II Figure 4.48 Correlation between high speed irrotational velocity fluctuations and (1)Z at n=0. -3() -2() -I() I865 () 41( .n.‘ I 0" E11(k1) 105 1/4 5/4 '04 8 V 103 lll“ lllll‘ llll‘ llllll" lllll‘ l l llll‘ l l lll'll‘ Wlll" lllllq I Lit Model Spectrum _______ Measured Spectrum J 111111l 1 1 1111111 1 1 1L11111 1 1 1 111111 10“ 10' 10'2 10" 100 1(an Figure 4.49 Measured and modeled one dimensional energy spectra: El 1(k1) E22(k1) '05 1/4 5/4 8 V ll'll‘ Tlllll‘ lllllII‘ llllll‘ lll l llll'l‘ l l llll‘ l l lll" l Tlll" lTllll'q 1 ll LA 10- Model Spectrum ....... Measured Spectrum 1(an Figure 4.50 Measured and modeled one dimensional energy spectra: E22(K1) I87 '06 f l l lllll' T l l llllll l l lllllll l l Tllllll l l l llllll l E3309) '05 10 1/4 5/4 3 8 V lll I II' I. l llllll‘ ’ I0’ lllllll‘ lllllll' 10 10' 10 10“ Model Spectrum _______ Measured Spectrum -2 I 0 10'3 j lllll‘ rnnrq llllll" l lllllI‘ lllm‘ l lllll‘ -4 1 1 1 11111I 1 l 1 llllll l 1 1111111 1 1 L111111 1 1 1111111 1 10‘4 10‘3 10'2 10'1 10° 1(an IO 5. til Figure 4.51 Measured and modeled one dimensional energy spectra: E33(K1) 0.3()0 f Y ffi v v v v v r v v l fi ‘ v v Y Y V Y I 0.250 _ .4 . ' o . 0.200 - - H12 . . . 0.150 _ 4 C . - 0.100 — o . .1 O .. 0.050 -— . . .. .0 I. O ().()()() b I ‘ ‘ ‘l L ‘ #1 1 L441. . . 1. .EW ‘0'} 10'2 104 km Figure 4.52 Spectral coherence function H12. Note the km rage is changed from previous figures. I88 2 _ '0 ,5» N x/00=707 1 .. .3. 10' x/90=384 10° Model Spectrum / 10" I I0-2 _6 111115 1 1.4 1 1“ “nu-2 1 IO 10‘ 10 IO‘ 10 10 lelK Figure 4.53 Autospectral function of vorticity. Model spectrum is from isotropic relation tOE“. 2 2 k¢wzm /un ) ().()5 0.04 0.03 0.02 0.0] 0.00 1 x/60=707 :1 ' \ ~ . k. I 108(km K) 0.0 Figure 4.54 Compensated vorticity spectra to show energy contribution to enstrophy inte- gral. I89 Pt 7) 0.00 Y -0.50 -1 .00 >- -I .50 I -2.()() I - ' - r - ‘ 1 . - ‘ - '1" ' P( '10 0.00 - '———— 4‘ Y 0.50 . ‘ -100 W 3 > 1 -150 ~ -200 - - - . - - - . - . - Nanci”! I I. _ PCY) 0.00 -0.50 Y -100 4.50 -2.00 r Figure 4.55 Schematic examples of yfor: (a) a thin uniform sheet of vorticity, (b) a sinuso- idally perturbed shear layer, (c) a rolled up shear layer into Gaussian vortices. The PW) curves on the right show a likely pdf that would be experimentally recovered for each con- dition with Gaussian noise. I90 ——9 —me m—a —_—> ‘0.0....IOOOOOOICCCOOOOOCOOO0.008 C3 1),.CIOOOOCCCOOOOOOIOOOOOOO.0.-COOS? IIITIIIIITITITITIITTIIT Figure 4.56 Schematic of showing the finite area of integration for the calculation of y Lght Sheet / Comero l Feud OF \Hew § . Mean Veuxflty H PPOFHe Area 0? Integroflon Comero 8 FEfld OF \New L. Loser Figure 4.57 Experimental configuration for 7 measurement I9I Number of realizations Figure 4.58 Histogram of measured 7 values. Note that the measured vales have been cor- rected by a factor of 1.19 to account for the limited field of view; see Section 4.7.2. 0.03 (02* F Bruns (1991) O 0.02 ~— IJOUCkS ( I008) resent meaSUI‘émenIS ‘1’ Balint and WaIIace 0.01 k ,«v g, “Ms" [/4/ 1,4111 1' I91“ “\I 0.00 1 1 1 1 , 1 1 2.0 2.5 3.0 3.5 4.1) 4.5 5.0 5,5 6.0 Log(Re) Figure 4.59 Peak vorticity measured by various researchers. I92 113:1: Thhrc semed inge Ther how enUe thes fine dd. mot dus gut thn It is mm 0101 Appendix A This appendix introduces a final conclusion which is not given in the body of this thesis. This result is related to the mechanism of entrainment, and is not central to the results pre- sented in section 4.8. It is therefore given here as a final point of interest. The conclusion is given as the following. The phenomenon of entrainment, defined as the movement of fluid towards the shear layer from the low speed side, would exist even if the fluid viscosity were zero. It is supposed that the observed magnitude of the entrainment would not change if the fluid viscosity were zero. The motivation for arriving at this conclusion originated from the lack of understanding of how fluid is entrained into the shear layer. One concept is verbalized by the term “viscous entrainment.” This implies the high speed fluid acts to “drag along” the fluid adjacent to the shear layer through the viscous diffusion of momentum. A second, and distinctly dif- ferent concept of entrainment is termed “engulfment” by Roshko (I979). This is an invis- cid entrainment mechanism in which the unsteady pressure field created by the large scale motions act to move the entrained fluid into the shear layer. As stated in the previous con- clusion, fluid is also extrained from the shear layer to the high speed side. The analysis given in the following paragraphs will seek to clarify physical mechanisms that lead to this asymmetrical movement of fluid. It is noted that Dimotakis (I986) utilized a model of point vortices to explain an inviscid mechanism of entrainment. Although the predictions of the entrainment rate given by this model are in agreement with many experimental results, the argument is heuristic in I93 nature, and is too simple to describe many other features of high Reynolds number shear layers. The present conclusion and the following discussion are not given to support any particular model, but simply to provide a solid argument that the mechanism which causes entrainment is an inviscid one. The proof of the conclusion stated above begins with the results of stability theory. Specif- ically, it has been shown by several authors (see. for example Michalke (1965)), that the inviscid, linear stability theory implies that the shear layer is inherently unstable and that this instability will lead to disturbances which grow spatially. Huerre and Monkewitz (1985) have provided a general result for shear layer instability based on the velocity ratio AU/ZU where AU is the velocity difference between the two streams and U is their aver- age velocity. The results of Huerre and Monkewitz (I985) have shown that for values of AU/ZU>I .3 15, the linear instability mechanism will lead to a temporally growing shear layer; values of AU/2(7<1.315 will lead to a spatially growing shear layer. For the present case (AU/2 0:1), this result implies that disturbances will undergo spatial growth. The nonlinear growth that results from this will lead to a nonzero value of dex. Equation 4.4 therefore requires that the entrainment is nonzero, just as equation 4.8 requires that the shear layer is tilted at a non—zero angle with respect to the constant veloc- ity free stream. It is concluded from these statements that if a shear layer could exist with a fluid which has zero viscosity. the shear layer would be convectivly unstable. The resulting spatial growth would lead to entrainment. However, this does not imply that the observed rate of entrainment (020.035) would be unchanged by the hypothetical case of zero viscosity, as I94 I!" I, 1 _.1 indicated by the supposition above. The following arguments will support the notion that entrainment is an emir'lv inviscid phenomenon. The first argument is given by the observation that the growth rate is linear. That is, dex is not a function of the Reynolds number, and therefore neither is the entrainment velocity. It is not correct to literally take this limit to Rezoo. because the Navier-Stokes equations become singular. However, the physical implications of the singular nature of the momen- tum equation are usually only important near solid walls where the boundary conditions must be applied. In summary, since the value of the Reynolds number does not effect the velocity of entrainment, it can be inferred that the phenomenon responsible for entrain- ment is not effected by the viscosity of the fluid. In addition to supporting the inferences above, additional insight can be obtained by observing the mechanism of viscous entrainment in two model flow fields. First, the idea of viscous entrainment can be studied from the “scraping plate” problem. This is described as steady flow of a viscous fluid in the upper half plane (y>0), with the boundary condi- tions (u,v)=(U0,0) at y=0, xe (—oo,oo). The entrainment rate is found to be ve = v(x, oo) ~ A. This solution shows that a viscous boundary conditions with constant velocity leads to an entrainment field which is dependent on the streamwise coordinate. The second idealized flow considered is the “stretching plate” problem. The boundary conditions are similar to the scraping plate problem. except the wall boundary condition is accelerating spatially at a rate a. That is, u(x,o)=ax, v(x,0)=0. The solution of this problem for large distances from the wall yields the entrainment rate: vc = v(x, oo) = x/av, where v is the kinematic viscosity of the fluid. In this solution, the entrainment velocity is a con- I95 stttttl. as I the work observed That is. t In sumn through morem ena. wl scale 1 the re oi the stant, as in the actual shear layer. It is then instructive to note that if air were to be used as the working fluid, the acceleration parameter that would be required to generate the value observed in the present physical shearlayer: VC 2 0.035 U0 = 0.25m/s, is a=4I60m/s/m. That is, the wall velocity would be: u(x=l,0)=4l60m/s. u(x=2,0)=8320m/s, etc. In summary, these model flows indicate that the induction of fluid towards the shear layer through the viscous diffusion of momentum is not likely to be significant. Rather, the movement of fluid towards the shear layer from the low speed side is an inviscid phenom— ena, which is believed to be related to the unsteady pressure field associated with the large scale motions of the flow. In this way, the present conclusion serves to motivate several of the results that were described in the bode of this thesis regarding the large scale motions of the flow. I96 REFERENCES Adrian, R. J., C. D. Meinhart, and C. Tomkins. 2000. Vortex organization in the outer region of the turbulent boundary layer. Journal anluid Mechanics 422, 1—53. Ali, S.K., Klewicki, C.L., Disimile, P.J., Lawson, I.. Foss, J.F., I985, “Entrainment region phenomena for a large plane shear layer,” proceedings, Fifth symposium on Turbulent Shear Flows. Andreopoulos, Y., Honkan, A., 2001, “An experimental study of the dissipative and vorti- cal motion in tubulent boundary layers,” vol. 439, pp. 131-163. Antonia, R.A., Browne, L.W., Shah, D.A., (1987) “Characteristics of vorticity fluctuations in a turbulent wake,” J. Fluid Mech, vol. 189, pp349-365. Antonia, R.A., Shafi, H.S., Zhu, Y., (1996) “A note on the vorticity spectrum,” Phys. Flu- ids 8 (8). P2196. Antonia, R.A., Zhou, T., Zhu, Y., (I998) “Three-component vorticity measurements in a turbulent grid flow,” J. Fluid Mech, vol. 374, pp29-57. Balint, J.L., Wallace, J .M., 1988 “Statistical properties of the vorticity field of a two- stream turbulent mixing layer,” Advances in Turbulence 2, Springer-Verlag, NY. Batchelor, G K., ( I946) “The theory of axisymmetric turbulence,” Proc. R. Soc. Land. A 186, 480-502. Bejan,A., 1995, Convective Heat Transfer, John Wiley and Sons, New York. Bradshaw, P., 1966, “The effect of initial conditions on the dvelopment of a free shear layer,” J. Fluid Mech., vol. 26, pp25-236. Browand, F.K., Lati go, B.O., 1979, “Growth of hte two-dimensional mixing layer from a turbulent and nonturbulent boundary layer,” Phys. Fluids, vol. 22, No. 6, pp 1101-1019. Browand, F.K., Trout T.R., 1980, “A note on spanwise structure in the two-dimensional mixing layer,” J. Fluid Mech., vol ”7, pp77l-781. Browand, F.K., Trout T.R., 1985, “The turbulent mixing layer: geometry of large vorti- ces,” J. Fluid Mech., vol 158, pp489-509. Browand, F.K., Weidman, PD, 1976, “Largest scales in the developing mixing layer,” J. Fluids Mech. Vol 76, ppl27-l44. Browne, L.W.B., Antonia, R.A., and Shah, D.A., ( I987) “Turbulent energy dissipation in a wake,” J. Fluid Mech., vol. 179, pp307-326. I97 Brown, GL., Roshko, A., “On density effects and large structure in turbulent mixing lay- ers,” J. Fluid Mech., vol 64, pp775-816. Bruns, J., 1990, “Evaluation of hot-wire velocity and transverse vorticity measurements in a single stream shear layer and its parent boundary layer,” Diplomarbeit, Technische Hochshule Aachen. Bruns, J .M., Haw, R.C., Foss, J .F., 1991, “The velocity and transverse vorticity field in a single stream shear layer,” in the Eighth Symposium on Turbulent Shear Flows. Technical University of Munich, September 9-l l. Bruun, H.H., I995, Hot-wire Anemometry, Oxford University Press. Cao, N., Chen, S., Sreenivasan, K.R., I996, “Properties of velocity circulation in three- dimensional turbulence,” Phys. Rev. Let. Vol. 76, No. 4, pp 6l6-6l9. Champagne, F.H., Pao, Y.H., Wygnanski, I.J., I976, “On the two-dimensional mixing region,” J. Fluid Mech. Vol. 74, pp.209-250. Dimotakis, PE, 1986, “Two-dimensional shear-layer entrainment,” AIAA Journal, Vol. 24, No. II, Ppl79l-l796. Dimotakis, P., Brown, G, 1976, “The mixing layer at high Reynolds number: large-struc- ture dynamics and entrainment,” J. Fluid Mech., vol 78, pp535-560. Foss, J .F., 198 I , “Advanced techniques for transverse vorticity measurements,” Proceed- ings, 7th Biennial Symposium on Turbulence, Univ. of Missouri-Rolla. Foss, J .F., 1994, “Vorticity considerations and planar shear layers,” Experimental Thermal and Fluid Sciences, Vol. 8, pp. 260-270. Foss, J.F., 1998, Private communication. Foss, J.F., Haw, RC, 1990, “Transverse vorticity measurements using a companct array of four sensors,” In The Heuristics of Thermal Anemometry, ed. DE STock, SA Sherif, AJ Smits, ASME-FED 97: 71-76, PP98. Foss, J .F., Bohl, D.G, Kleis, 8.1., 1995, “Activity Intermittency in a two-stream shear layer,” Proceedings of the Tenth Symposium on Turbulent Shear Flows, The Pennsylvania State University, University Park, PA. Freymuth, P., I966, “On transition in a separated laminar bloundary layer,” J. Fluid Mech. Vol. 25, pp.683-704. Hama, F.R., I962, “Streaklines in a perturbed shear flow,” Phy. of Fluids, Vol. 5, No. 6, pp 644-650. I98 Haw, R.C., Foss, J.K., Foss., J.F., I989, “Vorticity based intermittency measurements in a single stream shear layer,” in Advances in Turbulence 2, eds. Femholz, H., and Fiedler, H., Springer-Verlag, Berlin. HO, O, Huerre, P., 1984, “Perturbed free shear layers,” Ann. Rev. Fluid Mech., vel l6, pp365-424. Holmes, Lumley, Berkooz, I992, Coherent Structures. Symmetry, Dynamical Systems and Turbulence. Cambridge University Press. George, W.K., Hussein, H.J., (1991) “Locally axisymmetric turbulence,” J. Fluid Mech. V0]. 233, pp- 1'23. Hamelin, J. Alving, A.E., 1996, “A low-shear turbulent boundary layer,” Phys. Fluids, vol. 8,No. 3 pp 789-804. Hileman, J., Samimy, M., 2001, “Turbulence Structures and the Acoustic Far-Field of a Mach 1.3 Jet,” AIAA Journal, Vel. 39, No. 9, ppl7l6-l727. Huerre, P., Monkewitz, P.A., 1985, “Absolute and convective instabilities in free shear layers,” J. Fluid Mech., Vol. 159, pp. 151-168. Hussain, A.K.M.F., Zedan, M.F., 1978, “Effects of the initial condition on the axisymmet- ric free shear layer: effects of the initial momentum thickness,” Phys. Fluids, Vol. 21, No. 7. Hussain, A.K.M.F., Zaman, K.B.M.Q., 1985, “An experimental study of organized motions in the turbulent plane mixing layer,” J. Fluid Mech. Vol. 159, pp.85-104. Kim, J.H., 1989, PhD thesis, University of Berlin. Klewicki 1989, PhD thesis, Michigan State University. Klewicki, J .C., Falco, R.E., Foss, J .F., 1992, “Some characteristics of hte vortical motions in the outer region fo turbulent boundary layers,” J. ofFluids Engineering, Vol. 114, pp. 530-536. Kock am Brink, B., Foss, J .F., 1993, “Enhanced Mixing via Geometric Manipulation of a Splitter Plate,” AIAA paper no. 93-3244, AIAA Shear Flow Conference, July 6-9, Orlando, FL. Kolmogorov, A.N., I941, “The local structure of turbulence in incompressible viscous fluid for very large Reynolds number,” translated in Turbulence and Stochasitc Processes: Kolmogorov’s ideas 50 years on, ed. Hunt, J.C.R., Phillops, O.M., and Williams, D., The Royal Society, London, 1991. I99 Koochesfahani, M.M., Catherasoo, C.J., Dimotakis, P.E., Gharib, M, and Lang, DB, (1979) “Two-point LDC measurements in a plane mixing layer,” AIAA J., Vol. 17, No. 12, ppl347-l351. Koochesfahani, M.M., Cohn, R., MacKinnon, C., 2000, “Simultaneous whole-field mea- surements of velocity and concentration fields using a combination of MTV and LIF,” Meas. Sci. Technol. Vol. II, pp 1289-1300. Lang, DB, 1985, PhD Dissertation, Cal. Inst. of Tech. Liepmann, H.W., Laufer J., I947, “Investigations of free turbulent mixing,” NACA Report No. 1257. Loucks, RB, 1998, “An Experimental Examination of the Velocity and Vorticity Fields in a Plane Mixing Layer,” Ph.D Thesis,University of Maryland. Lui, X., 2000, Ph.D. Thesis, Department of Aerospace and Mechanical Engineering, Uni- versity of Notre Dame. Lumley, J.L., (I965) “Interpretation of time spectra measured in high-intensity shear flows,” Phys. Fluids. 8, 1056-1062. Mehta, R.D., Westphal, R.V., 1986, “Near-field turbulence properties of single and two- stream plane mixing layers,” Experiments in Fluids, 4, pp. 257-266. Michalke, A., 1965, “On spatially growing disturbances in an inviscid shear layer,” J. Fluid Mech, vol. 23, pp521-544. Monin, A.S., Yaglom, A.M., 1975, Statistical Fluids Mechanics: Mechanics of Turbu- lence. Vol 2, The MIT Press, Cambridge. Naguib, A., 1992, PhD thesis, Illinois Institute of Technology, Chicago, IL. Narayanan, S., Hussain, F., 1996, “Measurements of spatiotemporal dynamics in a forced plane mixing layer,” J. Fluid Mech., vol 320, pp7 l -I IS. Ong, L., Wallace, J .M., 1998, “Joint probability density analysis of the structure and dynamics of the vorticity field of a turbulent boundary layer,” vol 367, pp. 291-328. Ong, L., Wallace, J .M., 1995, “Local isotropy of hte vorticity field in a boundary layer at high Reynolds number,” in Advances in Turbulence V, edited by R. Benzi (Kluwer, Dor- drecht), p. 392. Panton, R.L., 2001, “Overview of the self-sustaining mechanicsm of wall turbulence,” Progress in Aerospace Sciences, Vel. 37, pp. 341-383. 200 Pope, S., 2000, Turbulent Flows, Cambridge University Pres. Pozrikidis, C., Higdon, J.J.L., 1985, “Nonlinear Kelvin-Helmholtz instability of a finite vortex layer,” J. Fluid Mech., Vol. 157, pp. 225-263. Pui, N.K., Gartshore, 1.5., (1979) “Measurements of the growth rate and sturcture in plane turbulent mixing layers,” J. Fluid Mech., vol. 9l, ppl 1 l-l30. Rogers, M.M., Moser, RD, 1994, “Direct simulation of a self-similar turbulent mixing layer,” Phys. Fluids, Vol. 6, No. 2. Roshko, A., 1979, “Structure of turbulent shear flows: a new look,” AIAA Journal, Vol. l4, No. 10, PPI349-l357. Saddoughi, S.G., Veeravalli, S.V., I994, “Local isotropy in turbulent boundary layers at high Reynolds number,” J. Fluid Mech, vol. I89, pp333-372. Schmidt, T.P., Ali, S.K., and Foss, J.F., 1986, “Relative efficiencies for parallel and per- pendicular entrainment flow paths,” AIAA Journal, Vol. 24, No. 9. Slessor, M.D., Bond, C.L., Dimotakis, PE, 1998. “Turbulent shear-layer mixing at high Reynolds numbers: effects of inflow conditions,” J. Fluid Mech, vol. 376, pp115-138. Sreenivasan, K.R., Antonia, R.A., I997, “The phenomenology of small-scale turbulence,” Ann. Rev. Fluid Mech., Vol 29, pp. 435-472. Tennekis and Lumley, I972, A First Course in Turbulence, MIT Press. Thomas, F.O., I99 I , “Structure of mixing layers and jets.” App]. Mech. Rev, vol 44, No 3.. pp] 19-153. Townsend, I975, The Structure of Turbulent Shear Flow. Cambridge University Press. Wallace, J .M., Foss, J .F., 1995, “The Measurement of Vorticity in Turbulent Flows,” Annu. Rev. Fluid Mech. 27, pp469-514. Winant, C.C., Browand, F.K., 1974, “Vortex pairing: the mechanism of turbulent mixing layer growth at moderate Reynolds number,” J. Fluid Mech., Vol. 63, pp. 237-255. Wygnanski, I., Fiedler, HE, 1970, “The two-dimensional mixing region,” J. Fluid Mech. Vol. 41, pp.327-36l. Wyngaard, J. C., and Clifford SF, (1977) “Taylor’s hypothesis nad high-frequency turbu- lence spectra,” J. Atmos. Sci. 34, 922-929. 20I