MODELING AND MEASUREMENT OF SURFACE PRESSURE FLUCTUATION IN AN IMP INGING JET By Nasem A. Aukla A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Mechanical Engineering Master of Science 2018 ABSTRACT MODELING AND MEASUREMENT OF SURFACE PRESSURE FLUCTUATION IN AN IMPING ING JET By Nasem A. Aukla Imping ing jets are used in many engineering and industrial applications, including heating, cooling , drying, food processing , and s urface cleaning, among others . The present thesis work is focused on modeling and studying the unsteady wall - pressure signature produced by a jet impinging normally on a flat surface . This study is divided into two main parts : A theoretical part, to establ ish the beginning step towards building a physics - based, mathematical model to calculate the surface - pressure fluctuation on the impingement surface. The mathematical model is used to explore the effects of changing the flow and jet - vortices parameters, on e at a time, on the characteristics of the surface - pressure fluctuation. Three main parameters are examined: vortex - passage frequency, jet R eynolds number, and vortex circulation. An experimental part, to measure the unsteady surface pressure fluctuation on the impingement surface for an axisymmetric jet at normal incidence. Measurements are done, for Reynold numbers 8272 and 24818 (based on the jet diameter ( ) and jet exit velocity), using a microphone array extending radially from the stagnation point ( ) into the wall - jet zone ( ). Comparison of the model and the experimental results sho ws that, despite of the model simplicity, certain qualitative features of the unsteady wall pressure are similar within the stagnation zone. This outcome establishes confidence to continue further development of the model in the future. Copyri ght by NASEM A. AUKLA 2018 iv DEDICATED TO THE MEMORY OF MY FATHER AND MOTHER v ACKNOWLEDGMENTS I would like to sincerely acknowledge my family , advisor, committee members, friends, and colleagues, because without their help and support this work could not have been done . Many people have helped me to come at this point, and I cannot go further before expressing my most profound gratitude to them. Firstly, to Prof. Dr. Ahmed Naguib for giving me an opportunity to be one of his students and to work on an exciting and challenging topic. His advice, direction, and encouragement were much supportive of my thesis. My profound gratitude to my committee members Prof. Dr. Farhad Jaberi and Prof. Dr. Ricardo Mejia - Alvarezwhos e for being on my committee . Many thanks to Dr. K e Zhang who helped me with my experimental setup. My profound gratitude to Dr. Sami Al - Araji, Mr. Dhia Al - Ashaqr, and Mr. Nazar Hussain who helped me to achieve my goal, without their support I would not h ave accomplished this work. Finally, Many thanks to my colleague and best friend, Fady Hindo. The honest and purest gentleman who I met in my entire life. Thank you for your tru e friendship. vi TABLE OF CONTENTS . . ..v iii . ... i x KEY TO SYMBOLS .. . xi v CHAPTER 1: Introduction ... . .1 1.1 Free Jet s ..1 1.1.1 . 1 1.1.2 Literature Review 1.2 Impinging 4 1.2.1 . ... ...4 1.2.2 Literature Review . ...5 1.2.3 . 9 1.3 . 12 1.4 Objective s .. . 13 CHAPTER ... 15 15 17 2.2.1 17 2.2.2 Wall (Vortex Rings Images) ... 19 2.2.3 Advection Field (Stagnation Point Potential Flow) ... 22 2.2.4 Vortex Advection and Flow - 22 2.2.5 Wall - . 26 28 2.3.1 Computation of the Time - Depend ent Stream Function and Velocity Field 29 29 30 2.4. 1 Impinging - .. 31 31 2.4.1.2 Validation of the Wall - Pressure - Grid Resolution 2.4.1.3 Validation of the Time Step Size 37 2.4.2 Application to A Line Vortex Above A Flat Wall CHAPTER 4 5 49 .. . 64 3.3 Circulation Effect .. . . 66 CHAPTER .. ... 7 1 4.1 .. 71 .. 72 vii 4.3 Hot - wire Setup and Calibration 76 83 89 4.6 Procedure to Acquire the Microphone Signal s 89 CHAPTER 5: Jet Characteristics and Wall - .. 90 5.1 Initial Shear Layer Self - Similarity 90 94 5.3 Time ... . 99 . 99 103 107 5.4 S 112 5.5 Comparison between Experiment and Mathematical Model ... . 123 CHAPTER 6: Conclusion s and Recommendation s . 1 32 .137 . . 142 viii LIST OF TABLES Table 2.1 Comparison of high - fidelity and present model features .. 16 Table 3.1 Model parameters for frequency , Reynold number , and circulation ... 49 Table 4.1 Sample of the microphones sensitivity .. ix L IST OF FIGURES Figure 1.1 Schematic of a free - jet flow, depicting various flow development zones ..2 Figure 1. 2 Schematic of the impinging jet at normal incidence (taken from Al - Aweni [7] 5 Figure 2.1 Schematic drawing of the full mathemati cal model of the impinging jet 18 Figure 2. 2 Radial wall - pressure distribution at the time instant when the negative pressure spike (minimum pressure) associated with vortex passing is located at r=5 cm location. Different color lines represent differ ent radial domain size s used for computing the wall pressure (as given by the legend). The inset shows the overall distribution, while the main plot depicts a magnified view to show the change between the different cases. 32 Figure 2. 3 N ormalized Root Mean Square Error ( NRMSE ) resulting from comparing the results of domain sizes 10 and 15 cm (blue) and 15 and 20 cm (red). The results are shown up to the computation time corresponding to the time instant where the negative pressure spike a ppears at r = 5 cm . 33 Figure 2. 4 Radial wall - pressure distribution at the time instant when the negative pressure spike (minimum pressure) associated with vortex passing is located at . Different color lines re present a different number of wall grid points used for computing the wall pressure (as given by the legend). The inset shows the overall distribution, while the main plot depicts a magnified view to show the change between the different cases. 3 5 Figure 2. 5 Normalized Root Mean Square Error ( NRMSE ) resulting from comparing the results of grid resolution 300 and 600 points (blue) and 600 and 1200 points (red). The results are shown up to the computation time corresponding to the time i nstant whe n the negative pressure spike appears at 36 Figure 2. 6 Radial wall - pressure distribution at the instant when the negative pressure spike (minimum pressure) associated with vortex passing is located at . Different color lines represent a different number of computational time steps in the period between injecting two successive vortices (as given by the legend). The inset shows the overall distribution, while the main plot depicts a magnified view to show the change between the different cases 38 Figure 2. 7 Normalized Root Mean Square Error ( NRMSE ) resulting from comparing the results of time step 4.066e - 06 s (1000 PPC ) and 2.033e - 06 s (2000 PPC ) (blue) and 2.033e - 06 s (2000 PPC ) and 1.0161e - 06 s (4000 PPC ) (red). The results are shown up to the computation time corresponding to the time instant whe n the negative pressure spike appears at 39 Figure 2. 8 Schematic of a line vortex centered at ( x o , y o ) above an infinite flat wall at y = 0. An image vortex centered at ( x o , - y o ) models the presence of the wall 41 x Figure 2. 9 Computational and analytical results of the wall pressure and associated streamlines for a line vortex above a wall. Different subplots represent different times (as indicated beneath the plots). The streamlines are shown in a frame of reference convecting with the vortex .. 43 Figure 2. 10 N ormalized Root Mean Square Error (NRMSE) between a nalytical and numerical values of the wall - pressure distribution beneath a line vortex above a flat wall. The time duration shown is from the start of the vortex motion until the vortex center is located at r 44 Figure 3.1 Streamlines and normalized pressure at different time step s for frequency , Re =5000, and circulation 51 Figure 3.2 Stremlines and normalized pressure at different time step for frequency f = 980 Hz , Re =5000, and circulation 52 Figure 3.3 Effect of vortex passage frequency on the radial distribution of the mean pressure coefficient . The green broken lines outline the boundary of the sta 54 Figure 3.4 Effect of vortex passage frequency on the radial distribution of the RMS pressure coefficient 56 Figure 3.5 Pressure time series at various r/D locations within the stagnation zone, Re =5000, , and . The small discontinuities in the time series for correspond to the time instant when a new vortex is injected at the top of the computational domain. Because the d iscontinuities are small, no attempt was made to get rid of them by using a larger wall - normal extent for the computational domain 58 Figure 3.6 Pressure time series at various frequencies , Re =5000, , and . The small discontinuities in the time series for the two larger frequencies correspond to the time instant when a new vortex is injected at the top of the computational domain. Because the disc ontinuities are small, no attempt was made to get rid of them by using a larger wall - normal extent for the computational domain. .62 Figure 3.7 Induced v elocity by an infinite line - vortex array along a line parallel to the array. Results are sho wn over one vortex spacing ( ) with the vortex located at the center ( ). 63 . Figure 3.8 Individual and collective induced wall velocity by three - line vortices placed relative to a wall a - frequency case when the leading vortex is located at ..64 Figure 3.9 Effect of Reynolds number on the radial distribution of the mean pressure coefficient . 65 Figure 3.10 Effect of Reynolds number on the radial distribution of the RMS pressure coefficient .66 xi Figure 3.1 1 Effect of vortex circulation on the radial distribution of the mea n pressure coefficient 67 Figure 3.1 2 Effect of vortex circulation on the radial distribution of the RMS pressure coefficient Figure 3.1 3 Pressure time series at various r/D locations within the stagnation zone, Re =5000, , and 69 Figure 3.1 4 Pressure time series for various vortex circulation s , Re =5000, , and ..70 Figure 4.1 F low configuration for normal - incidence impingement jet 72 Figure 4.2 Experiment al facility 73 Figure 4.3 An image of the impingement plate and the sliding ta ble attached to the Velmex manual traverse system 74 Figure 4.4 Block diagram of the setup used for calibration and measurements of the hot - wire 76 Figure 4.5 An image of the motorized system for traversing the hot wire 77 Figure 4.6 Block diagram of the setup for the square wave test 79 Figure 4.7 An image of the oscilloscope screen showing a typical square wave test result 79 Figure 4.8 An image of the h ot - wire, temperature sensor, nozzle, impingement plate, and the conditioning box .81 Figure 4.9 Sample of hot - wire calibrations before and after an experiment 83 Figure 4.10 Front and cross section ( B - B ) view (top and bottom respectively) of the microphone array configuration used in the present work 84 Figure 4.1 1 Block diagram of the calibration setup for the microphone array 86 Figure 4.1 2 An image of the calibration setup for the microphone array .87 Figure 4.1 3 A sample of the microphone calibration results: sensitivity (top) and phase (bottom) 88 Figure 5.1 Shear - layer mean velocity profile at X/D =0.2 for various Reynolds num bers .92 xii Figure 5.2 Shear - layer fluctuati ng - velocity root - mean - square profile at X/D =0.2 for various Reynolds numbers 92 Figure 5.3 Shear - layer mean velocity profile at various X/D locations and 93 Figure 5.4 Shear - layer fluctuati ng - velocity root - mean - square profile at various X/D locations and .94 Figure 5.5 Normalized Root Mean Square Pressure versus for different and two Reynolds numbers: (a) Re D = 8272, and (b) Re D = 24818 96 Figure 5.6 Normalized Root Mean Square Pressure versus at Re D = 8272 and 24818 , for : and .98 Figure 5.7 Sample normalized Pressure signals at Re D = 8272 , for for: and 100 Figure 5.8 Sample normalized Pre ssure signals at Re D = 24818 , for for: and 102 Figure 5.9 Probability Density Function for the pressure signal at Re D = 8272 , for , and: and .. .105 Figure 5.10 Probability Density Function for the pressure signal at Re D = 24818 , for , and: and 106 Figure 5.11 Radial distribution of skewness for and and: , and 109 Figure 5.12 Radial distribution of Kurtosis for and and: , and ..111 Figure 5.13 Normalized PSD contour plots at for: , and .. 115 Figure 5.14 Normalized PSD contour plots at for: , and .117 Figure 5.15 Normalized PSD at for , and at: 118 Figure 5.16 Normalized PSD at for and at: .119 Figure 5.17 Normalized PSD at for: and .121 xiii Figure 5.18 Normalized PSD at for: and ..122 Figure 5.1 9 Samp le time series from microphone measurements (top) and the vortex - array model (bottom) at the stagnation point ( ) and end of the stagnation zone ( ). Experimental data are shown for the and , and model results for the refe .. 125 Figure 5. 20 Sample wall - pressure signature characteristic of that produced by vortex - induced separation from microphone measurements at , and 126 Figure 5. 21 Comparison of the effect of varying on RMS wall - pressure fluctuation between the experimental (top) and the model (bottom) results at . In the model, increasing frequency corresponds to decreasing . The br oken green lines outline the end of the 128 Figure 5. 22 Comparison of the effect of varying on RMS wall - pressure fluctuation between the experimental (top) and the model (bottom) results at (experiment) and (model) . The broken green lines outline the end of the stagnation zone ..131 xiv KEY TO SYMBOLS Hot - wire calibration coefficient s a Constant in the potential stagnation - point flow eq uations C Volumetric constant related to the volume of the vortex ring c Sound speed Mean - pressure coefficient Root - mean - square - pressure coefficient D Jet diameter E Hot - wire output v Temperature - c orrected h ot - wire output voltage Complete elliptic integral of the second kind Measured h ot - wire voltage Strain - rate tensor f Frequency H Distance between the impingement plate and the jet exit Complete elliptic integral of the fir st kind l Side length of the plane wave tube N Number of samples in a time series P Mean wall pressure D ynamic pressure of the jet PSD Power spectral density of the wall - pressure fluctuation Wall pressure xv Wall - pressure fluctuation Root mean square of the wall pressure fluctuation q Flow s ource of pressure Jet Reynolds number based on the jet diameter and the jet exit velocity Vortex core radius Initial vor tex core radius r, Impingement - plate polar coordinate s Strouhal number based on jet diameter Temperature of the fl ow during h ot - wire calibration Temperature of the fl ow during h ot - wire measurements Hot - w ire temperature Radial component of velocity induced by the image on the real vortex ring Dimensionless radial component of velocity induced by the image on the real vortex ring Normal component of velocity induced by the image on the real vortex ring Dimensionless normal component of velocity induced by the image on the real vortex Ring Radial component of se lf - induced velocity Normal component of self - induced velocity Dimensionless normal component of self - induced velocity Jet exit v elocity Radial v elocity component (parallel to the plate) Axial v elocity component (normal to the plate) Azimuthal v elocity component xvi Initial coordinates of the vortex ring in polar coordinate s Initial coordinates of the vortex in Cartesian coor dinate x, y, z Cartesian coordinate system with origin at the center of the impingement plate C artesian coordinate s of the pressure source t Time Velocity p otential Rotation rate tensor Vortex Circulation Circulation around a contour that encompasses the shear laye r V orticity Wavelength Momentum thickness of the jet shear layer Fluid density Fluid dynamic viscosity Fluid kinematic viscosity Velocity difference across t he shear layer Stream function Dimensionless stream function 1 CHAPTER 1: Introduction Experimental, computational and analytical research has been used to study impinging jets for the past decades because of their significance t o many applications, including cooling, heating, drying, air conditioning, and ventilation. The present research is concerned with modeling , and understanding the physics of wall - pressure generation in impinging jets. This knowledge is significant for appl ications involving flow - induced noise and vibration. To motivate the current research, a summary of relevant previous literature is necessary. However, some understanding of the basic characteristics and flow features of free (non - impinging) jets is essent ial. Therefore, the present discussion starts with an overview of the latter. This is followed by a brief summary of the relevant work on impinging jets and their wall - pressure characteristics. Finally, the motivation and the specific objectives of this th esis are outlined at the end of the chapter. 1.1 Free Jets 1.1.1 Background Jets can be classified into different categories according to: the nozzle exit shape (e.g. circular, square, triangular, lobed, etc.), the nozzle contour (e.g. smooth versus sharp - edge d), the pipe length from which the jet emerges (long or short; if any), and the initial discharge condition ( free, wall and surface jet). The most basic type of jets is the axisymmetric free jet, driven by pressure to emerge at the end of a contoured nozzl e into a quiescent ambient. For a free jet, after exiting the nozzle, a free shear layer with uniform pressure surrounds the jet . As the flow develops farther downstream, mixing of the jet fluid with the ambient causes an increase in the mass flow rate in the jet stream (i.e. via flow entrainment), and, in conjunction with viscous and turbulence effects, leads to jet spreading and decreasing of the flow speed to conserve momentum. 2 Depending on the characteristics of the jet development may be divided into three zones as depicted in Figure 1.1 . see Shih - I Pai [1] Zone 1: in this zone , there is a potential core in the central part of the jet and a mixing zone is sandwiched between the potential core and the surrounding me dium. The potential core has a centerline velocity equal to the jet exit velocity . This zone extends up to 4 6 jet diameters ; Zone 2: represents the transition where the velocity profile gradually changes until self - similarity is established . This zone e xtends from the end of zone 1 to 20 jet diameters ; Zone 3: represents the self - similarity zone where the transverse mean velocity profile is similar at different axial distances when normalized using the local centerline velocity and jet width. Figure 1.1 Schematic of a free - jet flow, depicting various flow development zones 3 1.1.2 Literature Review The primary and fundamental interest in studying free jets has been focused on the instability of the shear layer downstream from the nozzle exit, and the t urbulence development farther downstream. Grant [2] theoretically studied the shear layer instability in axisymmetric jets . This study shows the formation of vortex structures from an instability wave originating at the beginning of mmetric shear layer. The initially weak instability wave amplifies with downstream distance, ultimately leading to creation of the eddies. Popiel and Trass [3] visualized free and impinging round jets using a smoke wire . For the free jet, t hey showed the vortices to form in the potential - core region within the free shear layer. The downstream merging of these vortices causes the creation of larger eddies. These vortices increase mixing and enhance the entrainment rate . Also, the axial symmetry of the near field of the nozzle exit is found to be created by the generation of the toroidal vortices, because they cause a significant upstream interaction . Zaman and Hussain [4] studied the natural large - scale structures in the axisymmetric mixing layer surrounding the jet. They found the flow from the jet exit to the end of the potential core to be controlled by two characteristic length scales: the jet diameter and the initial shear layer thickness. Near the jet exit, the momentum thickness controls the flow struc ture. Initially, the momentum thickness is thin, and as the shear - layer rolls - up into vortical structures, the momentum thickness grows . The vortices resulting from the roll - up start to interact and create larger eddies. After a distance x , which is compar able to the jet diameter, the diameter length scale controls the flow structure. As the jet Reynolds number ( ; where is the jet exit velocity, the jet diameter, and the fluid kinematic viscosity) increases, the ratio of the shear layer thickness at the nozzle lip to the diameter of the jet decreases, which leads to a smaller in itial instability wavelength, relative to the jet diameter, and larger number of vortex 4 pairings before the vortices size becomes comparable to the jet diameter, at the end of the potential core. 1.2 Impinging Jet 1.2.1 Background In an impinging jet, the f low is incident on a, typically, flat wall at a distance H from the jet exit. Though the angle of incidence of the jet relative to the wall - normal direction may change, the present work is only concerned with normal incidence; i.e. where the jet symmetry a xis is perpendicular to the wall. The flow field for an impinging jet may be divided into three zones (see Figure 1. 2 ): Free - jet zone: which represents the domain stretching from the nozzle exit to the point where the existence of the plate does not infl uence the flow. Within this zone, the jet flow and associated flow features are as discussed in the previous section; Stagnation zone: which represents the domain where the mean flow direction changes from being normal to being parallel to the plate. This zone, which extends up to in the radial direction, has the maximum mean pressure of the flow ( at the stagnation point on the wall ) ; Wall - jet zone: w hich corresponds to the domain . Unsteady separation of the boundary layer is known to o ccur in this region due to the interaction of the jet vortices with the wall; See Didden and Ho [5] and Landreth and Adrian [6] . 5 Figure 1. 2 Schematic of the impinging jet at normal incidence (taken from Al - Aweni [7 ] ) 1.2.2 Literature Review A brief revi ew is provided here of the current knowledge on wall - pressure fluctuation in impinging jets. Similar to the current work, focus will be on H/D values extending to the end of the potential core ( ). To characterize the strength of the pressure fluctuation, typically the radial distribution of the root mean square of the fluctuation ( ) is examined. The recent work of Krishna [8] provides the most comprehensive data, co vering the Reynolds number range , and . Krishna showed that for , the largest pressure fluctuation are found in the wall - jet zone in the range for all Reynolds numbers. A second peak within the stagnati on zone emerged, near , when reached 25,000 and the impingement wall was sufficiently far from the jet ( ). When present, this secondary peak magnitude was higher than that in the wall - jet zone. 6 Consistent with the work of Krishna [8], several earlier studies have identified the wall - jet - zone peak of the pressure fluctuation. These include the work of Hall and Ewing [9] and [10], who found the peak at for and , El - Anwar et al. [11], at for and , and Al - Aweni [7], for and . The variation in the exact location of the peak between the different studies might be due to Reynolds number, the jet initial condition, or t he radial spacing between the measurements. The RMS level of the pressure fluctuation associated with this peak is quite large, dynamic pressure (based on jet exit velocity), but this level, along with the overall level of pressure fluctuation in the wall - jet zone, decrease with increasing both and . This Reynolds number trend is seen in the data of Krishna [8], while the dependence is reported in all aforementioned studies. The peak in the stagnation zone is most clearly seen in the data of Krishna [8], which e xtend into a high - enough range for the peak to be observed. The presence of this peak is also implied in the data of Hall and Ewing [9 , 10] at , although there were not sufficient measurement points to ascertain the specific locatio n of the peak inside the stagnation zone. Furthermore, Hall and Ewing employed a jet that exits at the end of a fully - developed turbulent pipe flow, which exhibits significant pressure fluctuations at the stagnation point (relative to jets exiting from a c ontoured nozzle) due to the absence of a potential core. Within the stagnation zone, increases with increasing ; opposite to the trend in the wall - jet region. This behavior is generally associated in the literature with the growing influence of the vortical structures on the stagnation zone as the vortices grow to a size co mparable to the jet diameter through successive pairings (e.g. see Al - Aweni [7]). On the other hand, the 7 Reynolds number influence is found to decrease the stagnation - zone pressure fluctuation (Krishna [8]). No specific explanation for this trend is known at this point. Frequency spectra analysis of the wall - pressure time series measured in the studies referenced above show that the frequencies of the spectral peaks are consistent with the passage frequency of the jet vortices. In addition to depending on t he initial conditions and Reynolds number of the jet, the latter frequency is predominantly affected by the distance between the jet exit and the impingement plate. As the distance increases from H/D= 2 to the end of the potential core, the number of vortex ring parings ahead of impingement increases, decreasing the passage frequency of the vortices. Al - Aweni [7] used simultaneous time - resolved flow visualization and wall - pressure measurement, employing a microphone array , for a jet Reynolds number of 7970, and H/D= 2,3, and 4. He showed that for H/D= 2 , the first vortex merging occurred within the wall - jet zone, as the vortices traveled parallel to the impingement wall. For , the first pairing was complete d before reaching the impingement plate, while f or H/D=4 , the second merging took place ahead of the plate. When pairing happened ahead of reaching the wall, each merging resulted in halving the fundamental frequency in the pressure spectra. When merging happened while the vortices traveled past, and in teracted with the wall ( ), the spectrum contained the original vortex formation frequency (or Strouhal number ) and its sub - harmonic ( ). The lowest frequency observed, at after two pairings, corresponded to . The drop of the dominant pressure - fluctuation Strouhal number with and the overall order of magnitude of the Strouhal number values reported in Al - Aweni [7] is consistent with the findings of Hall and Ewing [9] and [10], El - Anwar et al. [11] and Krishna [8]. The ea rliest explanation for the large pressure fluctuation associated with the peak in the wall - jet zone came from the work of Didden and Ho [5]. These authors studied a normally 8 impinging axisymmetric forced air jet at Reynolds number of 19000, and H/D= 4. Using phase averaged pressure and hot - wire measurements, they showed that the high level of pressure fluctuation in the wall - jet region is associated with the unsteady boundary layer separation, and subsequent formation of a n opposite - signed second ary vortex, when a jet (primary) vortex interact ed with the wall. Such an unsteady boundary - layer separation process , and associated phenomena, was first noted by Harvey and Perry [12] in relation to trailing wing - tip vortices interacting with the ground. In impinging jets, formation of secondary vortices was also reported in the work of Landreth and Adrian [6] , for Reynolds number of 6500 using particle image velocimetry, and the flow visualization of Popiel and Trass [3]. Didden and Ho [5] showed that the boundary layer separation was associated with the adverse pressure gradient imposed by the jet vortices on the boundary layer . Naguib and Koochesfahani [13] used whole - field velocity data of an isolated axi - symmetric vortex ring interacting with a flat wa ll to understand the fundamental surface - pressure generation mechanism associated with vortex - - pressure generating sources and wall - pressure signature from the velocity - field data. In addition to connecting negative wall - pressure peaks with the primary and secondary vortices, they were able to identify an important source of positive pressure fluctuation, not known before the n. Specifically, they showed that the high strain zone associated with the separation of the boundary layer (induced by the main vortex ring) was an important source of positive pressure fluctuation. The results , however , did not show a strong negative pre ssure peak, as was found in the work of Didden and Ho [10] beneath the separated zone. This may have been due to the limited spatial resolution of the experimental data within the separating boundary layer. 9 Al - Aweni [ 7 ] conducted a comprehensive study of the wall - pressure fluctuation and associated generation mechanisms in impinging jets. The work utilized both experimental data (from simultaneous time - resolved flow visualization and wall - pressure sensor - array data) and axi - symmetric laminar CFD calculatio n of isolated vortices interacting with a flat wall. An interesting and new finding from this study is that pressure fluctuation where the wall - jet peak is observed are especially strong when vortex - wall interaction happens while two vortices are in the process of pairing. He showed that during such pairing, which happened at , the resulting secondary vortex is much stronger than wh en a single vortex interacts with the wall . In particular, he saw that, occasionally, vortices may pass without pairing, in which case, the pressure spikes were not as strong. The switch between pairing/no - pairing seemed to happen randomly in time . The nea r - wall pairing produced strong negative pressure spikes that reach ed a magnitude comparable with the dynamic pressure , based on the jet exit velocity. Al - Aweni also found that the positive - pressure source identified earlier by Naguib and Koochesfahani [13] , which is associated with the high strain rate within the separating boundary layer, has a significant influence on the shape and evolution of the strong negative spikes. 1.2.3 Governing Equations To get better understanding of the physical mechanisms le ading to wall - pressure fluctuation, it is insightful to connect the pressure fluctuation generation to the vortical structure s . in incompressible turbulent flow: (1.1) 10 Where, is the velocity gradient tensor and Einstein tensor notation is used . The forcing term in equati on (1.1) ( right - hand side) can be divided into two parts: symmetric (in terms of the strain rate ), and antisymmetric (in terms of the rotation tensor ); see Bradshaw and Koh [14]: (1.2) Where, (1.3) (1.4) It is known that the rotation tensor is connected to the vorticity vector through: (1.5) Thus, equation (1.2) can be written as (1.6) From equation (1.6), the pressure source strength ( q ) is given by: (1.7) Equation (1.6) may be written in vector form as follows, (1.8) 11 function can be used to solve equation (1.8) (e.g. Blake [15]) to get the pressure on a solid wall beneath an unsteady flow, where the wall - normal coordinate , for pressure source distribution within the flow . More specifically, (1.9) The volume integral (first term) in equation (1.9) rep resents the contribution to the wall pressure by the flow structure within the body of the flow , while the surface integral is computed over the wall beneath the flow . For a flat wall, the second term is negligible (from boundary layer approximation perspe ctive). The volume integrand shows that there are two wall - pressure generation mechanisms: one related to strain rate, and the other to vorticity. The former results in the generation of positive, and the latter in negative wall pressure. Thus, flow featur es where rotation dominates strain effects (e.g. in the core of vortices), generate negative pressure, while those associated with dominant strain (e.g. the zone in between interacting vortices) results in positive pressure generation; e.g. see Naguib and Koochesfahani [13]. Another important feature of the volume integrand in equation (1.9) is that the source term effect is inversely proportional to the distance between the point of wall - pressure observation and the pressure source location ( as seen from t he numerator of the integrand). Thus, as this distance increases , the observed wall pressure decreases. Therefore , the pressure observed at a point on the wall is a global quantity , related to all features within the flow with the net pressure being the re sult of the integrated effect of the strength and proximity of the different sources. 12 1.3 Motivation The motivation for this research is to understand the physics of, and to predict the unsteady surface pressure generation in impinging jets due to their signif icance in flow - induced noise and vibration. As described earlier in this chapter, the basic connection between the dominant flow features in impinging jets and wall - pressure generation is fairly well understood. However, there is practically no effort that capitalizes on this understanding to develop physics - based (also known as structure - based) models to compute the wall pressure in impinging jets. Such models, if sufficiently accurate, could be valuable as engineering design tools for flow - induced noise a nd vibration applications, since the models are much more efficient to run than direct numerical simulations, and they are more robust than non - physics based turbulence models. Additionally, physics - based models could be used to understand the underlying flow physics from a point of view that is not possible with experiments or numerical simulations. Specifically, in real flows, or their simulations, it is difficult to vary certain flow or structure parameters one at a time because of the interdependence o f these parameters. With a mathematical model, such variations are possible, which could lead to a clearer understanding of the effect of individual parameters on quantities of interest; the wall - pressure fluctuation in the present work. An example of such model - based insights may be found in the work of Monnier et al. [16], where a Gaussian - core vortex array was used to model the wake of a harmonically pitching airfoil. The model helped to understand the connection between the parameters describing the wak e - vortex configuration (vortex streamwise and cross - stream spacing, vortex circulation, and core radius) and the mean thrust acting on the airfoil. 13 1.4 Objectives The objectives of this study may be summarized as follows: 1 - Developing a simple mathematica l model, which represents the first step towards a high - fidelity model, for predicting wall - pressure fluctuations in normally impinging jet flows. The model will be used to explore the effect of vortex passing frequency, vortex circulation, and Reynold num ber on the unsteady wall - pressure characteristics, while varying one parameter at a time. In addition, the model results, which are only applicable to the stagnation zone due to inherent simplicity/ limitations of the model, will be compared to experimental data obtained in the present study. 2 - To conduct measurements of the unsteady surface pressure in a normally impinging jet using a microphone array. The measurements are done for two Reynolds numbers: 8272 and 24818 for H/D values of 2, 3 and 4 (reaching to the end of the potential core). The measurements will be used to examine the influence of the Reynolds number on the characteristics of the wall pressure flucution (primarily the radial di stribution of the root mean square pressure fluctuation, the probability density function of the pressure fluctuation and the power spectral density) . The lower Reynolds number is selected to match this of the earlier study by Al - Aweni [ 7 ] in the same jet facility. Al - Aweni utilized time - resolved flow visualization and microphone - array measurements to connect the flow features to the wall - pressure characteristics. Thus, combining Al - those from the present work, it is possible to infer the effect of Reynolds number o n the sources of wall - pressure generation. The remainder of the thesis is organized to show the mathematical model details in Chapter 2, the mathematical model results in Chapter 3, the experimental apparatus and procedure i n Chapter 4, 14 the experiment results and their comparison with the model results in Chapter 5, and conclusions and recommendations for future work in Chapter 6. 15 CHAPTER 2: Mathematical Model 2.1. Modeling Backgroun d Prior to describing the features of the physics - based model developed here, a recap of the flow details that should be captured by a good model is described . Near the jet exit, vortices are generated due to the instability of the shear layer and its subs equent roll - up into axisymmetric vortex rings. As the vortex rings travel downstream , the vortex size and strength increases by merging. Several such mergings may occur, with the number of successive pairing s depending on Reynolds number and distance betwe en the nozzle and the impingement plate. When the Reynolds number decreases, the coherence of the flow structure increases, and the number of successful pairings increases. Nearing impingement, within the stagnation zone , the flow changes its direction to be parallel to the plate. Further downstream as the flow advects through the wall - jet zone, boundary layer separation leads to secondary vortex formation and maximum wall - pressure fluctuation. Based on the literature review of Chapter 1, surface pressure f luctuation in impinging jets is caused by both inviscid phenomena (direct influence of vortices in stagnation and wall - jet zones) and viscous phenomena (vortex - wall and vortex - vortex - wall interaction in the wall - jet zone). The above description highlights that a complex model is needed to capture the impinging jet flow features and associated wall - pressure fluctuation. More specifically, elements of a high - fidelity model should consist of A viscous vortex ring model ; Ability to model vortex pairing ; Advec tion velocity field consistent with the actual mean jet velocity field ; 16 An impingement - plate boundary layer model ; A vortex - boundary layer interaction model ; - pressure from the velocity field . However, develop ing such high - fidelity model is an ambitious goal that requires several stages of development. This work focuses on the starting step by considering only the simplest possible model. Table (2.1) shows a comparison of the features of a high - fidelity and the present model. Feature High - Fidelity Model Current Model Vortex ring model Viscous - core vortex model Potential vortex model Vortex - vortex interaction Mutual induction Pairing model Not m odeled Advection field Based on impinging - jet mean flow; e.g., fro m CFD or experiment Potential stagnation - point flow Impingement - plate boundary layer Modeled Not m odeled Vortex - boundary layer interaction Modeled Not m odeled Wall - pressure calculation equation Table 2. 1 Comparison of high - fidelity and present model features Table (2.1) demonstrates that the present model is very simple and ignores all viscous and vortex - vortex interactions effects. However, evidence suggests that within the stagnation - zone wall - 17 pressure fluctuations are primarily generated via potential flow mechanisms for . Therefore, the present model may be successful in at least reproducing the same qualitative features of stagnation - zone wall - pressure fluctuations. One of the main goals of this work is to assess the ability of the present simple model to do so. 2.2 Model Details 2.2.1. Vortex Rings The full inviscid mathematical model for the impingement - jet problem is depicted schematically in figure (2.1). A cylindrical coordinate system i s used to describe the problem mathematically. However, becau se of symmetry, the model equations have no dependence on . Each of the jet vortices is modeled employing a potential vortex ring that has an axis of symmetry perpendicular to the impingement wall. Each ring has circulation , radius , zero cor e radius , and vertical core coordinate . In the absence of the wall, the ring is in free space and translate s downwards (for the sense of circulation depicted in figure 2.1 ) due to self - induction effects. The stream function for the ring is gi ven by Helmholtz equation [17] 18 Figure 2.1 Schematic drawing of the full mathematical model of the impinging jet (2.1) where (2.2) represe nts complete elliptic integral of the first kind, and represents complete elliptic integral of the second kind: 19 (2.3) (2.4) To make equation 2 .1 dimensionless , the initial vortex - ring radius is used as a length scale. S ince in jet flow, the jet vortices have a radius approximately equal to the jet opening radius, this length scale is equivalent to half the jet diameter, D /2 . The velocity scale is taken as the velocity of the advection field (potential stagnation flow, at the top of the computational domain , ), which is given by (see section 2.2.3) . This velocity scale correspond s to the mean exit jet veloci (2.1) is: (2.5) The above equations are for zero - core radius potential vortex ring. The infinitesimal core leads to an infinite induced velocity at the core , which makes it impossible to track the vortex movement. To overcome this problem, a finite, but small, vortex core radius is assumed . Following the work of Walker et al . [18], the velocity distribution inside the core is assumed to be uniform. This leads to the following self - induced velocity of the ring: (2.6) 2.2.2. Wall ( V ortex R ings I mages) The equations discussed in section 2.2.1 are valid for a vortex in free space . These equations need to be modified for the presence of the impingement wall; i.e., by enforcing a zero wall - normal velocity, or no - penetration, condition at the wall. This is done by adding to the model an image vortex, relative to the wall with equal but opposite circulation to the real 20 vortex (see figure 2.1). this velocity induced by the image on the real vortex ring has radial and wall - normal components . Thus, as the real vortex ring approaches the wall, the image vortex ring will affect the real ring in two ways: by stretching the ring radially outwards and reducing the approach velocity of the ring toward the wall. The radial stretching of the ring causes the ring radius to increase. According to Helmholtz, for inviscid incompressible flow, the vortex lines move with the fluid particles, and the vortex ring must have constant volume, i.e. (2.7) Where is constant. Thus, in the presence of the wall, increases and decreases with time. Both of these quantities affect the self - induced velocity (see equation 2.6). The resulting self - induced velocity componen ts , using equations (2.6 and 2.7) respectively are: (2.8) (2.9) The stream function for the image vortex is given by (2.10) where (2.11) 21 And as before, and represent the complete elliptic integrals of the first and the second kind respectively . The velocity induced by the image on the real vortex is obtained from the stream function due to the image by applying equations (2.12 and 2.13) and setting and ; i.e. , the coordinates of the real vortex core. (2.12) (2.13) This lead s to the velocity components produced by the image vortex ring on the real vortex (see Walker et al . [ 18 ] ) : (2.14) (2.15) 22 2.2.3. Advection Field (Stagnation Point Potential Flow) The next phase of the model involves adding steady flow to represent advection of the vortex ring s by the jet mean flow. To this end, potential point - stagnation flow is selected . The stream function of the flow is given by [19] as (2.16) Applying equations (2.12 and 2.13) to (2.16) leads to (see Naguib et al [ 20] ) : (2.17) (2.18) where is a constant with dimension . Equation 2.1 7 shows that the mean flow radial velocity increases linearly and unboundedly with . In reality, this kind of variation is expected to hold only near the stagnation point with deviation from the model increasing with the radial distance. Ther efore, we anticipate that this crude, yet very simple, model to be reasonable only within the stagnation zone. 2.2.4. Vortex A dvection and F low - F ield E volution in T ime For a single vortex convecting through the computational dom ain shown in figure 2.1, the velocity field at any time instance consists of three components due to: 1. t he real vortex ; 23 2. t he image vortex ; 3. t he stagnation flow . The stream function of the system is the sum of the three corresponding stream functions; specific ally (2.19) The velocity components of the resulting flow are found using equations (2.12 and 2.13) (2.20) 24 (2.21) where (2.22) 25 (2.23) The summation in (2.20) and (2.21) is over the number of vortices present in the domain at any given time instant. At the start of the computation, a vortex ring is pla ced at the top of the domain at . as time progresses, the ring core coordinates change with time according to (2.24) (2.25) To model the periodic formation of vortice s in a jet, as the leading vortex moves toward the impingement plate , subsequent vortices are added at the top of the domain at a selected frequency . All vortices approach the wall, gradually changing their dominant travel to be radially outwards. The comp utation is continued until the first vortex passes the end of the domain ( i.e., ) . 26 2.2.5. Wall - Pressure Calculation and Evolution in Time equation. (2.26) Where , is the velocity potential, is gr avity acceleration, and is a function of time. Evaluating the LHS on the wall and at a suitable reference point, and ignoring gravity effect , equation (2.26) becomes (2.27) Or (2.28) Equation (2.28) shows that to obtain the wall pressure at a given point at any instant in time requires: 1 - Knowledge of the radial v elocity component at the wall . This information is readily available from the velocity field equations (2.20) and (2.21) ; 2 - Selection of a reference point where: the velocity magnitude is known and the pressure is fixed so it can be used as a reference pressure. The first of these conditions is easily satisfied using equations (2.20) and (2.21). The second condition is more difficult to satisfy and , as will be seen below , meeting this condition sets a lower limit on the radi al computational domain size and an upper limit on the computation al time. 27 3 - Calculation of the time rate of change of the velocity potential difference between the point of interest and the reference point. At every time instant, the velocity potential can be computed by inte grating the velocity field. This leads to two potential issues. First, the integration yields the potential with an unknown additive integration function of time. This issue is not problematic since the time function is subtracted out when computing the po tential difference . Second, though the velocity field is known analytically, its form is not easy to integrate. Therefore, integration is done numerically to get the potential difference at each time instant. The resulting fields are subsequently different iated numerically in time to arrive at the first term on the RHS of equation (2.28). Further details follow. As depicted in figure (2.1), the reference point is selected on the wall at the end of the computational domain . This point rem ains unaffected by the vortex rings until the first vortex ring convects through the entire domain, reaching near . Thus, the pressure is steady up to the point of arrival of the first vortex ring, at which point the computation is st opped. This imposes a limit on the computational time, leading to a tradeoff process. Given the periodic influence of the vortices, the computation must be run until at least two vortices pass by a given point on the wall in order to obtain a full period o f pressure fluctuations. Thus, must be large enough for the computation time to be larger than the lowest period of vortex passage of interest. On the other hand, cannot be made arbitrarily large since the radial domain must be discretized finely for the integration leadi ng to the calculation of the velocity potential difference, which could lead to a prohibitively long integration time. Therefore, is set only as large as necessary by allowing enough running time for the core of at least the second vortex injected into the domain to reach the largest radial location of interest. As discussed previously, it is expected that the applicability of the present model is limited to the stagnation zone. Accordingly, the extent of the 28 radial domain of interest (see f igure 2.1) is , where is the equivalent of the jet diameter. The unsteady potential function difference , is calculated using: (2.29) which leads to , (2.30) Where, is a dummy variable for integration along the radial coordinate . The integral (2.30) is evaluated numerically, where the integrand is obtained by setting in equation (2.2 0 ). Notably, since it is the time derivative of (2.30) that is required for evaluating the wall pressure, the stagnation flow component, which is steady, does not affect the unsteady potential difference term. 2.3. Numerical Details The model was implemented numerically using Matlab. The implementation consisted of two main tasks. The first one involved periodically seeding vortices at the entrance of the domain and tracking the core centers of these vortices as they advect through the computational domain . K nowledge of the core centers locations at each time instant enabled computation of the instantaneous stream function and velocity field. This information is used to visualize the flow field concurrently with the wall pressure. The second task employed the distribution of the radial velocity component at the wall to compute the distribution of the wall - pressure at each time instant. The specifics of the numerical implementation of these two tasks are given in the two following sub - sections. 29 2.3.1. Computati on of the Time - Dependent Stream Function and Velocity Field Given the initial core coordinates ( R o , Z o ) of a vortex at the top of the computational domain, subsequent locations of the core ( R, Z ) at each time step was determined by solving equations (2.24 and 2.25) using fourth - order Runge - Kutta method for the two variables R ( t ) and Z ( t ) . Once ( R, Z ) was determined , the stream function was calculated by evaluating equation (2.19) on 100 × 100 - point grid using symbolic math tools in Matlab. The solution was independent of the time step and grid resolution, as demonstrated in Section 2.4. 2.3.2. Calculation of t he Wall Pressure at any radial location relative to the pr essure at the reference point requires knowledge of the radial velocity at the same radial location and at the reference point, as well as the rate of change of the velocity potential difference between the pressure observation and the reference point. The velocity information was straightforward to obtain. Specifically, once the vortex core locations were determined at every time instant in the first task, the radial velocity distribution on the wall could be determined by setting z = 0 in the analytical e quation (2.20). The reference velocity was further determined by setting r = r ref in the resulting equation. The radial velocity distribution, and subsequently the wall pressure, was evaluated at 600 points on the wall spanning from the axis of symmetry of the domain to the reference point. This number of points was sufficiently large for the calculated wall pressure to be independent of the number of grid points (see Section 2.4 for details). On the other hand, to find the unsteady term on the wa ll equation (2.28), was determined by numerically computing the integral (2.30). 30 Specifically, the integral was discretized on the 600 - point wall grid using the method of rectangles, leading to: (2.31) Where i is an index of the wall grid points starting from the point next to the reference point and increasing towards the axis of symmetry, n is th e number of grid points (600), and is the radial resolution of the grid. The negative sign on the right - hand side reflects the fact that the integration is in the direction of decreasing r coordinate. To evaluate the summation in (2.31) , the radial vel ocity u ri was evaluated by substituting for z = 0 and r = r i in equation (2.20). This enabled evaluation of using (2.31) for all time instants. Subsequently, the time derivative of the potential difference was computed using forward finite differencing. The solution was independent of the time step and grid resolution, as demonstrated in Section 2.4. 2.4 Validation of the Computational Approach Validation tests were done to ensure that the model results are independent of: - Radial c omputational domain size; - Wall - pressure resolution (number of points on the wall to calculate the pressure); - Time step. In addition , the outcomes of these validation tests, in terms of the domain size, radial wall - grid resolution and time step, were employ ed to numerically compute the unsteady surface pressure associated with a line vortex advecting parallel to a flat wall. This was done since: (1) the advecting line vortex problem has certain similarities with the current vortex - rings problem; and 31 (2) the wall - pressure is known analytically for the line vortex problem, enabling further verification of the numerical solution approach and implementation. 2.4.1 Impinging - Jet Model 2.4.1.1 Validation of the Computational Domain Size As described in Section 2.2, th e main influence of the radial domain size is related to the basic assumption that the pressure at the reference point, which is placed at the radial end of the domain, remains steady throughout the computation. Thus, a larger domain allows longer computat ion time since it takes the first vortex injected into the domain longer to reach the end and affect the pressure at the reference point. To ensure that the selected domain size is appropriate, three different computational domains having a radial extent o f 10, 15, and 20 cm are compared . These values correspond to approximately 20, 30 and 40 times the initial vortex radius. For each of these domains, the wall pressure is computed and the resulting radial distribution at a time is located at r = 5 cm is considered (Figure 2. 2 ). The selected location is 10 times the initial vortex radius; well beyond what would be considered the stagnation zone of the jet ( r = one jet diameter twice the initial vortex ring radius), which is the main focus of the present work. The three pressure distribution s from the different domains are compared in pairs at the selected time instant using a Normalized Root Mean Square error ( NRMSE ), defined as follows: (2.32) whe re subscripts domain 1 and domain 2 denote the two domains under comparison, is the magnitude of the negative pressure spike (minimum pressure) produced by the vortex, and the summation is over all wall grid points; n = 600. Other than the do main radial extent, other 32 model parameters are the same as used for the results presented in Chapter 3. It is noteworthy that the computational time for the three different cases varie d from about 10 to 45 minutes for the smallest and the largest domains r espectively for the reference case which is define in table 3.1 . Figure 2. 2 Radial wall - pressure distribution at the time instant when the negative pressure spike (minimum pressure) associated with vortex passing is located at r=5 cm location. Diffe rent color lines represent different radial domain size s used for computing the wall pressure (as given by the legend). The inset shows the overall distribution, while the main plot depicts a magnified view to show the change between the different cases. 33 Figure 2. 3 depicts the NRMSE % as a function of time up to the time corresponding to Figure 2. 2 . The maximum pressure magnitude at each time step is used in calculating NRMSE . As seen from the figure, the effect of increasing the domain size beyond 10 cm is practically zero (well below 0. 1% ) . Therefore, a 10 - cm wide domain was utilized for all the computations conducted here. Figure 2. 3 Normalized Root Mean Square Error ( NRMSE ) resulting from comparing the results of domain sizes 10 and 15 cm (blue) and 15 and 20 cm (red). The results are shown up to the computation time corresponding to the time instant where t he negative pressure spike appears at r = 5 cm . 34 2.4.1.2 Validation of the Wall - Pressure - Grid Resolution This step targets the validation of the resolution of the wall grid used to find the wall - pressure fluctuations. The higher the resolution, the better t he accuracy. However, increasing the resolution is expensive, consuming more computer resources and time. To validate the selected wall resolution , three wall resolutions were examined for a radial domain extent of 10 cm : 0.33 mm (300 points on the wall), 0 .16 mm (600 points on the wall), and 0.0833 mm (1200 points on the wall). These resolutions correspond to approximately 0.07, 0.034 and 0.017 of the initial vortex ring radius . All other model parameters are as described in Chapter 3. It is noteworthy that the computational time for the three different cases varie d from about 10 to 40 minutes for the resolution 0. 33 mm and 0. 0833mm respectively for the reference case which is define in table 3.1. The pressure distributions for the three cases considered are shown in figure 2. 4 for the time instant when the pressure spike is located at ( t = 0.0039 s ) . The difference between the different cases is quantified using NRM SE , which is shown in figure 2. 5 up to the time when the negative pressure spike appears at . Noticable from the plot, the maximum error decreased from approximately 10% for the case between 300 and 600 points to less than 1 % for the case between 600 and 1200 points . Therefore, 600 grid points on the wall were used for all wall - pressure results reported in Chapter 3. 35 Figure 2. 4 Radial wall - pressure distribution at the time instant when the negative pressure spike (minimum pressure) associated with vortex passing is located at . Different color lines represent a different number of wall grid points used for computing the wall pressure (as given by the legend). The inset shows the overall distribution, while the main plot depicts a magnified view to show the change between the different cases. 36 Figure 2. 5 Normalized Root Mean Square Error ( NRMSE ) resulting from comparing the results of grid resolution 300 and 600 points (blue) and 600 and 1200 points (red). The results are shown up to the computation time corresponding to the time instant whe n the negative pressure spike appears at . 37 2.4.1.3 Validation of the Time Step Size The final validation step is to verify the size of the time step to produce results with an acceptable accuracy while minimizing time and computer resources. For this evaluation, three - time s teps are used corresponding to 1000, 2000 and 4000 - time steps during the period between injecting two successive vortices in the domain. The corresponding time step is 4.06, 2.03, and 1.016 s , respectively, for the vortex passage frequency considered (det ails of how the physical value of frequency is determined will be given in Chapter 3) . All other model parameters are as described in Chapter 3. The computational time for the three different cases varie d from approximately 10 minutes for 1000 points/cycle to 40 minutes f or 4000 points/cycle for . Pressure distributions for the three - time steps utilized are shown in figure 2. 6 at the instant when the pressure spike is located at ( t = 0.0039 s ) . As before, t he difference between the different cases is quantified using NRMSE in figu re 2. 7 . As seen from the plot, the maximum error decreased from 3% when comparing between 1000 and 2000 point per cycle ( PPC ) to 1.65% between 2000 and 4000 point per cycle ( PPC ) . All computations done here utilized 2000 PPC . 38 Figure 2. 6 Radial wall - pres sure distribution at the instant when the negative pressure spike (minimum pressure) associated with vortex passing is located at . Different color lines represent a different number of computational time steps in the period between injecting two successive vortices (as given by the legend). The inset shows the overall distribution, while the main plot depicts a magnified view to show the change between the different cases. 39 Figure 2. 7 Normalized Root Mean Square Error ( NRMSE ) resulting from comparing the results of time step 4.066e - 06 s (1000 PPC ) and 2.033e - 06 s (2000 PPC ) (blue) and 2.033e - 06 s (2000 PPC ) and 1.0161e - 06 s (4000 PPC ) (red). The results are shown up to the computation time corresponding to the time instant whe n the negative pressure spike appears at . 40 2.4.2 Application to a Line Vortex Above a Flat Wall To validate the overall unsteady - wall - pressure computational approach and the implemented algorithms , it was desired to apply the method/algorithms to a closely related problem , where an analytical solution for the unsteady wall pressure is known. The problem of a line vortex above a wall seemed to be appropriate. Similar to the vortex - ring problem , the presence of the line vortex above a flat wall , as depicted schematically in figure 2. 8 , is modeled using an imaginary image vortex placed symmetrically relative to the wall. The presence of this image vortex causes advection of the vortex pair in the positive streamwise direction (for the given sense of circulation), producing an unsteady wall - pressure imprint on the wall. Physically, this scenario is similar to that of the vortex ring once it changes its advection direction from being predominantly towards, to being approximately parallel to the wall. In both the line and ring vo rtex cases, the unsteady pressure field at a given point on the wall is caused by the passage of vortices above the point . Therefore, the ability to reproduce the unsteady wall - pressure field of the line vortex accurately using the tools developed for the impinging jet problem provides further confidence in the implementation of the model developed in the present work. 41 Figure 2. 8 Schematic of a line vortex centered at ( x o , y o ) above an infinite flat wall at y = 0. An image vortex centered at ( x o , - y o ) m odels the presence of the wall. The velocity, and velocity potential for the line vortex problem are given by [21]: (2.31) (2.32) The corresponding wall - pressure distribution at a given time instant is computed using eq uations (2.31) and (2.32) in conjunction with equation (2.28) . The numerical implementation is the same as described in Section 2.3. For this calculation, the vortex and computational parameters are kept the same as for the vortex ring case (without the in clusion of the stagnation flow). Specifically, the parameter s values a re : 42 - Domain size: - Time step: - Circulation: - Wall resolution: - Vortex location above the plate Figure 2. 9 depicts the wall - pressure signature obtained an alytically and computationally, concurrently with the streamlines at selected time steps. The streamlines are depicted in a frame of reference convecting with the vortex but placed at the appropriate x location as the vortex travels in the positive x - direc tion pressure peak is found ; as expected. Two additional positive, but substantially weaker, peaks are found ahead and behind the vortex center. Overall, excellent agreement is found betwe en the computed and analytical wall pressure distribution even up to the time when the vortex is fairly close to the reference point (subplot f in figure 2. 9 ). This agreement is quantified by calculating the NRMSE , which is displayed in figure 2. 10 versus time. For this plot, the time show n is only up to the point when the vortex center reaches x = 5 cm . As seen from the figure, the NRMSE does not exceed 0.3% over the entire duration depicted. 43 (a) Time=0.00278 s (b) 0.00627 s (c) 0.01324 s (d) 0.02022 s (e) 0.0237 s (f) 0.02719 s Figure 2. 9 Computational and analytical results of the wall pressure and associated streamlines for a line vortex above a wall. Different subplots represent different times (as indicated beneath the plots). The streamlines are shown in a frame of reference convecting with the vortex. 44 Figure 2. 10 Normalized Root Mean Square Error (NRMSE) between a nalytical and numerical values of the wall - pressure distribution beneath a line vortex above a flat wall. The time duration shown is from the start of the vortex motion until the vortex center is located at r = 5 cm. 45 CHAPTER 3: Mathematical Model Results This chapter will focus on the interpretation of the mathematical model results including the mean and Root Mean Squa re ( RMS ) of the fluctuating pressure. The analysis was done for three main cases. For each case, the solution was obtained when changing only one parameter while the other parameters remained unchanged. This facilitates understanding the effect of changing each of these parameters on the wall - pressure characteristics. In fact, varying one parameter at a time is one advantage of the model because in the actual flow it is generally not possible to change these parameters independent of one another. The three parameters examined are: vortex passage frequency ( f ), jet Reynolds number , and vortex circulation . Though (vortex ring diameter) and initial jet velocity (stagnation flow velocity at top of the computation domain) in ord er to facilitate comparison with actual jet flow results. The nominal values of the model parameters were selected to be representative of an actual jet flow at the end of the potential core. The details of determining these values follow. Crow and Champag en [14] showed that the jet unsteadiness at the end of the potential core is dominated by the preferred, or column, mode with a Strouhal number of . Thus, (3.1) In the model, the jet diameter is reasonably equivalent to twice the vortex ring radius; i.e. , and the jet velocity to the centerline velocity of the stagnation (advection) flow at the top o f the computational domain; i.e. . Making these substitutions in equation (3.1): 46 (3.2) In equation (3.2), both and are known a priori. is the height of the computational domain, and is the initial vortex ring radius (equivalent to the jet exit radius), which is arbitrarily selected without loss of generality since it is also chosen as the length scale for making quantities non - dimensional. To determine , and hence the vortex pas sing frequency via equation (3.2), use is made of : (3.3) where, is the kinematic viscosity. Equation (3.3) leads to: (3.4) Hence, by specifying the ri ng radius, jet Reynolds number and knowing the computational domain height it is possible to compute the nominal vortex passage frequency (corresponding to the jet column, or preferred mode). Because the jet preferred mode frequency corresponds to the vort ex passage frequency at the end of the potential core, the frequency calculated as outlined above represents vortices near the end of the potential core (approximately downstream of the jet exit). To model vortex passage at locations closer to the exit of the jet, use is made of the fact that the terminal vortex passage frequency of the preferred mode is the result of successive pairing s of vortices, which initially form as an instability of the jet shear layer, rather than the jet column. With each pairing, the vortex passage frequency drops by a factor of 2. Hence, to represent vortex passage at distances closer to the jet, the nominal frequency is increased by a fa ctor of 2, 4, etc. To connect the nominal vortex circulation value to the jet flow, use is made of the approximation of the circulation for a shear layer (Koochesfahani and Dimotakis [22]): 47 (3.5) where, is the velocity difference across the shear layer, is the wavelength between vorti ces, and is the circulation around a contour that encompasses the shear layer and extends a length in the streamwise direction. Equation (3.5) assumes all vorticity is concentrated in the shear layer vortices and is equivalent to computing the velocity jump across a zero - thickness potential vortex - sheet from the circulation density (circulation per unit length) along the sheet. For the jet flow and [23], thus: (3.6) For all cases studied, parameters - maintained constant are - radial extent: (see figure 2.1 for definition). is selected to be much larger than and to ensure invariance of the results with further in crease in , as discussed in Chapter 2; - Initial vortex ring radius: ; - Initial wall - normal distance between the vortex ring and the impingement plate (also wall - normal extent of the computational domain ): m . This value was not critical . If is increased , the vortices will travel a longer distance towards the wall before they have any effect on the wall. Since the focus of the model is on the wall - pressure, needs to be larger than or equal to the height at which the vortex presence affects the wall pressure. The choice of five times the ring radius fulfills this requirement for all cases examined here; - Volumetric constant ( ) in equation (2.7): . This value was taken to be the same as used by Walker et al [18] . The specific value of C is not critical since it only 48 affects the self - induced velocity of the vortex, which in turn only affects how quickly the vortex ring lineally approaches the wall. - Air kinematic viscosity ( at ): . Air is selected as the fluid since the results of the model are compared to the measurements in an air jet in Chapter 5. With and values fixed, the selection of , sets (equation 3.4), which in turn se ts and (equations 3.2 and 3.6, respectively). A reference case was arbitrarily selected to correspond to , resulting in and . This reference case was repeated in three series of parametric investigations ai med at examining the influence of , and on the wall pressure. Each series contained three cases, including the reference and involved varying only one parameter, while all other parameters remained fixed in order to isolate the influence of each parameter. This one parameter at a time variation is only possible using a mathematical model since in the real jet , and are generally interdependent. Indeed, this independence is even reflected in equations (3.2), (3.4) and (3.6), whi ch attempt to mimic the real jet flow in a simplistic way. The parameter values for all cases investigated are summarized in table 3.1. For the frequency series, two other frequencies representing doubling and quadrupling the reference - case frequency ar e examined. As discussed earlier these higher frequencies represent vortex passage prior to the second and first vortex pairing, respectively, in the real jet (which corresponds to shorter jet to impingement plate distance). For the Reynolds number cases, the three values utilized are . Finally, the three cases in the circulation series are those from the reference case , and the others taken to be half and double of the reference circulation . For each series, the reference case is highlighted in table 3.1 using green color. 49 case Reynolds number Frequency Jet velocity Circulation Constant Frequency 5000 245.9 7.8698 0.1798 163.954 5000 491.8 7.8698 0.1798 163.954 5000 983.6 7.8698 0.1798 163.954 Reynolds number 5000 245.9 7.8698 0.1798 163.954 15000 245.9 23.6094 0.1798 491.862 25000 245.9 39.3490 0.1798 819.77 Circulation 5000 245.9 7.8698 0.0899 163.954 5000 245.9 7.8698 0.1798 163.954 5000 245.9 7.8698 0.3596 163.954 Table 3.1 Model parameters for frequency , Reynold number , and circulation cases. Rows highlighted in green depict the parameters for the reference case 3.1 Frequency Effect Figure 3 .1 and 3.2 show snapshots of the streamlines of the computed flow and associated wall - pressure for the lowest and highest frequency respectively. In each figure, the point - vortex location is indicated with an asterisk. In both figures, a given vortex appro aches the wall moving vertically downwards without much change in the radial location, until the vortex is very close to the wall where the main travel direction switches from towards to parallel to the wall. The main difference between the two frequency c ases is that multiple vortices are seen within the computational domain in the high - frequency case (figure 3.2). The spacing between the vortices initially decreases, as they approach the wall, resulting in packing them rather densely within the stagnation zone. Subsequently, as the vortices travel parallel to the wall, their spacing increases 50 substantially. The change in the vortex spacing is related to their convection velocity. Initially, as they move towards the wall, their convection velocity decreases (due to the decrease in the stagnation flow wall - normal velocity and the opposing influence of the image vortex). Once they by the image vortex both com within the wall - jet zone the spacing of the vortices is increased substantially. In the wall - jet zone, b ecause the vortices are spaced far apart, regardless of their frequency, their wa ll - pressure signature is similar in character in both the low - and high - frequency cases. This signature takes the form of a negative pressure spike immediately beneath the vortex (where the induced velocity by the vortex is highest on the wall) surrounded by two small positive peaks up and downstream of the negative peak. This signature , which is consistent with the expected focusing of the stream lines beneath a vortex (see figure 2.9), can be observed best at the last time instant (largest vortex location) in figures 3.1 and 3.2. The negative spike first appears when the vortex gets close enough to the wall (as will be seen later, this takes place within the stagnation zone, between ), then it monotonically increase s in magnitude as the vortex travels radially outward s . This monotonic increase is caused by the continuous decrease in the vortex distance to the wall with increasing time. Since potential vortices are employed in the present model, as the wall nears the vortex center, the induced velocity on the wall becomes higher and higher, and therefore the corresponding pressure is expected to become lower and lower. 51 t=0 s t=0. 0004 s t=0. 0008 s t=0. 0011 s t=0. 0015 s t=0. 0019 s t=0. 0022 s t=0. 0 026 s t=0. 003 s Figure 3.1 Streamlines and normalized pressure at different time step s for frequency , Re =5000, and circulation 52 t=0. 0 s t=0. 0004 s t=0. 0008 s t=0. 0012 s t=0. 0015 s t=0. 0021 s t=0. 0023 s t=0. 0026 s t=0. 003 s Figure 3.2 Stremlines and normalized pressure at different time step for frequency f = 980 Hz , Re =5000, and circulation 53 To examine the wall - pressure characteristics quantitatively, radial profiles of the m ean and RMS pressure are computed. Both of these quantities are calculated over one period of oscillation ( ) after the initial transients of starting the calculation have passed. The mean pressure distribution is shown in figure 3.3 in the form of a me an - pressure coefficient ; where is the dynamic pressure based on , for a truncated radial domain that focuses on the stagnation zone. The latter is delineated in these figures, as well as similar ones later in the chapter, using broken green lines. As seen in figure 3.3, the mean pressure has its maximum at the stagnation point (as expected). It should be clarified that the stagnation point pressure coefficient exceeds unity because of t hree reasons. First, the dyn amic pressure used for normalization is computed using a velocity scale that is representative of the stagnation (advection) flow only. Therefore, any added mean streaming velocity towards the wall due to the presence of the vortices does not affect the ve locity scale. The added velocity would increase the stagnation pressure beyond that of the stagnation flow alone, causing the pressure coefficient to go beyond unity. The difference between the overall pressure and that due to the stagnation flow alone can be seen in figure 3.3, where of the stagnation flow alone is also plotted (indicated with SP in the legend). As expected, this distribution does not vary with frequency. Second, the reference pressure used in defining is taken from a different spatial location ( than that where , and hence , is calculated ( . Third , the flow is unsteady, and therefore, generally speaking, the stagnation pressure coefficient need not be unity. 54 Figure 3.3 Effect of vortex passage fr equency on the radial distribution of the mean pressure coefficient . The green broken lines outline the boundary of the stagnation zone Figure 3.3 shows that at the stagnation point increases with increasing frequency. This effect can be understo od as being caused by the net streaming velocity induced by the vortices towards the wall. As the frequency increases, more vortices are present simultaneously within the stagnation zone (e.g. compare figures 3.1 and 3.2), and the superposition of their in duced velocity increases with the number of vortices, leading to larger stagnation pressure. Another interesting effect is the development of a local negative peak just outside the stagnation zone, which is seen clearly for the highest frequency. This effe ct is also understood to be caused by the net streaming velocity induced on the wall by the collective effect of the vortices, which becomes stronger with increasing frequency. Unlike the streaming effect towards the wall at , which understandably p roduces a positive pressure peak on the wall at the stagnation point, when the flow velocity is forced to zero, the stronger the wall - parallel velocity induced by the vortices, the 55 lower the pressure. The presence of a localized peak at a given radial loca tion, for large enough frequency, is indicative of the location where the collective influence of the vortices is highest. As discussed in figure 3.2, as the vortices approach the wall, they get packed densely resulting in maximum collective effect. Howeve r, subsequently the vortex spacing increases substantially reducing this effect. The negative peak in the mean pressure distribution is believed to be a manifestation of this behavior when enough vortices are present in the simulation (i.e. when the vortex passing frequency is high enough). T he negative peakon figure 3.3 can be explained alternatively based on figures 3.1. and 3.2. More tightly packed vortices means a larger "induced" flow. Since this flow cannot bypass the streamlines, the velocity is incr eased in regions where the spacing between streamlines is reduced. That is the place where the vortices first approach the wall. Since vortex injection frequency is higher, streamlines tend to stay closer to each other almost permanently at that point. The radial distribution of is depicted in figure 3.4. Two different trends ), the level of pressure fluctuation decreases with increasing frequency, wh ile the opposite takes place for larger near the stagnation point. As increases beyond this flat zone, two different behaviors are seen, depending on frequ ency. At low frequency, a decrease in the RMS level is observed with the increase in , before a monotonic increase is observed. The initial decrease with becomes smaller with increasing frequency, eventually disappearing at the highest frequency , where the RMS level increases monotonically with beyond the initial flat distribution near the stagnation point. 56 Figure 3.4 Effect of vortex passage frequency on the radial distribution of the RMS pressure coefficient . The green broken lines o utline the boundary of the stagnation zone Possible physical reasoning for the trends observed in the distributions could be developed via inspection of sample pressure times series. Figure 3.5 depicts pressure signals for one cycle of vort ex passage at various r/D locations within the stagnation zone. Two plots are included in the figure for the lowest and highest frequencies; (top) and (bottom), respectively. For both cases, the general features of the signals are quali tatively similar, depicting a fundamental difference in the signal shape between radial locations that are below approximately and those that at larger radial locations. For locations larger than 0.5, an energetic pressure signature that i s characterized by a strong negative peak is observed. As discussed in figures 3.1 and 3.2, the strong/narrow negative pressure peak is found directly beneath the vortex closest to the wall, and is produced by vortex passage. Once seen at the wall, this pe ak continuously increases in magnitude with 57 decreases as the vortex convects radially outwards. This increase in negative - peak strength is seen in figure 3.5 between , for both frequencies. Below , the pressure s it does not exhibit the strong negative peak characteristic of vortex passage. Since the only unsteadiness in the model originate from vortex passage, there is no doubt that the unsteady pressure is still connected to the vortices; it is just that the unsteadiness does not reflect the strong local effect, near the vortex core. Therefore, hereafter the pressure unsteadiness is characterized as remote for , and local for . The remote effects are more harmonic and weak in nature, and the local effect are strong and only felt at the wall for sufficiently large . For real (viscous core) vortices, the distinction between local and remot e is expected to depend on how far is the point of observation from the vortex - core center relative to the vortex core radius. Since the vortex core size increases with each merging, one would expect that local effects would extend farther and farther with merging. This physical hypothesis cannot be tested with the present model, which is based on zero - core - size potential point vortices. 58 Figure 3.5 Pressure time series at various r/D locations within the stagnation zone, Re =5000, , and . The small discontinuities in the time series for correspond to the time instant when a new vortex is injected at the top of the computational domain. Because the discont inuities are small, no attempt was made to get rid of them by using a larger wall - normal extent for the computational domain. 59 The discussion of figure 3.5 provides insight regarding the effect of on in the stagnation zone. To analyze t he influence of frequency, pressure time series at the same radial location but different frequencies are shown in figure 3.6. Two radial locations are chosen, at the start and end of the stagnation zone : (top) and (bottom), respectivel y. These two locations are chosen because one of them exhibits decrease in pressure fluctuation level with f requency ( ), and the other, the opposite. Focusing first at the stagnation point, the decrease in the amplitude of the sinusoidal - like pressu re variation with increasing frequency is obvious. This trend may be clarified as follows: per earlier discussion of figures 3.1 and 3.2 , as the frequency increases, the vortices become packed densely near the wall within the stagnation zone . As a result, the induced velocity on the stagnation stream line (i.e. the flow approaching the stagnation point) by the individual vortices become increasingly overlapping. This reduces the vortex - to - vortex velocity fluctuation, and leads to more of a steady streaming flow with increasing frequency. To substantiate this physical picture just described , figure 3.7 shows the induced velocity along a line that is parallel to an array of line vortices with varying inter - vortex spacing: 10 mm, 50 mm , and 500 mm . This simple situation is intended to emulate wall are parallel to the stagnation streamline in the present model ). The results in fig ure 3.7 are plotted for only one wavelength (i.e. for a length equal to the spacing between the vortices), since outside the shown range, the induced velocity would repeat periodically for an infinite array. For all vortices in the array, the circulation i s the same. As seen from figure 3.7, for the largest wavelength (smallest packing of vortices), the vortices are spaced so far away that th e induced velocity from the neighboring vortices has no influence within the wavelength shown (as seen 60 from the veloc ity practically decaying to zero at the ends of the wavelength). As the vortices become more packed ( ), the effect of neighboring vortices overlaps more, causing an overall rise in the mean velocity (i.e. increasing streaming effect), and a reductio n in the velocity fluctuation (since the induced velocity amplitude does not fall too much before the influence of a neighboring vortex is felt). This trend is particularly evident for the highest vortex packing ( ), where the induced velocity is p ractically steady, having the strongest streaming component and no fluctuation. On the other hand, within the zone where the local vortex effects are felt (bottom plot in figure 3.6), the pressure amplitude increases only slightly with increasing frequenc y, suggesting that this increase is not responsible for the strong increase in RMS with frequency seen in figure 3.4. Instead, a substantial broadening of the pressure peaks is found to take place with increasing frequency. This shows that the monotonic in crease in with frequency for is a - broadening is again a result of the increase in the number of packed vortices affecting the press ure as the frequency increases. However, unlike , the effect of vortex packing relates to the induced velocity along a line that is practically normal, rather than parallel, to the vortex array, which leads to a different influence of vortex packi ng. To further clarify the effect of vortex spacing on the induced velocity on the wall, a simple situation is considered where three line vortices are placed at the same positions relative to a wall as the vortices in the model when the leading vortex cen ter is at (i.e. the same location for which the time series are shown in figure 3.6). The induced velocity on the wall by each of these vortices is plotted in figure 3.8 versus distance along the wall, together with the velocity induced by all of them. If th e vortex spacing is so large such that only one vortex passes through 61 the domain at a time (as in the low - frequency case of the present model) , the induced velocity signature at the time when the vortex is located at , would be the same as given by the red line in figure 3.8. On the other hand, if the vortex spacing is so small such that when the leading vortex passes , two other vortices are trailing in close proximity (with relative spacing similar to the high - frequency case of the model), substantial broadening of the induced velocity is seen due to the effect of the trailing vortices. This should influence the wall pressure in the same way, clarifying the influence of frequency on the time series in figure 3.6. 62 Figure 3.6 Pre ssure time series at various frequencies , Re =5000, , and . The small discontinuities in the time series for the two larger frequencies correspond to the time instant when a new vortex is injected at the top of the computational d omain. Because the discontinuities are small, no attempt was made to get rid of them by using a larger wall - normal extent for the computational domain. 63 Figure 3.7 Induced v elocity by an infinite line - vortex array along a line parallel to t he array. Results are shown over one vortex spacing ( ) with the vortex located at the center ( ). Different colors indicate different spacing (i.e. different vortex packing density). 64 Figure 3.8 Individual and collective induced wall velocity by three - line vortices placed relative to a wall at the sa - frequency case when the leading vortex is located at . 3.2 Reynolds Number Effect Figure 3. 9 and 3. 10 show the radial distribution for the mean and RMS wall pressure when the jet Reynolds number i s varied . Since the velocity scale is defined as that based on the stagnation (advection) flow at the top of the computational domain, increasing corresponds to increasing the strength/velocity of the stagnation flow. In addition, because the mod el is inviscid, essentially jet relative to that induced by the vortices since all vortex parameters remain unvaried. Figure 3.9 shows that the stagnatio n - point ( SP ) flow does not vary with Reynolds number. This is expected given the normalization by a velocity scale based on the stagnation flow. In contrast, the overall exhibits a deviation from of SP flow that decreases with increasing 65 . Th is should not be too surprising given that the deviation is produced by the streaming flow induced by the vortices, which should remain invariant since the vortex characteristics are unchanged between Reynolds numbers. This invariance when normalized with an increasing dynamic pressure causes the deviation from SP mean pressure to decrease with increasing . Similarly, the decrease of the RMS pressure coefficient with Reynolds number (figure 3.10) can be attributed to an invariance in the unsteady pressure from vortex passage that produces smaller coefficient of pressure when normalized by the incre Figure 3.9 Effect of Reynolds number on the radial distribution of the mean pressure coefficient 66 Figure 3.10 Effect of Reynolds number on the radial distribution of the RMS pressure coefficient 3.3 Circulation Effect Figures 3. 1 1 and 3. 1 2 depict the mean and RMS pressure distributions for various vortex circulation magnitudes (i.e. vortex strengths). The case with the intermediate circulation value is the reference case (the same as the in section 3.1), where one vortex at a time is affecting the wall pressure. The other two cases are similar with the exception of the vortex strength, which is weaker in one case and stronger in the other. The overall features of the mean pressur e profile stay the same with varying circulation strength. Quantitatively, the stagnation pressure increases, and a local minimum develops with increasing circulation. The same trends were observed to take place with increasing frequency in section 3.1. In that case, these changes were attributed to the increase in induced velocity due to the increasing number of vortices with increasing frequency. Here, the effect is similar but since 67 the number of vortices does not change (the frequency is the same for al l three cases), the induced velocity increases due to circulation. Figure 3. 1 1 Effect of vortex circulation on the radial distribution of the mean pressure coefficient 68 Figure 3. 1 2 Effect of vortex circulation on the radial distribution of the RMS p ressure coefficient In regard to the RMS distribution (figure 3. 1 2 ), the overall shape remains invariant with changing circulation (consult figure 3.4 to see the details of the reference case/intermediate circulation more clearly) but the RMS level incre ases monotonically with increasing circulation (consistent with the stronger vortices). Overall, the pressure signal shapes within the stagnation zone remain similar with increasing circulation, while the magnitude of the pressure peaks increases (leading to the larger RMS level). This may be seen by comparing figure 3. 1 3 for the highest - circulation case to the top plot in figure 3.5 for the reference case. A more direct comparison between the time series shapes of the lowest and highest circulation cases a t the stagnation point and end of the stagnation zone is given in figure 3. 1 4 . The results clearly demonstrate the increasing strength of the pressure fluctuations without change in the pressure signal shapes. 69 Figure 3. 1 3 Pressure time series at variou s r/D locations within the stagnation zone, Re =5000, , and 70 Figure 3. 1 4 Pressure time series for various vortex circulation s , Re =5000, , and . 71 CHAPTER 4: Experiment Setup and Apparatus Thi s chapter exhibit s the experimental setup that was used to measure the unsteady wall pressure and the initial jet velocity profiles . The coordinate system and the configuration of the flow utilized in the experiment will be described. The measurement techn iques and procedures used to measure the velocity and the fluctuating pressure will be demonstrated. Moreover, the stepper motor and 2D traversing system used to traverse the hot - wire velocity probe over the measurement domain, the hardware and software us ed to coll ect experimental data, and the assembly of the experimental components will be explained . 4.1 Coordinate System and Flow Configuration This research is focused on the normal incidence impinging jet flow. The flow configuration is depicted in fig ure 4.1. Two coordinate systems are used . The first system is cartesian system with origin at the jet exit centerline, and the second is a p olar system originating at the center of the impingement plate and at (shown in figure 4.1 for the plane ). The impinging jet facility used in the curre nt experiments resides in the Flow Physics and Control Laboratory at Michigan State University. The facility underwent a major renovation and subsequent characterization by Al - Aweni [ 7 ]. The jet discharges through a round exit with diameter , at th e end of a contoured nozzle, resulting in an initial condition of a top - hat velocity profile. The flow from the jet imping es on a circular flat plate with diameter . The disc diameter is greater than the jet diameter by an order of magnitude to reduce the effect of the edge of the disc on the flow. The mean flow doe s not chang e in the direction and it i s assumed 72 to be axisymmetric (as verified in Al - Aweni [7]), and the distance between the jet exit and the impingement plate is adjustable. Figu re 4.1 F low configuration for normal - incidence impingement jet 4.2 Experiment General Assembly The experimental facility is presented in figure 4.2, and it consists of a centrifugal blower, type Dayton 4C108, driven by HP DC motor. The flow rate th rough the facility is adjusted by changing the speed of the motor . The blower provides provide air to the jet through a PVC pipe with a of diameter . The exit pipe diameter of the blower is smaller than , and it does not touch the inner diameter of th e PVC pipe to reduce the vibration effect from the blower . The air passes into a flow conditioning chamber with dimensions , before entering a nozzle with an area contraction ratio of 80, to decrease the turbulence intensity. The tur bulence intensity was measured at the nozzle exit using a hot - wire anemometer, 73 and was found to be less than 1% for the jet velocity of (based on the streamwise velocity fluctuations) Figure 4.2 Experiment al facility Referring to figure 4.3, th e impingement disc is embedded flush in a vertical square plate with dimensions of . The circular disc is fitted with 30 microphones for measuring the unsteady wall pressure, including eight microphones, employed in the current work, that ar e arranged as a line array along the radial direction (see section 4.4 for more details). The square plate is placed on a sliding table that can be moved in the x - direction using a manual traverse system ( as shown in figure 4.3 ) , model Velmex A1506P40 - S1.5 - TL, which allows to change the distance between the impingement plate and the nozzle exit ( H ). The Velmex traverse system can move total distance of 115 mm with an accuracy of 0.0254 mm . 74 Figure 4.3 An image of the impingement plate and the slid ing table attached to the Velmex manual traverse system A steel frame is used to hold the conditioning chamber, the nozzle, and the impingement plate. The frame is mounted on a table different from the one utilized to support the blower to isolate the fac ility from the vibration generated by the blower . In the course of this study, a stepper - motor - driven Velmex traversing system was added to the facility. The system was used to traverse a single hot - wire probe to characterize the jet velocity profiles nea r the nozzle exit (i.e. the initial condition) . The hot - wire wa s attached to the carriage of the Velmex system via a custom - made arm (see figure 4.4 and 4.5 ). The traverse system has three - degrees of freedom ( DOF ) : two - line a r and one rotation al ( model s MB4 027K2J - 56 and Square plate Circular impingement disc Velmex manual travers e system Hole with embedded Microphone s Slid ing table 75 B4872TS respectively ) . Only the linear DOF are used in the present work to move the velocity probe in X and Y directions. The traverse system is controlled by Velmex controller, type VP9000 , with the ability to control the system manually th rough a joystick, or programmatically through a n RS - 232 interface with a desktop PC computer. The jet mean exit velocity was determined from measurement of the difference between the stagnation pressure in the setting chamber and the ambient pressure (fur ther details are given in section 4.3) . Data acquisition was accomplished using National Instruments NI PCI - 6024E PC - based analog to digital ( A/D ) converter card. The A/D card was coupled with National Instruments BNC2080 analog breakout board to facilitat e signal connections via coaxial BNC cables . The card has 12 - bit resolution and input range that is changeable between to , and it can sample up to 16 single - ended multiplexed channels at a maximum rate of using LabVIEW version 8.2 . 76 4.3 Hot - wire Setup and Calibration A single hot - wire probe was used to measure the streamwise velocity p rofile of the jet near the exit of the nozzle. Figure s 4.4 and 4.5 depict a schematic of the hot - wire setup used during calibration and measurements , and an image showing the motorized traverse system respectively. Figure 4.4 Block diagram of the setu p used for calibration and measurements of the hot - wire 77 Figure 4.5 An image of the motorized system for traversing the hot wire The hot wire was built using tungsten with a sensing length and diameter of 1 mm and respectively, which provide a length to dimeter ratio of 200. The wire was used to measure the mean and the fluctuating streamwise velocity while operated using a Constant Temperature Anemometer ( CTA ) , model TSI 1750, at an overheat ratio , defined as: (4.1) Vertical movemen t Horizontal movement Hot - wire probe Jet exit 78 Where, represents the operating (heated) hot - wire resistance, and the hot - wire resistance at room temperature . The overheat ratio can be as high as 1 but it is generally kept between 0 .6 - 0 .8 . The higher the value, the better the velocity sensitivity and the smaller the temperature sensitivity. Too high of a value , however, could cause oxidation , and hence drift in the response of the wire . The bandwidth cut - off frequency of the hot - wir e was found using the following equation (4.2) Where, represents a response time , which was obtained from a square wave test. Figures 4.6 and 4.7 depict the diagram for the square wave test and the result respectively. During the tes t, square wave with frequency of was fed from Agilent WAVETEK function generator to the square wave input of the CTA . The output of the CTA was captured on a Tektronix TDS 1002B digital oscilloscope, as shown in figure 4.7. the figure also d emonstrates how was determined to be yielding a bandwidth of . 79 Figure 4.6 Block diagram of the setup for the square wave test Figure 4.7 An image of the oscilloscope screen showing a typical square wave test result 80 The hot - wire wa s calibrated in - situ by placing the wire near the center of the jet within the potential core where the jet velocity can be obtained from the difference between the stagnation and the ambient pressure. This pressure difference was measured using either hig h a (10 torr) or a low (1 torr) pressure transducer Baratron model 223BD - 00010ACU or 223BD - 00001ACU respectively. The pressure transducers have sensitivity of 0.75 mV/Pa and 7.5 mV/Pa respectively. The positive - input side of the pressure transducer was con nected via Tygon tubing to a pressure tap in the wall of the setting chamber, just upstream of the nozzle. The other, low pressure, side of the pressure transducer was left open to ambient pressure. Temperature of the air flow was measured sing a thermisto r, type Omega DP - 25 - TH, with a sensitivity of . The measured temperature was used to correct the hot - wire output for the variation of the flow temperature from that of the calibration . Figure 4.8 depicts an image of the hot - wire, temperature se nsor, stagnation pressure tap , nozzle, impingement plate, and the setting chamber. 81 Figure 4.8 An image of the hot - wire, temperature sensor, nozzle, impingement plate, and the conditioning box To acquire the pressure, the temperature, and the hot - wire signal, a LabVIEW program was developed for this purpose. Typically, the signals were acquired for eight different velocities depending on the velocity range in the experiments which varied between 5 to 15 m/s . All hot - wire voltages were corrected for the variation of the temperature during the period of calibration using (4.3) Hot - wire Setting chamber Temp. sensor Impingement plate Stagnation pressure tap 82 Where, represents the corrected hot - wire voltage, the measured hot - wire voltage, The hot - wire temperature calculated using equation ( 4.4 ) below, the flow temperature measured during the acquisition of data, the average temperature during the calibration process. The hot - wire temperature was found using (4.4) Where, represents the resistance - temperature coefficient (for tungsten 0.0045 º C - 1 ) , and is the ambient temperature. For the calibration, the j et velocity was found using (4.5) (4.6) Where, is the air density, the atmospheric pressure, the ideal gas constant for air , the a mbient air temperature , measured using the temperature sensor, and is the stagnation pressure in the settling chamber, measured using the pressure transducer . Note that is one over the contraction area ratio (1/80), which is neglec ted in the above calculation. The data pairs of the jet velocity and the corrected voltage of the hot - wire sensor were fitted with equation ( 4.7 ) - squares : (4.7) Where, r epresents the corrected voltage of the hot - wire, the measured velocity using the pressure transducer, and n (typically in the range = 0.4 to 0.45 ) , are the equation constants found using the least - squares method . 83 The calibration was done before and after the experiments. Figure 4.9 depicts a sample of typical calibrations after and before the experiments , which agree within a maximum difference of 0.8 %. Figure 4.9 Sample of hot - wire calibrations before and after an experiment 4.4 Mi crophone Setup and Calibration The fluctuating pressure was measured using eight electret Panasonic WM - 61A microphones embedded in the impingement plate , sensing hole and package diameter are 2 mm (0.08 D ) and 6 mm (0.24 D ) respectively. The microphones have a flat frequency response between 20 - 20,000 Hz and a typical manufacturer - provided sensitivity of @ 1kHz (corresponding to 11.22 to 28.18 mV/Pa ,) with DC supply voltage in the range 2 V and 10 V. The eight - microphone array was arranged starting from the center of the disc ( r/D =0) along the radial direction with a spacing of 0.33 D between the centers 84 of each successive microphones. Figure 4.10 depicts the configuration of the microphones on the impingement p late. The microphones were connected to a homemade 16 - channel microphone circuit powered by 9 DC volts to provide power to the microphone and connect the microphone to the NI - 6024E board . Figure 4.10 Front and cross section ( B - B ) view (top and bottom r espectively) of the microphone array configuration used in the present work Since the sensitivity of the microphones provided by the manufacturer is nominal , the individual microphones had to be calibrate d before each experiment. This also accounts for p ossible change in sensitivity due to variation in temperature, humidity, dirt, and installation. The eight microphones were calibrated, one at a time by using a plane wave tube ( PWT ) and a reference microphone that has known sensitivity. The reference micr ophone used for the calibration was 85 Brüel and Kjær ( B&K ) model 4923 - A - 001 with sensitivity of at a polariz ation voltage of . The microphone has a bandwidth of . The PWT was built by Al - Aweni [ 7 ] using a PVC square tube, which fastens to the impingement plate , by using two clamps , as shown in figures 4.1 1 and 4.1 2 . The tube has eight holes for mounting the B&K microphone at the same radial locations of the Panasonic microphones. On the other hand, the wall of the tube that is in contact with the impingement plate was removed allowing the Panasonic micrphone array to be embedded in the PWT . Acoustic waves were generated in the tube using Agilent model HP - 33120A function generator coupled to a Hafler - P100 0 amplifier, which drives a Dayton model RS150S - 8 audio speaker . The speaker was placed at one end of the plane wave tube , generating w hite noise acoustic waves in order to excite all frequencies of interest simultaneously inside the PWT . According to [24] , for example, if an acoustic wave with a wavelength ( corresponding to a frequency ; where represent the tube cross - section side length, the sound frequency, and the speed of sound respectively) propagates in a square so lid duct, then the wave will remain planar at any given cross section. This means, the phase and the magnitude of pressure will remain constant over the cross section at the given plane. So, the reference microphone ( B&K ) and the microphone to be calibrate d (Panasonic) will be exposed to the same pressure magnitude and phase because they are located at the same cross section. During calibration, each microphone was calibrated individually by inserting the reference microphone into the PWT at the same cross section of the Panasonic microphone . One hundred records, each having 4096 point data points sampled at 50 k Hz , were acquired from two A/D channels , to which the microphones were connected. Applying the analysis shown in Al - Aweni 86 [7], the acquired data pr oduced the magnitude and phase response of the Panasonic microphones ; as exemplified in figure 4. 1 3 . Figure 4.1 1 Block diagram of the calibration setup for the microphone array 87 Figure 4.1 2 An image of the calibration setup for the microphone a rray Table 4.1 show s a sample of the eight microphones sensitivit ies, where microphone 1 is located at the stagnation point and microphone 8 is placed at the end of the measurement domain . The values given, which fall within the manufacturer - reported nom inal range , are found by averaging the magnitude response over the frequenc y range 100 - 5000 Hz . Microphone 1 2 3 4 5 6 7 8 Sensitivity mV/Pa 18.625 18.447 23.789 19.623 18.837 22.652 25.666 25.624 Table 4.1 Sample of the microphones sensitivity obtai ned from calibration . B & K microphone Plane wave tube Speaker Nozzle Impingement plate 88 Figure 4.1 3 A sample of the microphone calibration results: sensitivity (top) and phase (bottom) 89 4.5 Procedure to Acquire Velocity Profile After calibration, the hot - wire was placed near to the edge of the shear layer. The latter was found approximately by monitoring the hot - wire signal on an oscilloscope as the wire was traversed across the shear layer. Once the edge of the shear layer was found , the wire was traversed to 50 to 100 different positions (with a resolution of ), depending on the X location, such that measurements were conducted across the entire shear layer. The movement and the data acquisition were automated using a LabView program. At each location, time series containing 409600 data point s of jet veloci ty , jet temperature, and hot - wire signal w ere acquired at 5000 samples/second . The duration of the acquisition is such that almost 5000 vortices travel across the measurement point at the lowest frequency of vortex passage ( St = 0.3 and the low est velocity of ). The sampling rate is also selected to be much larger than any frequency of interest ( , based on the highest velocity). 4.6 Procedure to Acquire the Microphone Signal s After calibration of the microphones, the PWT was removed and the impingement plate was fixed at a desired distance ( H/D ). Before acquiring the microphone signal s , the mean jet velocity was fixed to provide the desired Reynolds number. Once the jet reached steady state, 800 data records contai ning 512 - point per record were acquired at a rate of 5000 samples/s from all microphones in the array using a LabVIEW program. 90 CHAPTER 5: Jet Characteristics and Wall - Pressure Measurements Before investigating the pressure measuremen ts, it is important to demonstrate the jet characteristics as it emerges from the nozzle; i.e. the initial condition. This chapter will explore the characteristics of the jet and the wall - pressure measurements. The jet characteristics were found by measur ing the streamwise flow velocity component using a single hot wire, traversed across the shear layer at a few streamwise locations. The data were used to demonstrate the self - similarity in agreement with the literature. T he pressure measurements from this study were compared with previous studies using different statistical quantities, including power spectra , root mean square of the pressure fluctuation, probability density functions ( pdf ), skewness, and kurtosis. 5.1 Ini tial Shear Layer Self - Similarity It is important to examine the characteristics of the shear layer and the initial flow conditions. To achieve this goal, the mean and fluctuating streamwise velocit y were measured, and the data were used to obtain the cor responding cross - stream velocity profiles across the shear layer. For these measurements, the hot - wire was initially moved using relatively large steps, with a resolution of 0.5 mm/step , in the transverse direction and the wire output signal was monitored on the oscilloscope to approximately locate the edges of the shear layer. Subsequently, the movement resolution was refined gradually to verify the location of the edges of the shear layer. Once the edges were found , the probe was controlled to traverse ac ross the whole shear layer with even higher resolution (0.01 mm/step ) to properly resolve the high - shear zone within the shear layer. The data were recorded at Reynolds numbers of 8272, 165454, and 33090 which represented velocities of 5 m/s , 10 m/s , and 2 0 m/s , respectively. Profiles were obtained at three different 91 streamwise locations ( X/D = 0.2, 0.4, and 0.8). The self - similarity of the profiles obtained at different Reynolds numbers was verified using normalized mean and fluctuati ng velocity plots, as s hown in figure s 5.1 and 5.2 respectively. For these plots, the origin of the cross - stream coordinate is taken at the shear layer centerline; defined as the y location where the mean velocity is half of the jet exit velocity. The y coordinate is normalized by the momentum thickness , which is found using (5.1) Where is the mean streamwise velocity profile, and is the jet exit velocity. To minimize the erro r resulting from hot - wire data near the shear layer outer edge, where reverse velocity may occur due to the energetic shear - layer vortices , the momentum thickness was calculated by truncating the integration limit to the location where is 10% of the jet velocity; i.e. the lower integral limit in equation 5.1 . As seen from figure s 5.1 and 5.2, the velocity profiles for the various Reynolds numbers at the same streamwise location ( X/D =0.2) collapse well. The collapse of the veloci ty profiles demonstrates the initial self - similarity of the jet. 92 Figure 5.1 Shear - layer mean velocity profile at X/D =0.2 for various Reynolds numbers Figure 5.2 Shear - layer fluctuati ng - velocity root - mean - square profile at X/D =0.2 for various Reynolds numbers 93 To check the self - similarity at different X/D locations, mean and fluctuati ng - velocity profiles of the shear layer were measured at three streamwise locations ( X/D = 0.2, 0.4, and 0.8) for Reynolds number of 8272. The results are depicted in figure s 5.3 and 5.4 respectively, which demonstrate that the velocity profiles collapse for the distance X/D =0.2 - 0.8, further confirming self - similarity. Figure 5.3 Shear - layer mean velocity profile at various X/D locations and 94 Figure 5.4 Shear - layer fluctuati ng - velocity root - mean - square profile at various X/D locations and 5.2 The Root Mean Square of the Fluctuating Pressure The root mean square pressure is calculated from the pressu re time series using (5.2) Where, is the instantaneous pressure, is the mean pressure, i indicates the time index, or sample number in the digitized pressure time series, and n is the total number of samples in the time series. The number of samples is 409600 samples with a sampl ing frequency of 5000 samples/sec . Note that since microphones are used for measuring the pressure, they are incapable of capturing . However, the measured time series typ ically have a small offset voltage error, 95 Figure 5.5.a depicts the radial distribution of the root mean square results for the pressure fluctuation for Reynolds number 82 72. The fluctuati ng - pressure RMS value is normalized by the dynamic pressure and the radial location ( r ) is normalized by the jet dimeter ( D ). Results are shown for different H/D values, represented by different colors. The re sults show that, though the overall shape of the root mean square pressure distribution is similar for all H/D values, the magnitude depends on H/D . A peak is noticed for all three H/D = 2, 3 and 4 cases at r/D 1.33. The magnitude of this pea k, and in the wall - jet zone in general, decreases with increase in H/D . Also, the location of the maximum seems to shift towards smaller r/D with increase in H/D of the maximum cannot be determined with the present measurement resolution. On the other hand, within the stagnation zone , the trend with H/D is reversed, where the magnitude of the root mean square pressure increases, rather than decreases, with increasin g H/D . For example, the magnitude of at the stagnation point ( r/D = 0) increases to 54% of the maximum for H/D= 4, in contrast to 14% at H/D= 2 . Also noteworthy , for all H/D values, the normalized drops rapidly for the range to a magnitude of around 5% by the end of the measurement domain. Figure 5.5.b depicts for the higher Reynolds number of 24818. The distribution has the same behavior as for Re D = 8272, except for H/D =4, where a second peak emerges wi thin the stagnation zone, at r/D =0.67. Significantly, unlike all other cases, at the higher Reynolds number and H/D = 4, the maximum is found in the stagnation, instead of the wall - jet zone. The magnitude of at the stagnation point increases to 42% of the maximum for H/D =4, in contrast to 12% at H/D =2. 96 ( a ) ( b ) Figure 5.5 Normalized Root Mean Square Pressure versus for different and two Reynolds numbers: (a) Re D = 8272, and (b) Re D = 24818 97 To better examine the effect of Reynolds number, Figure 5.6.a depicts a comparison between the radial distributions for Re D = 8272 and 24818 when H/D =2. The figure shows that as Reynolds number increases, the profile does not change, for all practical purposes. For both cases, the minimum pressure is at the stagnation point, where r/D = 0, and the maximum pressure is measured at r/D =1.33. Figures 5.6.b and 5.6.c depict a similar comparison for H/D =3 and 4 , respectively. For both cases, increasing the Reynolds number, results in decreasing the level of pressure fluctuations within both the stagnation and wall - jet zones. In addition, as noted previously, a second local peak emerges at r/D = 0.67 for the higher Reynolds number and . It is unclear if this peak reflects a change in the physics of wall - pressure generation within the stagnation zone at the higher Re D , or simply that the peak becomes observable due to reduction on the level of pr essure fluctuations in the wall - jet zone. Overall, the characteristics of the RMS profiles and how they change with and is very consistent with those reported in the literature, and summarized in section 1.2.2. 98 Figure 5.6 Normalized Root Mean Square Pressure versus at Re D = 8272 and 24818 , for : and ( a ) ( b ) ( c ) 99 5.3 Time Series Analysis 5.3.1 Time Series Figure 5.7.a depicts sample normalized - pres sure signals at Re D = 8272 at the stagnation point ( ) and the location of maximum ( ) for As expected from the distribution (figure 5.5 . a), the pressure signal at r/D =0 is low compared to that at r/D =1. 33. Aside from this, inspecting the time series enables extraction of additional interesting information. Specifically, Figure 5.7.a demonstrates that the signal shape is completely different between the stagnation point and r/D = 1.33. At the former locat ion, the signal is relatively symmetric around zero level and looks like a distorted sinusoid. In contrast, at r/D = 1.33, the dynamic pressure! In his study in the same jet facility at a similar Reynolds number and H/D , Al - Aweni [7] used simultaneous time - resolved flow visualization and wall - pressure measurements and numerical simulations to demonstrate that these very strong pressure spikes are a result of t he interaction of the jet vortices with the wall and the formation of secondary vortices (see section 1.2.2) . Similar strong negative spikes can also be seen at the same r/D = 1.33 location but the larger H/D values of 3 and 4 (figures 5.7.b and 5.7.c res pectively). However, a significant difference between the signals observed at the three H/D values is that the average time period between spikes increases with increasing H/D . Based on the flow visualization of Al - Aweni [7], it is known that this increase is due to two successive pairings of the jet vortices before reaching the impingement wall: one pairing taking place between H/D = 2 and 3, and the other between H/D = 3 and 4. Interestingly, the magnitude of the spikes in figure 5.7 is almost unaffected by the H/D values, suggesting that the strength of the pressure produced from the vortex - wall interactions is 100 maintained with increasing H/D . These observations lead to the following explanation for the decrease in the maximum (and likely in general for the whole wall - jet region as well) with increasing H/D (see figure 5.5): with the strength of the spikes remaining invariant with H/D but becoming less frequent, the value must decrease. Figure 5.7 S ample normalized Pressure signals at Re D = 8272 , for for: and ( a ) ( b ) ( c ) 101 Figure 5.8 displays plots similar to those in figure 5.7 for the higher Reynolds number. Generally spea king, observations similar to those made in relation to figure 5.7 can be made from figure 5.8. However, there are also some notable differences. One of these relate to the strength of the negative pressure spikes. Unlike the time series for Re D =8272, the strength of the negative pressure spikes at r/D = 1.33 decreases noticeably with increasing H/D for Re D =24818. This decrease is associated with the signal becoming more irregular and the appearance of high - frequency fluctuation. This is consistent with the decrease in the level with increasing Re D observed in figures 5.6.b and 5.6.c at H/D = 3 and 4 respectively. The signal forms in figure 5.8.b and figure 5.8.c suggest that this decrease is associated with weakening of the pressure spikes at th e higher Reynolds number and larger H/D , which might be related to the jet vortices breaking up and becoming irregular/turbulent with increasing Reynolds number (as implied from the irregularity of the signal and appearance of high frequency fluctuations). 102 Figure 5.8 Sample normalized Pressure signals at Re D = 24818 , for for: and ( c ) ( b ) ( a ) 103 5.3.2 Probability Density Function (pdf) The t ime series analysis in the previous section shed s light on the general characteristics of the pressure signal at the stagnation point, as well as where is highest in the wall - jet zone. However, these observations are based on short, randomly se lected time series samples. Therefore, to ensure that the observations made in section 5.3.1 are statistically relevant, Probability Density Function ( pdf ) results are examined. The pdf was estimated by finding the maximum and minimum fluctuating - pressure values in a given time series. The range bound by these values was then divid ed into 30 equal - width bins . Finally, the number of data points falling in each bin was divided by the total number of points in the time series and the bin width to obtain the pr obability of the pressure value occurring within a given bin per bin width, or the pdf . Figure 5.9 depicts the pdf for the pressure time series at Re D = 8272 and for all three values. By examining the plots, we can see that a t , the pdf is approximately symmetric and narrow, which is consistent with the symmetric, low - level character of the corresponding signals observed in figure 5.7. On, the other hand, at , the signal has a pronounced negative skewness , as reflected in the long negative tail and consistent with the strong negative pressure spikes noted earlier in figure 5.7. In addition, the pdf is substantially wider than that at , consistent with the smaller at ( figure 5.5.a). Another interesting observation at H/D = 2, is the presents of a plateau with a hint of a peak at . This suggest the presence of a bi - model phenomenon. Interestingly, Al - Aw e ni[7] found the vortex structures to either merge as they convect past or to pass without merging. This may explain the subtle bi - modal feature of the at in figure 5.9.a 104 At , the width of the pdf increases with increasing H/D , demonstrating the increasing lev el of at the stagnation point with larger H/D (see figure 5.5.a). At , the pdf remains negatively skewed with increasing H/D , approximately reaching negative pressure values as high as P d for all H/D values. However, the p df appears to become overall narrower with increasing H/D , consistent with the corresponding reduction in (see figure 5.5.a). The narrowing of the pdf primarily manifests itself in the reduction in the pdf value for the large negative pressure spikes ( approximately less than ), which reinforces the idea discussed in section 5.3.1 of these spikes becoming less frequent at larger due to vortex pairing. 105 Figure 5.9 Probability Densit y Function for the pressure signal at Re D = 8272 , for , and: and Figure 5.10 depicts the results for Re D = 24818 at , and all values. By exami ning the se plots, we can see the same general behavior as seen for Re D = 8272. However, there are some notable differences. Overall, the pdf at r/D = 1.33 is not as strongly skewed as for the lower Reynolds number case. Additionally, the long negative tail of the pdf ( a ) ( b ) ( c ) 106 extends to smaller negative - pressure magnitudes with increasing H/D ; in contrast to reaching approximately the same value for Re D = 8272. The reduction in the magnitude of the negative spikes with increasing H/D was also noted earlier from the time series plots in figure 5.8. Figure 5.10 Probability Density Function for the pressure signal at Re D = 24818 , for , and: and ( a ) ( b ) ( c ) 107 5.3.3 Skewne ss and Kurtosis The overall features of the pdfs presented in section 5.3.2 may be expressed in terms of their skewness and kurtosis. The skewness provides a measure of the symmetry of the pdf . A skewed pdf exhibits a long negative tail (if negatively s kewed) or a positive one (if positively skewed). On the other hand, the kurtosis indicates how flat (as opposed to having a prominent peak ) a pdf distribution is. A pdf with large kurtosis tends to have long tails. For reference, a Gaussian pdf has a skewn ess of zero (due to its symmetry) and a kurtosis of 3. In the present work, the skewness and kurtosis were calculated as follows (5.3) (5.4) where, is the instantaneous pressure, is the mean pressure, i indicates the time index, or sample number in the digitized pressure time series, is the standard deviation of the pressure time series points , and n is the total number o f samples in the time series Figure 5.11.a depicts the variation of skewness over the measurement domain ( r/D range 0 to 2.33) for a Reynolds number of 8272 at H/D = 2, 3, and 4. The plots show that at H/D =2, the pressure signal has positive skewness in th e stagnation zone ( ), but it switches sign and becomes negative in the wall - jet zone. For H/D =3 and 4, the switch from positive to negative skewness happens near the end of the stagnation zone . The negative skewness is particularly strong at r/D = 1.33 and 1.67. This is consistent with the long negative tail of the pdfs and the strong negative pressure spikes in the time series discussed earlier for r/D = 1.33. However, as r/D increases further, the skewness magnitude decreases monotonically; thoug h it remains negative. 108 Figure 5.11.b depicts the skewness results for the higher Reynolds number of 24818. With the exception of one apparently errant data point (at r/D = 0.33 and H/D = 4), the general qualitative behavior of the skewness distribution is similar to the lower Reynolds number. However, at Re D = 24818, the largest negative skewness is not as large in magnitude as for Re D = 8272. Also, the skewness becomes zero (or low valued) by r/D = 2.0; implying the pdfs reach symmetry by the end of the me asurement domain. 109 ( a ) ( b ) Figure 5.11 Radial distribution of skewness for and and: , and 110 Figure 5.12.a depicts the kurtosis results for Reynolds number of 8272. W ithin the stagnation zone ( ), the kurtosis is similar to that of a normal (Gaussian) distribution with a magnitude near 3. In the wall - jet zone, the kurtosis initially increases substantially (up to the radial location r/D = 1.67) then it decays mon otonically. For the higher Reynolds number, the kurtosis behavior (figure 5.12.b) is surprisingly different. In this case, the kurtosis is highest at the stagnation point and decays monotonically with increasing r/D . The reason for the fundamental change i n kurtosis behavior with Reynolds number is not clear. Measurements at intermediate Reynolds numbers would be recommended in order to observe if this change is gradual or abrupt , and attempt to understand the reasons behind it. The large magnitude of negat ive skewness and kurtosis near the location of maximum in the wall jet zone indicates the presents of strong pressure spikes. This is consistent with the strong negative peaks found in the work of Didden and Ho [5], Hall and Ewing [ 9,10 ] and Al - Aw e ni [7]. The latter study connected the formation of the spikes to the secondary vortex formation (as also Didden and Ho [5] and Hall and Ewing [ 9,10 ]), and the formation of high strain zone in the boundary layer beneath the jet vortex. 111 ( a ) ( b ) Figure 5.12 Radial distribution of Kurtosis for and and: , and 112 5.4 Spectra Analysis Analyzing spectra of the pressure fluctuations is considered one of the useful tool s to get information about the frequency content of the pressure signal. In this work, spectral information is presented as power spectral density ( PSD ). The PSD is calculate d as an average of the PSDs obtained from different pressure data records, as f ollows: (5.5) Where, is the Fourier transform of the i th pressure data record, is the number of points in the data record , is an index indicating frequency and f s is the sam pling frequency . The physical frequency corresponding to each value is given by (where is the frequency resolution of the PSD ). To compute the PSD , 409600 samples of a given pressure signal were acquired at a sampling frequency o f 5000 Hz . The resulting data were divided into 800 records , each containing 512 points. The PSD was obtained for each record by taking its Fast Fourier Transform ( FFT ), multiplying the transform by its conjugate, and dividing the result by the square of t he number of points (512 points) and the spectrum frequency resolution ( = 5000/512 = 9.77 Hz). The resulting spectrum random uncertainty is 3.5%. Figure 5.13 shows the normalized PSD in the form of contour plots versus r/D and Strouhal number ( St D = fD /U j ) at and all values. This way of presenting the spectra provides a global perspective of the entire measurement domain, but it does not allow clear observation of some of the less dominant spectral features. To see these, the spect ra are presented using line plots at r/D = 0 and 1.33 in figure 5.15. The contour plots in Figure 5.13 show a 113 dominant spectral peak at (the actual value varies between 0.64 for H/D = 2 and 3, and 0.58 for H/D = 4, which is approximately wit hin the spectrum resolution of 0.05). From the work of Al - Aweni [7] , it is known that the corresponds to that of the jet vortices after the first pairing. As discussed in section 5.3.1, Al - Aweni also found that for approximately th e same Reynolds number, the first vortex pairing took place as the vortices traveled parallel to the wall, within the wall - jet zone, at H/D = 2, and ahead of reaching the impingement plate, at H/D = 3. A second pairing was observed before the vortices reac hed the plate at H/D = 4. These conclusions suggest that the spectrum should be dominated by fluctuations at 0.6 at H/D = 2 and 3, and at H/D = 4. The former expectation is consistent with the results in figure 5.13.a and 5.13.b. H owever, the dominance of 0.6, instead of 0.3 at H/D = 4 (figure 5.13.c) seems inconsistent with Al - fact that the dominant spectrum peak in figure 5.13.c is seen within the wall - jet zone ( r/D > 1). Inspecting a sample of the corresponding time series (figure 5.7.c), it is evident that the time series is highly irregular and dominated by strong negative spikes with varying strength (i.e. modulation) from one spike to anoth er. The spectrum of such a signal is not expected to yield a clean peak at the dominant frequency, but rather a broader spectrum of multiple peaks. Indeed, although is dominant in figure 5.13.c, another, barely visible peak is seen at 0.3 (pointed to by a white arrow in figure 5.13.c ). This peak can also be observed more clearly in the line plots of figure 5.15.b. Although the observed peak is weaker than that at , this could be caused by the vortices being less coheren t after the second merging (which was observed in the flow visualization videos of Al - Aweni). However, this reasoning cannot be ascertained at this stage. 114 A better indication of the presence of pressure fluctuation at the second - pairing frequency ( 0.3) at H/D = 4 may be seen in the stagnation zone, particularly at r/D = 0, in figure 5.13.c. Due to the simplicity of the signal at the stagnation point (figure 5.7.c), the corresponding spectrum has a more straight forward interpretation. The contou r plot in figure 5.13.c and the line plots in figure 5.15.a show that indeed when H/D = 4, the dominant spectral peak shifts to 0.3 at r/D = 0. 115 Figure 5.13 Normalized PSD contour plots at for : , and Figure 5.14 shows normalized PSD contour plots similar to those in figure 5.13 but for . Consistent with the lower Reynolds number, the spectrum peak at is dominant within t he wall - jet zone at H/D = 2 and 3. At H/D = 4, the dominant peak clearly shifts to within the stagnation zone at . Recall that this is also the radial location ( a ) ( b ) ( c ) 116 where the maximum is found for the larger Reynolds number an d H/D = 4 (figure 5.5.b), which is different from all other cases. These results reinforce earlier observations regarding the overall weakening of the pressure fluctuation in the wall - jet zone with increasing H/D and Reynolds number. Since it is known from the work of Al - Aweni that the vortex - wall interactions dominate the pressure fluctuations in the wall - jet zone at low Reynolds number, the present results suggest that the ability of these interactions to generate unsteady pressure weakens with increasing Reynolds number, leading to the observed dominance within the stagnation zone. Another notable characteristic of the spectra at the higher Reynolds number is the rise of low - frequency fluctuations. These are seen clearly in both the contour plot in figure 5.14.c and the line plots in figure 5.16. The latter plots also show the general shift of the dominant frequency towards low values with increasing H/D (figure 5.16.a). 117 Figure 5.14 Normalized PSD contour plots at for: , and ( a ) ( b ) ( c ) 118 ( a ) ( b ) Figure 5.15 Normalized PSD at for , and at: 119 ( a ) ( b ) Figure 5.16 Normalized PSD at for and at: 120 Figures 5.17 and 5.18 depict a direct comparison of the normalized PSD at the two different Reynolds number ( . These plots are provided to facilitate understanding of the Reynolds number effect on the spectra. Overall, both figures show that the Strouhal number band of the pressure fluctuations is consistent for both Reynolds number. However, the PSD has hi gher level and sharp peaks at the lower Reynolds number; in comparison to being broader and having lower level with increasing Reynolds number. This suggests that the basic wall - pressure generating mechanisms remain the same with increasing Reynolds number , but they become weaker and more stochastic in nature. In addition, at the largest H/D of 4, and the high Reynolds number , the vortex - wall interaction effectiveness in generating pressure in the wall jet zone weakens substantially. 121 Figure 5.17 Normalized PSD at for: and ( a ) ( b ) ( c ) 122 Figure 5.18 Normalized PSD at for: and ( a ) ( b ) ( c ) 123 5.5 Comparison between Experiment and Mathematical Model In the present section, a comparison is conducted between the results of the experiments and the model. The purpose of this comparison is to assess t he degree by which the model is successful in capturing the underlying physics of wall - pressure generation in impinging jets. It is emphasized here that, given the fairly crude nature of the model, the comparison is focused on qualitative features and tren ds and is constrained to the stagnation zone of the impinging jet. The discussion in section 3.1 demonstrated that the character of the modeled wall - pressure time series is different depending on whether the radial location is near the stagnation point ( ) or the end of the stagnation zone ( ). Near the stagnation point, the pressure signal was weaker and characterized with sinusoidal like variation, and the end of the stagnation zone, the signal featured a prominent negative pressure peak and a strong, but less prominent positive peak. The pressure variation (which are replicated at the bottom of figure 5.19) were connected to ) effects of vortex passage. Comparing similar signals obtained expe rimentally at the lower Reynolds number and (top of figure 5.19), we see very similar qualitative features, suggesting that he basic physics of wall - pressure generation in the stagnation zone are captured by the present model, notwithstanding its h igh level of simplicity. Some of the differences between the experimental and model results in figure 5.19 include the broadness of the positive and negative peaks at . The experimental time series clearly exhibit much broader peaks, which is not su rprising given that the real jet vortices have a finite core size, in comparison to the point vortices employed in the present model. Thus, one of the important future improvements of the model is to utilize finite - core vortices; for example, Oseen type, h aving Gaussian vorticity distribution. This point is also important from the perspective that 124 depend on how far is the point from the vortex center, relative to the vortex core size . 125 Figure 5.1 9 Sample time series from microphone measurements (top) and the vortex - array model (bottom) at the stagnation point ( ) and end of the stagnation zone ( ). Experimental data are shown for the and , and model results for the reference case (emulating the conditions at the end of the potential core). 126 Though shown over a short time period, the experimental time series in figure 5.19 is characteristic of the remainder of the time series, with the shown signature repeating quasi - periodically. However, at , there are moments in time were the signature character is different. This is exemplified in figure 5.20 for the same time - window size as the top of figure 5.19. The s ignal shape during such periods is very similar to that associated with the vortex - induced separation of the boundary layer. This conclusion can be made based on the work of Al - Aweni [7]. Such vortex - boundary - layer interactions cannot be captured by the pr esent model, which is both inviscid and does not include any modeling of boundary layer effects. Figure 5. 20 Sample wall - pressure signature characteristic of that produced by vortex - induced separation from microphone measurements at , and Another point concerning the comparison in figure 5.19 is that it is done using the experimental data at the lower Reynolds number. At the higher Reynolds number, as discussed previously, small - scale pressure fluctuations start to appear, implying the formation of small - scale 127 turbulence. Such turbulence, which obviously cannot be reproduced by the present model, makes it more difficult at times to identify the vortex - passing signature in the time series. Another interesting qualitative fea ture of wall - pressure fluctuation in the stagnation zone that is captured by the vortex - array model relates to the effect of on . The reader is reminded that the effect of increasing in the model is simulated by decreasing the vortex passing frequency; i.e. emulating the reduction in frequency via vortex pairing. This frequency effect, which was presented in figure 3.4 and is reproduced in the bottom of figure 5.21, is qualitatively similar to that seen in the experimental results (fi gure 5.21 top). At high frequency (small ), both model and experimental results show the pressure fluctuations to be lowest at the stagnation point and rise to be highest at the end of the stagnation zone. As the frequency decreases ( increases), the level of pressure fluctuations increases at the stagnation point. Also both the model and the experimental results show that within the zone , the pressure fluctuations increase with at highest - frequency (smallest ), while they dec rease with for the middle and lowest frequency ( ). For all cases, the RMS pressure level increases with when (within the stagnation zone) for both the model and the measurement results. 128 Figure 5. 21 Comparison of the effect of varying on RMS wall - pressure fluctuation between the experimental (top) and the model (bottom) results at . In the model, increasing frequency corresponds to decreasing . The broken green lines outline the end of the stagnation zone 129 Figure 5.21 also show consistency between the calculation and the measurement in that the trend of increasing RMS level with increasing (decreasing frequency) at the stagnation point, reverses by the end of the stagnation zone. T he switching point, which happens at a specific location in the model results, happens at different locations in the measurements, depending on which two cases are considered. This difference in the location of switching might be due to t he increasing vortex core size with increasing (due to vortex pairing); an effect that is not captured in the present model but need to be included in future development of the model. Finally, f igure 5. 22 demonstrates the general consistency between the experimental (top plot ) and the model (bottom plot ) results regarding Reynolds number effect . In both cases, as Reynolds number increases , the fluctuati ng pressure level decreases . Based on the discussion of the model results in section 3.2, the overal l decrease in the normalized RMS pressure fluctuation with Reynolds number is predominantly due to the increase of the normalization scale (the jet dynamic pressure) while the level of pressure fluctuation remaining invariant. Of course in the real jet it without affecting the strength of the vortices, and hence the wall - pressure fluctuation. However, a more realistic interpretation of the clue from the model results is that perha ps the strength of the jet vortices increases with hypothesis: the vortex strength (vorticity/circulation) is expected to increase in proportion to the shear in th e separating shear layer, which is expected to scale as the viscous shear stress in the separating boundary layer. The latter is expected to be proportional to , based on laminar boundary layer theory (Al - Aweni [7] showed that the boundary lay er at the jet exit is laminar over the Reynolds number investigated). In contrast, the dynamic pressure should increase as . Thus, if the wall pressure increases in proportion to the square of vorticity/circulation (based on 130 the pressure source term in equation 1.7) of the jet vortices, the normalized pressure would decay as . This could explain the decay of the normalized pressure fluctuation with Reynolds number. 131 Figure 5. 2 2 Comparison of the effect of varying on RMS wall - pressure fluctuation between the experimental (top) and the model (bottom) results at (experiment) and (model) . The broken green lines outline the end of the stagnation zone 132 CHAPTER 6: Conclusion s and Recommendation s The present investigation is focused on studying the unsteady surface pressure fluctuation in jets impinging normally on a flat wall. The study is divided into two parts : t he first part is concerned with developing a simple physics - based mathem atical model of the impinging - jet wall pressure, and the second part involves measurements of the unsteady wall pressure in an existing impinging jet facility. The intent of developing the mathematical model is twofold: (I) as a first step in a multi - step pr ocess of developing a high - fidelity, efficient model that may be used as a design tool for predicting wall - pressure fluctuation for problems involving flow - induced noise and vibration by impinging jets; (II) to utilize the model for understanding the connectio n between the characteristics of the jet vortex structures and the surface pressure by varying the main jet and vortex parameters one at a time, and investigating the influence of this variation on the wall - pressure characteristics. As the first - step in th e development process, the present model is very simple, consisting of an array of potential, point - vortex rings that are advected under their own influence and that of their image vortices (due to the presence of the impingement wall), in addition to a st eady advection field (emulating the mean - jet flow) consisting of potential stagnation point flow. This at each time instant as the vortex rings advect periodica lly towards the wall then radially outwards. In comparison to the real jet, the present model does not account for several significant phenomena. These include the viscous core of real vortices, vortex pairing, vortex - boundary layer interaction, and the sp ecific mean advection field of the jet flow. As such, the model is not expected 133 to be useful, in quantitative or qualitative sense, in characterizing the pressure fluctuations outside the stagnation zone of the impinging jet, where vortex - wall interaction is known from literature to play a significant role in wall - pressure generation. However, the model results are expected to be qualitatively consistent with the characteristics of the wall - pressure fluctuation within the stagnation zone. Establishing this point, which is done here using experimental data from the second part of the investigation, would provide the necessary initial confidence to continue the development of the model in the future. The model was utilized to examine the effects of changing th e vortex - passing frequency, jet Reynolds number, and the vortex circulation on the fluctuati ng wall - pressure in the stagnation zone. The effect of varying the frequency effectively corresponded to varying the spacing between the jet exit and the impingemen t plate ( ). Overall, the model results revealed that the stagnation zone pressure fluctuation are either low - level, sinusoidal - like and rather symmetric for , or energetic, featuring strong positive and negative peaks, with the negative peak being more prominent, for influences remained fairly invariant with , local effe cts increased with increasing because of the increased proximity of the vortices to the wall as they convect radially outwards. For , a s the frequency increase d ( decreased) , was found to decrease due to the increasing pac king of vortices near the wall within the stagnation zone . The corresponding decrease in the inter - vortex spacing caused the induced flow along the stagnation streamline (which is parallel to the vortex array) by the individual vortices to overlap , reducing the vortex - to - vortex velocity and pressure fluctuation . An opposing frequency ( ) trend was found on locally produced fluctuation ( ), where increased vortex packing/passage frequency 134 resulted in increasing . This opposite effect was connected to the broadening of the induced velocity peaks on the wall under neath (which is normal, rather than parallel to the array) with the packing density of vortices. When changing the vortex circulation ( ) and the Reynolds number ( ) , the overall radial distribution of remained qualitatively the same. The main influence of increasing these two parameters was to either increase ( or decrease ( ) the strength of the pressure fluctuation relative to the dynam ic pressure of the jet. In t he second (experimental) part of the investigation, the unsteady wall pressure was measured at two Reynolds number s ( ; based on exit jet velocity and the jet diameter ) for a jet at norm al impingement incidence. The pressure was measured using an array of eight microphones over a radial domain range for three separation distances between the jet exit and the i mpi ngement wall : The result s yielded radial distributions of that were consistent, both in their shape as well as their trends with and , with the literature . Within the stagnation zone, several observations showed good qualitative agreement with the mathematical m odel results. These include, the wall - pressure time series features at versus , and the trends in the distribution with both and . On the other hand, it appeared that discrepancy in finer qualitative details be tween the experimental and the model results may be accounted for by including a vortex viscous - core in the model. Analysis of the experimental data alone, using time series and power spectra , suggest ed that the basic wall - pressure generating mechanisms r emain the same with increasing Reynolds number. However, the overall weakening of the level of pressure fluctuation with Reynolds number was inferred to be due to two reasons. First, the slower increase of the pressure source 135 strength (square of vorticity/ vortex circulation) with pressure ( respectively). Second , the weakening of the vortex - wall interaction effectiveness in generating pressure in the wall jet zone . This was hypothe sized to be related to the jet vortices breaking up and becoming irregular/turbulent with increasing Reynolds number . This was implied from the appearance of small - scale random fluctuations in the pressure time series, and the irregularity of the negative pressure spikes at the higher Reynolds number (particularly at ) . Another interesting observation from the experimental data was the fact that the strong negative pressure spikes in the wall - jet zone did not weaken with increased , notwithstan ding that the overall pressure RMS level decreased with . This was observed in both time series as well as in the wall - pressure probability density functions and skewness results. Since it is well understood in the literature that these spikes are pro duced during vortex - wall interaction, and given that the passage frequency of vortices decreases with due to vortex pairing, it was concluded here that the decrease in RMS is due to the reduction in frequency of the spikes. Overall, the results of t he present investigation, in addition to providing some new insights into the connection between the unsteady wall pressure and the jet vortical structures , establishes an encouraging first step towards developing a physics - based model of impinging jets wa ll - pressure fluctuation . However, as a first step, the present model is very simple and it does not include phenomena that are known to be important for wall - pressure generation: vortex - vortex interaction, impingement - plate boundary layer, vortex - boundary layer interaction, viscous - core calculate the wall pressure. Th ese elements should be added in future development of the model . The present study suggests that, as far as the stagnation zone is concerned, the n ext two highest priority elements would be the inclusion of 136 viscous - core vortices and the use of a more realistic advection field. Both of these items are expected to not only enhance the qualitative agreement with physical observations in the stagnation z one, but to possibly also lead to reasonable quantitative comparisons. 137 APPENDIX 138 Appendix: Uncertainty Calculation 1 - Finding the uncertainty for the root mean square pressure [ (see 28)] Where represent s the uncertainty and represent s the number of independent sample s and was found assuming to be equal to the number of vortices passing during the measurement period. Thus, Where represent s the time period between t he passage of successive vortices, and 139 2 - Finding the uncertainty for the third order moment ( s kewness ) , assuming Gaussian variation [(see 28)] More general ly Finding by a ssum ing 140 Where Finding and assuming Gaussian variation 141 3 - Finding uncertainty for the fourth order moment ( k urtosis ) , assuming Gaussian variation and following the same procedu re for calculating the uncertainty of skewness [(see 28)] 142 BIBLIOGRAPHY 143 BIBLIOGRAPHY [1] Pai, S. I. (1954). Fluid dynamics of jets. [2] Gra nt, A. J. (1974). A numerical model of instability in axisymmetric jets. Journal of Fluid Mechanics , 66 , 707 - 724. [3] Popiel, C. O., & Trass, O. (1991). Visualization of a free and impinging round jet. Experimental Thermal and Fluid Science , 4 (3), 253 - 264. [4] Zaman, K. B. M. Q., & Hussain, A. K. M. F. (1984). Natural large - scale structures in the axisymmetric mixing layer. Journal of Fluid Mechanics , 138 , 325 - 351. [5] Didden, N., & Ho, C. M. (1985). Unsteady separation in a boundary layer produced by an im pinging jet. Journal of Fluid Mechanics , 160 , 235 - 256. [6] Landreth, C. C., & Adrian, R. J. (1990). Impingement of a low Reynolds number turbulent circular jet onto a flat plate at normal incidence. Experiments in Fluids , 9 (1 - 2), 74 - 84. [7] AL - Aweni, M. O. (2013). simultaneous flow visualization and unsteady surface - pressure measurements in normally and obliquely impinging jets (Doctoral dissertation, MSU). [8] Krishna, S. (2012). - Investig - sensor measurements (Unpublished master's thesis). Delft University of Technology. [9] Hall, J. W., & Ewing, D. (2005). The development of the large - scale structures in rou nd impinging jets exiting long pipes at two Reynolds numbers. Experiments in fluids , 38 (1), 50 - 58. [10] Hall, J. W., & Ewing, D. (2006). On the dynamics of the large - scale structures in round impinging jets. Journal of Fluid Mechanics , 555 , 439 - 458. [11] E l - Anwar, M., K. Zhang, A. Naguib, and A. Abouel - Fotouh. "Pressure fluctuation produced by a jet impacting on a wall at normal and oblique incidence." ExHFT - 7 7th World Conference on Experimental Heat Transfer, Fluid Mechanics, and Thermodynamics, 28 June - 03 July 2009, Krakow, Poland: proceedings . Krakow: AGH U of Science and Technology Press, 2009. N. page. Print. [12] Harvey, J. K., & Perry, F. J. (1971). Flow field produced by trailing vortices in the vicinity of the ground. AIAA journal, 9(8), 1659 - 16 60. 144 [13] Naguib, A. M., & Koochesfahani, M. M. (2004). On wall - pressure sources associated with the unsteady separation in a vortex - ring/wall interaction. Physics of Fluids , 16 (7), 2613 - 2622. [14] ation for pressure in a turbulent flow. The Physics of Fluids , 24 (4), 777 - 777. [15] Blake, W. K. (1986). Mechanics of flow - induced sound and vibration. Volume 1 General concepts and elementary source. Volume 2 - Complex flow - structure interactions. Aplikace Matematiky, Applied Mathematics, 17. [16] Monnier, B., Naguib, A. M., & Koochesfahani, M. M. (2015). Influence of structural flexibility on the wake vortex pattern of airfoils undergoing harmonic pitch oscillation. Experiments in Fluids , 56 (4), 80. [17] Sa ffman, P. G. (1970). The velocity of viscous vortex rings. Studies in Applied Mathematics , 49 (4), 371 - 380. [18] Walker, J. D. A., Smith, C. R., Cerra, A. W., & Doligalski, T. L. (1987). The impact of a vortex ring on a wall. Journal of Fluid Mechanics , 181 , 99 - 140. [19] White, F. M. (2017). Fluid mechanics . New Delhi, India: McGraw - Hill Education. [20] Naguib, A. M., Jiang, W., & Zhang, K. (2012). Source advection effects on unsteady - surface - pressure asymmetry in oblique impinging jets. AIAA journal , 50 (7), 1634 - 1638. [21] Milne - Thomson, L. M. (1996). Theoretical hydrodynamics . Courier Corporation. [22] Koochesfahani, M., & Dimotakis, P. E. (1985). Laser - induced fluorescence measurements of mixed fluid concentrationin a liquid plane shear layer. AIAA journal , 23 (11), 1700 - 1707. [23] Crow, S. C., & Champagne, F. H. (1971). Orderly structure in jet turbulence. Journal of Fluid Mechanics , 48 (3), 547 - 591. [24] Kinsler, L. E., Frey, A. R., Coppens, A. B., & Sanders, J. V. (1999). Fundamentals of acoustics. Fundame ntals of Acoustics, 4th Edition, by Lawrence E. Kinsler, Austin R. Frey, Alan B. Coppens, James V. Sanders, pp. 560. ISBN 0 - 471 - 84789 - 5. Wiley - VCH, December 1999. , 560. [25] Liepmann, D., & Gharib, M. (1992). The role of streamwise vorticity in the near - fi eld entrainment of round jets. Journal of Fluid Mechanics , 245 , 643 - 668. 145 [26] Jung, D., Gamard, S., & George, W. K. (2004). Downstream evolution of the most energetic modes in a turbulent axisymmetric jet at high Reynolds number. Part 1. The near - field reg ion. Journal of Fluid Mechanics , 514 , 173 - 204. [27] Cooper, D., Jackson, D. C., Launder, B. E., & Liao, G. X. (1993). Impinging jet studies for turbulence model assessment I. Flow - field experiments. International Journal of Heat and Mass Transfer , 36 (10), 2675 - 2684. [28] Benedict, L. H., & Gould, R. D. (1996). Towards better uncertainty estimates for turbulence statistics. Experiments in fluids, 22(2), 129 - 136.