Download as pdf or txt
Download as pdf or txt
You are on page 1of 72

The Iby and Aladar Fleischman Faculty of Engineering The Zandman-Slaner School of Graduate Studies

Lagrangian measurements of Kolmogorov size neutrally buoyant particles in turbulent flow

A thesis submitted toward the degree of Master of Science in Mechanical Engineering by David Ratner

May 2010

The Iby and Aladar Fleischman Faculty of Engineering The Zandman-Slaner School of Graduate Studies

Lagrangian measurements of Kolmogorov size neutrally buoyant particles in turbulent flow


A thesis submitted toward the degree of Master of Science in Mechanical Engineering by David Ratner

This research was carried out in the School of Mechanical Engineering Under the supervision of Dr. Alexander Liberzon

May 2010

ACKNOWLEDGMENT

This work was carried out in the Turbulence Structure Laboratory at the Tel Aviv University. I am most thankful to my supervisor Dr. Alex Liberzon, for his support, help and guidance; without them my work would not have been possible. I also would like to thank all my colleagues from the laboratory, to Reut Elfassi for her guidance and help, to Mark Kreizer for his friendship and help, to Dikla Kersh, David Altura, Oleg Babin and Vitaly Haslavsky for their wonderful company, assistance and moral support. My special thanks goes t to all the technical and management staff in the Department of Fluid Mechanics and Heat Transfer that have made it possible for me to accomplish all the duties for the M.Sc. degree. They were all kind and helpful. Also, I want to thank Ira, my family and friends for their support and help; without them I wouldnt be here today.

ABSTRACT There has been increased effort to the develop measurement techniques for turbulent flows research, both in an attempt to achieve accurate small-scale measurements as well as to extract the maximum possible information from the processes observed. Small-scale measurements are especially important in two-phase flow in which the second phase consists of very small particles (in the range of a tenth of a millimeter). Despite extensive research, it is difficult to predict the velocity or distribution in space of the particles in turbulent flow, due to their interaction with the carrying fluid. Two-phase interaction bears significance in a range of applications including conceptual understanding of the physics of raindrops, aerosol sprays, and dust storms, among others. Measurement problems become far more complex when the flow is turbulent and the particles in question are on the smaller end of the flow length scales. The Maxey & Riley (1983) approach and similar, which assume that the particles are point-like, are no longer applicable to equation derivations for the motion of relatively large particles. We must ask to what extent such approach is applicable and whether it might help to adjust the definition and the physical significance of the Stokes number, St / .

The present work deals with the question: if neutrally buoyant particles with a diameter in the range of the Kolmogorov length scale behave differently than the flow, what is the main difference and which quantities underlie the physical mechanisms responsible for the observed differences? Particles movement was measured using a threedimensional PTV (particle tracking velocimetry) system that allows for identification and tracking of particles as they move through a control volume. The system can also identify the particles trajectory and kinematic properties, such as location, velocity and acceleration with great precision. Carrying fluid flow properties were measured using the same technique with neutrally buoyant flow tracers (with a mean diameter of 10 microns). The study was performed in a turbulent LDC (lid-driven cavity) flow at a Reynolds number of 10,000 based on the cubic cavity width (80 mm) and the velocity of the lid. We demonstrate differences between the various quantities we measured among which the most striking results came from the second order structurethe large particles do not behave like the flow tracers despite their low Stokes number and in fact show relatively higher correlation to the velocity-time increments.

ACKNOWLEDGMENT ABSTRACT...I TABLE OF CONTENTS..II LIST OF FIGURES..IV NOMENCLATURE...VIII 1 2 Introduction .............................................................................................................. 1 Literature review ...................................................................................................... 3 2.1 2.2 Turbulent particle-laden multiphase flows ........................................................ 3 Basic equations ........................................................................................... 7 Three-dimensional particle tracking velocimetry (3D-PTV) ..................... 9 Principles of 3D-PTV ........................................................................ 10 Tracking algorithm ............................................................................ 11 Overview of measurement techniques ............................................................... 8 2.2.1.1 2.2.1.2 3 4

2.1.1 2.2.1

Objectives ............................................................................................................... 12 Experimental method.............................................................................................. 12 4.1 4.2 4.3 Preliminary experiment.................................................................................... 12 3D-PTV experimental setup ............................................................................ 15 Calibration ................................................................................................ 16 3-D PTV program processing analysis ..................................................... 17 Matlab analysis ......................................................................................... 19 Methods of analysis ......................................................................................... 17

4.2.1 4.3.1 4.3.2 5

Results and Discussion ........................................................................................... 22 5.1 5.2 Preliminary experiment.................................................................................... 22 3D-PTV experiment ......................................................................................... 23 Eulerian Results ........................................................................................ 24 Average velocity maps ...................................................................... 24 Root-mean-square velocity contours................................................. 27 Reynolds stresses .............................................................................. 31 Summary of Eulerian analysis .......................................................... 31 Scales evaluation ............................................................................... 33 Distribution analysis.......................................................................... 34 Estimation of the settling velocity effect .......................................... 36 Spatial derivatives ............................................................................. 38 II 5.2.1.1 5.2.1.2 5.2.1.3 5.2.1.4 5.2.2 5.2.2.1 5.2.2.2 5.2.2.3 5.2.2.4

5.2.1

Lagrangian results .................................................................................... 32

5.2.2.5 5.2.2.6 5.2.2.7 5.2.2.8 5.2.2.9 5.2.2.10 5.2.2.11 6 7

Lagrangian correlation of velocity .................................................... 40 Autocorrelation of acceleration ......................................................... 42 Acceleration analysis ........................................................................ 44 Velocity increments .......................................................................... 46 Structure function .............................................................................. 47 Logarithmic derivatives .................................................................... 48 Mixed velocity-acceleration structure function................................. 50

Summary and conclusions ...................................................................................... 52 Bibliography ........................................................................................................... 54

APPENDIX A: Uncertainty analysis

III

LIST OF FIGURES Figure 1: Basic elements of a PTV system, components of the system are described in section 4.2 (Dracos 1996). .............................................................................................. 10 Figure 2 : Determination of the observation volume (Dracos 1996).............................. 11 Figure 3 Scheme of particle trajectory evaluation (Dracos 1996). ................................. 11 Figure 4 Schematic view of the preliminary experiment setup along with combined images from the experiment. x2-represents measuring volume, x1- represents distance between measuring volume and water surface, xn- xn-1 distance between two adjusted mark lines. ...................................................................................................................... 13 Figure 5 Electronic microscope images of two particles................................................ 14 Figure 6 3D-PTV experimental setup............................................................................. 15 Figure 7 Schematic view of cameras and cavity with axes origin and cavity location with respect to the origin. ............................................................................................... 16 Figure 8 Isometric view of calibration body of 0.060.060.006m, the body is 3D shaped and has channels of 1mm depth on each side with distance between white dots in x-y plane of 5 mm and of 1 mm in z plane. .................................................................... 17 Figure 9 Four calibration images (taken at the same time by 4 cameras), a-d) are cameras 1-4, respectively, according to the camera positions........................................ 17 Figure 10 Flow chart of the 3D-PTV software procedure (Willneff, 2003)................... 18 Figure 11 Example of an image (captured by camera 1) in tracers and particles experiment. ..................................................................................................................... 18 Figure 12 Lagrangian (right) to Eulerian (left) transformation of system of coordinates. ........................................................................................................................................ 19 Figure 13 Settling velocities of the particles as function of the depth. .......................... 23 Figure 14 Velocity vector maps. (a) z mid plane map of Ux , Uy , (b) y mid plane map Ux , Uz , (c) x mid plane map of Uz , Uy . Blue color represents tracer velocity and red color represents particle velocity, top arrow represent velocity scale. .................................... 25 Figure 15 Average velocity profiles comparison: (a)- Prasad and Koseff (1989) results ( SAR 1:1 Re=10000), tracers velocity profile using the Eulerian analysis, particles velocity profile using the Eulerian analysis. (b)- particles average profile for two different voxels sizes. ..................................................................................................... 26 Figure 16 ux r.m.s in z mid-plane (a) r.m.s velocity of tracers, (b) r.m.s velocity of particles, (c) r.m.s velocity of particles calculated using average velocity of tracers, (all axes are normalized by cavity width D). .................................................................. 28

IV

Figure 17 uy r.m.s in z mid-plane (a) r.m.s velocity of tracers, (b) r.m.s velocity of particles, (c) r.m.s velocity of particles calculated using average velocity of tracers (all axes are normalized by cavity width D). ........................................................................ 29 Figure 18 uz
r.m.s

in x mid-plane (a) - r.m.s velocity of tracers, (b) - r.m.s velocity of

particles, (c) - r.m.s velocity of particles calculated using average velocity of tracers (all axes are normalized by cavity width D). ........................................................................ 29 Figure 19 Profiles of r.m.s velocity for middle z plane of the cavity. (1) uy_r.m.s in cross-section of z and y middle planes, (2) - ux_r.m.s in cross-section of z and x middle planes. (a) - Prasad and Koseff (1989) results ( SAR 1:1 Re=10000), (b)- tracers velocity profile, (c)- particles velocity profile. ............................................................... 30 Figure 20 Reynolds stress contour maps at the z mid-plane (a)- tracers, (b)-particles. . 31 Figure 21 Particles trajectories in 3D cavity (a)- trajectories Longer than 100 frames, (b) - longest particle trajectories. Red circle represent the end point of trajectory. Cube denotes the limits of the cavity. ...................................................................................... 33 Figure 22 Particles/tracers histogram of spatial distribution, along (a)- x, (b)- y, (c)- z, directions, respectively. .................................................................................................. 35 Figure 23 PDF of spatial particle/tracer location. (a)- x location, (b)- y location, (c)- z location. .......................................................................................................................... 36 Figure 24 PDF of vertical velocity component Uy. (a)- tracers (average value is 0.012 m/s), (b) particles (average value is 0.007 m/s).............................................................. 37 Figure 25 Number of detected particles as function of frame number, with average number line (black) of 274 detected particles. ............................................................... 38 Figure 26 (Left) Plot of x component of particle velocity for several insulated trajectories in time, each blue line represent single trajectory. (Right) Velocity fluctuation v in respect to the linear fit representing average velocity. ........................ 39 Figure 27 Spatial derivatives. (1) Spatial derivative of full velocity, (a)- tracer (average value is 25.5 1/s2 ), (b)- particle (average value is 25.8 1/s2). (2) Spatial derivatives of velocity fluctuations. (a)- tracer (average value is 23.9 1/s2 ), (b)- particle (average value is 24.4 1/s2 ). ................................................................................................................... 40 Figure 28 Particle location and velocity at two different times on its trajectory............ 41 Figure 29 Lagrangian velocity autocorrelation coefficient. (a)- tracer (represented by solid lines curves) and particle (represented by symbols) autocorrelation coefficient of Lagrangian velocity vector and its components, (b)- tracer and particle autocorrelation coefficient of fluctuating velocity vector and its components. (full-Rvv , Rxx , Ryy , Rzz time is normalized by Kolmogorov time scale which is equal to 0.1s ) . .................... 42 V

Figure 30 Sketch of acceleration vector with its components of two points on particles trajectory. ........................................................................................................................ 43 Figure 31 Acceleration autocorrelation function. (a) Autocorrelation function of full acceleration vector, Raa (b) autocorrelation function of tangential to trajectory component of acceleration vector RaTaT. Time is normalized by Kolmogorov time scale which is equal to 0.1s. .................................................................................................. 44 Figure 32 Acceleration in direction of x axis (ax), tracers r.m.s <ax2>0.5=0.16 m/s2, particles r.m.s <ax2>0.5=0.17 m/s2 . ................................................................................ 45 Figure 33 Acceleration in direction of y axis (ay), tracers r.m.s <ay2>0.5=0.14 m/s2, particles r.m.s <ay2>0.5=0.17 m/s2 . ................................................................................ 45 Figure 34 Acceleration in direction of z axis (az), tracers r.m.s <az2>0.5=0.3 m/s2, particles r.m.s <az2>0.5=0.3 m/s2 . .................................................................................. 46 Figure 35 Velocity increments. (a) velocity increments of tracers, (b) velocity increments of particles. Time lags are 0.07 0.15 0.3 1.5 3.5 , and displayed with a vertical shift from 0.07 (up) to 3.5 (down) for clarity. ................................................... 47 Figure 36 Lagrangian second order structure function of particles and tracers x and y velocity components, lag is lag time from minor time interval between adjusted point to maximum interval between start and the end point of the trajectory, is Kolmogorov time scale of the flow. .................................................................................................... 48 Figure 37 Logarithmic derivatives of order 4. (a)-tracers, (b)-particles, lag is lag time and is Kolmogorov time scale of the flow. .................................................................. 49 Figure 38 Logarithmic derivatives of order 6. (a)-tracers, (b)-particles, lag is lag time and is Kolmogorov time scale of the flow. .................................................................. 49 Figure 39 Lagrangian time mixed structure function, lag is a lag time and is Kolmogorov time scale of the flow. ............................................................................... 50 Figure 40 Probability density function (F) of the cosine of the angle between v and a ................................................................................................................................... 52

VI

NOMENCLATURE a vi ax,ay,az particle/tracer acceleration projection on x y z axes respectively Lagrangian velocity i components. vn,vt are normal and tangential projection on , vx vy vz projection on the Cartesian axes.

the trajectory ui

Eulerian velocity components.ux - x component of velocity, uy y component of

velocity, uz - z component of velocity Ui Average velocity component Gravitational acceleration Density [ ]

f , p density of the fluid and particles accordingly PIV PTV LDA Re Kinematic viscosity Dynamic viscosity

Particle Image Velocimetry Particle Tracking Velocimetry Laser Dopler Anemometry Reynolds number =
V L

(V-velocity, L- length, -kinematic viscosity)

Rep , Rel

- Reynolds number which length scale is particle diameter and flow

representing length accordingly St 2D, 3D l, , Stokes number ( - particle response time, -fluid response time)

2 and 3 dimensional Integral , Taylor and Kolmogorov flow length scales accordingly, Kolmogorov time scale

fps dp Ub r

Mean rate of dissipation per unit mass frame per second Particle diameter Velocity of the leading band increment time function defined as f f t f t .

increment length function defined as f

f x

f x .

VII

1 Introduction
Development of experimental techniques capable of measuring the smallest scales is especially important in two-phase flows where the second phase of bubbles, particles and droplets typically affects the flow on small scales. Extensive research is needed because we do not have yet the correct prediction of particle position or velocity in complex and turbulent flows. As of late we have some statistical functions of velocity and concentration for some cases. We have no general theory for the turbulent flow itself. Although the equations of the flow are known (i.e., Navier Stokes), it is a nonlinear partial differential equation which has no general solution; this equation is one of seven millennium problems and a million dollar prize is currently being offered for its solution (Clay Mathematical Institute). In general, solutions are gained from neglecting the non-linear components of the equation, but this is not possible in turbulent cases. When dealing with turbulent two-phase flows, we have not only the unknown motion of the carrier fluid, but also the second phase with its own inertia, density, etc. Interaction between the two phases is a problem of great interest, since most of the natural and engineering flows include more than one phase. Such flows we see in aerosols, raindrops, dust particles in atmosphere, raindrops, plankton in the oceans and much more. The problem becomes much more complicated when the flow is turbulent where many factors and scales are involved, and particles are of the size of the flow scales or larger. It is noteworthy that, in order to derive the equation of motion of particles significantly larger than the smallest scale of the flow, it is not possible to use the approach of Maxey & Riley (1983), which is based on the assumption of point-like particles. It is therefore natural to ask to what extent an extension of the above approach is possible, and whether it might help to maintain the definition and the physical significance of the Stokes number. There are many questions we can ask about the interaction starting from governing equations and continuing with the drag coefficient where a lot of un-explained phenomena such as drastic reduction in drag coefficient of a rigid sphere at high Re numbers. Our way of understanding the phenomena is by measuring velocities, accelerations and locations of bodies in turbulent flow and trying to find some regularity and statistical order, so that we might say that in certain cases, flow with bodies will have some same predicted properties, with hope of predicting not only the same cases but also slightly different ones, trying to match formulas, nondimensional numbers like Re, St, Pa (Tanaka and Eaton, 2008). There are many measurement techniques of measuring flow and bodies in flow, from classical ones such 1

as anemometers and Pitot tubes, requiring insertion of a probe into the place interested to measure, to later and more complicated techniques that need not intrude the flow as, PIV, LDA and PTV (Dracos 1996), but insert into the fluid a second phase, follow after it with high speed cameras, analyze it's motion and conclude on the flow motion. With development of advanced techniques, better cameras and strong computers to process the data, we can now get to smallest scales with miniature particles that follow the smallest eddies thereby measuring and exploring phenomena that we could only dream of measuring before. Therefore, the development of these methods, which allow for the measurement of parameters that could only be estimated in the past, can be extended further for the exploration of new concepts and ideas.

2 Literature review
The literature review focuses on the studies of turbulent multiphase flows of fluids and solid particles, which are in the focus of the present study. Recent experimental, numerical and theoretical studies are reviewed in the following section.

2.1 Turbulent particle-laden multiphase flows


Single-phase turbulent flows of Newtonian fluids are very complicated and require extensive experimental, theoretical and numerical research in order to obtain relatively simple empirical models. It is even more difficult when the flows include a second phase, e.g., particles. The Brennen (2005) textbook, for example, mentions two main groups of complications related to the particle-laden turbulent flows: 1. Complex unsteady motions of the particles that may result in non-uniform spatial distribution of the particles and possibly even particle segregation. It can also result in particle agglomeration or in particle fission, especially if the particles are bubbles or droplets. 2. Modifications of the turbulence itself caused by the presence and motions of the particles. Turbulence could be reduced by the presence of particles, or it could be enhanced by the wakes and other flow disturbances that the motion of the particles may introduce. The question that remains asks under which conditions each of these scenarios might occur and what are the important parameters that dominate the various phenomena. The effect of particles on the flow was studied by direct numerical simulations (DNS) as well as through a variety of experimental methods. Since the particles are Lagrangian objects, Lagrangian properties such as dispersion of particles, their velocity and acceleration history become of primary interest. The first numerical studies in isotropic turbulence were performed by Yeung and Pope (1989) for various Reynolds numbers. The authors found a strong correlation of the acceleration statistics with the Taylor micro-scale Reynolds number, . The connection between Lagrangian velocity

autocorrelation and Eulerian velocity autocorrelation was examined experimentally by Shlien and Corrsin (1974). They found that the major difference in Taylor micro-scales calculated from autocorrelations resulted from the different definitions between Lagrangian and Eulerian autocorrelations, where Eulerian autocorrelation relies on

constant distances in pre defined control volume, and Lagrangian autocorrelation relies on distances particle makes on its way (trajectory). Particle-liquid interaction is the key to understanding major multiphase flows. It includes different issues, starting from fundamental aspects of modeling dispersion for numerical studies Shirolkar et al. (1996), which reviewed and summarized leading works in order to describe the problem, to experimental research of turbulent transport of particles with different diameters in isotropic turbulent flows; Qureshi et al. (2007) who found that there is no shape change in acceleration probability density function in contrast with its variance and chaotic flows; Ouellette et al. 2008, who studied millimeter-scale particles compared to finite sized particles and found a systematical difference between their statistical properties. The effect of particles on the flow was researched also by Gore and Crowe (1989) who studied the effect of particles on turbulent intensity and claimed that the value of the ratio d (particle diameter to Kolmogorov length scale) can indicate whenever

turbulent intensity is increased or decreased by the particles: where this ratio is larger than 0.1, turbulent intensity of the carrier fluid (first phase) is increased, and smaller than 0.1 is decreased. Measures have been carried out in an attempt to understand the problems theoretically. Tanaka and Eaton (2008) added new parameters to try to classify turbulence modification by sphere particles with the novel dimensionless particle moment number Pa , which was created in order to simplify the carrier-phase Navier

Stokes equations in the presence of particles, along with Re number. They found a turbulence attenuation region between augmentations for two critical particle momentum numbers. The problem of assessing large and heavy particle distribution and clustering regions in turbulent flow has been examined by Guala et al. (2008) who found clustering of these particles in strain-dominating regions. The continuation of this work with heavier-thanfluid Kolmogorov sized particles at fully developed turbulent flows was performed by Gibert et al. (2010) who found that the mixed Eulerian structure function of two particle velocity and the acceleration difference vector v a increase significantly with alignment between these dynamical quantities.

particle inertia, and showed that this increase can be attributed to a preferential

One of the major characteristic dimensionless parameter describing particles in flow is the Stokes number, St, which has many definitions and significant impact on its motion. The Stokes number is defined as a ratio of particle time scale p to a representative time scale of the flow f , or as viscous drag term in the equation of motion (Maxey and Riley, 1983). The time scale of the particle reflects how long it takes the particle to react to flow conditions. Its value is as follows: For small particles which Reynolds number based on settling velocity (Vs) is very small O

O 1 , time scale estimation is:

(1)

When particle Reynolds number based on settling velocity is of order of one O

O 1 , we include Oseen correction to particle time scale:

(2)

According to Fallon and Rogers (2002), the definition of the fluid time scale is not always apparent and depends on limiting factors in the flow and there are many ways of choosing time scales based on different definitions. The same problem exists in definition of Reynolds number, and there are many definitions of Stokes number based on many fluid time scales, usually defined with same values as Reynolds number with characteristic velocity and length scale. Ouellette et al. (2008) defined Reynolds number as Re=ur.m.s L/ where ur.m.s is r.m.s velocity and L is magnet spacing (the flow was induced by a Lorentz force who set into motion KCL solution), and Stokes number defined as Re .

Guala et al. (2008) defined particle time scale as: location) and as Kolmogorov time scale of the flow. (3)

(Neglecting the Basset and Faxn forces as well as acceleration of the fluid in particle

between 1 as:

Gibert et al. (2010) defined Stokes number for solid particles with densities varying / 8, and particle diameters of the order of Kolmogorov length scale St (4)

where

3 / 2

is the modified density ratio that takes into account the added

mass effect, and is the Kolmogorov length scale. Bourgoin et al. (2009) suggested that Stokes number effects (defining Stokes number as p/D where the particle time scale p has been corrected in agreement with Schiller and Nauman (1933) and D is the flow eddy turnover time at a scale of particle diameter) cannot be considered alone to properly characterize and model inertial particles. Optical experiments can supply the direct experimental data of particle velocity location and acceleration properties. For example, optical particle tracking methods were used by Sato and Yamamoto (1987) who performed Lagrangian measurements of fluidparticle motion and determined r.m.s. values of fluctuating velocity and estimated velocity autocorrelations. More complicated properties such as velocity structure functions and logarithmic derivatives, which are derivatives of basic properties (location, velocity and acceleration), are often used to shed light on subjects like physics of dispersion and mixing, (e.g. Toschi and Bodenschatz 2009). One of the parameters requiring greater attention is acceleration and its derivatives. The statistical parameters, such as autocorrelation and mixed structure functions, are typically used. These provide us clues about the forces particles experience while moving in flow. For example Calzavarini et al. (2009) demonstrated that point particle models do not capture the effects of particle finiteness on their statistical properties if the Faxn forces are not taken into account. The impact of parameters, such as particle density and diameter on these forces was studied by Bourgoin et al. (2009), who demonstrated that Stokes number alone is not enough to properly characterize and model inertial particles, and by Brown et al. (2009), who, in contrary, found little dependence between particle size and the scaled acceleration probability density function.

2.1.1 Basic equations


The majority of the research reviewed above describes particle motion according to the well-known equations of a small rigid sphere proposed by Maxey and Riley (1983). The equations are written for a small rigid sphere of radius , which is present in the fluid

flow and which, in absence of the particle, is otherwise defined by the velocity field ,

. The presence of the sphere (or particle) and its motion through the fluid will .

modify the flow locally and lead to a new velocity field, for incompressible flow of uniform density:

The authors (Maxey and Riley, 1983) applied the momentum and continuity equations

(5)

x, t u x, t as |x Y t |

with the following boundary conditions surface and

on the sphere

If the fluid stress tensor is defined as:

(6) then the equation of motion for the spherical particle is: m m g
.

n dS

(7)

where surface integral is over the surface of the sphere and

is the normal to the

surface direction. The problem is: in order to evaluate the fluid stress tensor on the sphere one can transform the coordinates to a frame moving with the center of the sphere with given transformation equations: , , , ,

. Also, in order to take advantage of small parameters, it is further

convenient to separate the flow field into two parts, q(0) and q(1) ; q(0) being the undisturbed flow without application of the boundary condition and q(1) is the disturbance made by the sphere , incompressibility. The equation for the disturbance flow reduces to 7 (9) the two components of the

flow that together satisfy the equation of motion, and each separately satisfies

provided that

1 Reynolds number defined on sphere radius and velocity / U /L 1 where L is representative length scale of

of undisturbed flow) and

the undisturbed flow, and U /L is the scale for the corresponding velocity gradient.

Besides the disturbed flow, there is also a contribution to the fluid force from the undisturbed flow q , which may be found quite generally and without specific

assumptions about low Reynolds number, such as the conditions used to derive the is F

previous equation. From the above, the contribution to the fluid force on the particle from q ; F F n p a (10)

which may be converted to a volume integral and approximately evaluated as (11)

under the assumption that the terms in parentheses are nearly uniform over the sphere, provided that the sphere is sufficiently small. This assumption is valid if the size of the sphere is small compared to the length scale of variations in the undisturbed flow (Maxey and Riley, 1983). This is the key assumption, which does not exist in turbulent flows with particles of the order of scales of the undisturbed flow (particles of this size are the key interest in this thesis). The smallest scales of the turbulent flows are Kolmogorov length scales, whose dimensions are of the order of the particles diameter and even for tracers the difference is not very significant (around 1:50-100), but sufficient for the Maxey and Riley (1983) equations. For flows with large Reynolds number and not point like particles, flow around the sphere is not uniform over the sphere. That is the reason why we cannot use analytical approach predicting the motion of the particles.

2.2 Overview of measurement techniques


There are many techniques that can be utilized for the measurement of flow/particle velocity. Flow field measurements are mainly based on non-intrusive optical techniques, among which are particle image velocimetry (PIV), which is designed to determine the movement of small groups of particles embedded in a fluid element during a short time as well as its stereo, multi-plane and holographic applications that allow to measure velocity vectors in three-dimensional space. Particle tracking velocimetry (PTV) allows for determination of the paths of individual particles inside

an illuminated volume. Laser Doppler anemometry (LDA, or LDV), allows for the measurement all three components of the velocity at one location. Another technique is acoustic Doppler in which measurements are limited to particles approximately five times larger than the Kolmogorov scale. All aforementioned techniques have their advantages and disadvantages which are reduced with technology advance, (i.e. faster cameras with higher resolution, smaller sizes of the probes, higher illumination, and faster and bigger storage devices).

2.2.1 Three-dimensional particle tracking velocimetry (3D-PTV)


Three-dimensional particle tracking velocimetry (3D-PTV) determines the velocity and acceleration from the displacement of individual particle in moving fluid during a prescribed time interval. 3D-PTV method has its limitations: the relatively low number of tracked particles per unit volume (as compared to the PIV, for example), illumination must be pulsed at high rate or continuous so that cameras can follow the particle in the flow and the optical access requirement adds constraints to the experimental setup. Moreover experiment duration time is also limited by the capacity of storage devises and their recording speed. Snyder and Lumley (1971) provided the first systematic set of particle tracking velocity measurements from wind-tunnel grid turbulence, and later Sato & Yamamoto (1987) reported similar measurements in water-tunnel grid turbulence. Three-dimensional Lagrangian particle tracking using multiple cameras for the stereoscopic reconstruction of particle tracks was pioneered by Virant and Dracos (1997) and further developed by Ott and Mann (2000). The first measurements at high Reynolds numbers, using direct optical imaging were accomplished by means of silicon-strip-detector technology (Mordant et al. 2004a, b; La Porta et al. 2001; Voth et al. 2001, 2002); however, only single-particle dynamics could be investigated. Today the technology of complementary metal-oxide semiconductor (CMOS) cameras allows the tracking of many particles at rates equivalent to those of the silicon strip detectors, but with only one quarter of the spatial resolution (Bourgoin et al. 2006). Recently, improved three dimensional particle tracing velocimetry techniques allowed access to velocity gradients along particle trajectories (Lthi et al. 2005). The same technique has been extended to scanning particle-tracing velocimetry to reach even larger particle densities (Hoyer et al. 2005).

A recent development involves instrumented particles which are promising tools in exploring Lagrangian turbulence (Gasteuil et al. 2007). These particles can include sensors capable of measuring local quantities such as temperature or acceleration and transmitting the signal in real time to receiving stations (current large size of the particles limit their utility for turbulent flows) (Toschi and Bodenschatz, 2009).

2.2.1.1 Principles of 3D-PTV 3D-PTV tracks the trajectories of individual particles in three-dimensional space. Particle identification and determination of its position in space is a very important step in PTV, so in order to achieve the needed accuracy: therefore, the seeding density of particles is limited. In order to measure flow velocity, the use of very small and neutrally buoyant particles (called tracers), which claim to be a perfect trace of the flow and react with its minor changes is necessary. Using them we can say that their velocity and trajectories are velocity and streamlines of the flow.

Figure 1: Basic elements of a PTV system, components of the system are described in section 4.2 (Dracos 1996).

Detection of these tracers and bigger particles in the observation volume depends on the phenomena studied and the equipment used (such as the power of a laser or another light source, focus range, and quality of CCD cameras), so the images produced are suitable for later analysis. Examples of an array of this type are shown in Figure 1 and the observation volume with its dimensions and important parameters in Figure 2.

10

Figure 2 : Determination of the observation volume (Dracos 1996).

After the tracers/particles are identified and located in the pre-set coordinate system the second stage analysis role is to find a match between tracers/particles in time and thus to create trajectories that can later be used to determine velocity, acceleration and other parameters. 2.2.1.2 Tracking algorithm

Figure 3 Scheme of particle trajectory evaluation (Dracos 1996).

Tracking algorithms requires assuming the maximum velocity of the particle which determines the searching sphere in which the particle might be found in the next time step. Knowledge on the previous particle location and velocity provides the algorithm with possible location where particle may be in the next (and if this analysis goes back in time then the previous) frame is received. Figure 3 demonstrates a three-frame algorithm basic, including particle searching spheres and five possible variants of trajectory path. From frame f to frame f+2, the particles are located inside the searching sphere and each one could represent the particle for which the trajectory is being built. The best trajectory is chosen by its smoothness and minor possible changes in velocity and acceleration. Typically the trajectory is smoothed using a polynomial or another low-pass filter from which the tracer location, velocity and acceleration are calculated. 11

3 Objectives
The primary objective of this study is to provide a set of properties that will unequivocally define whether the solid particle motion is different from the carrying fluid motion. It is of our interest to compare the Eulerian and Lagrangian approaches to the 3D data obtained using 3D-PTV experimental method. Specifically, in this work we intend to established 3D-PTV to acquire the motion of the solid particles in a turbulent lid-driven cavity flow, attempting to measure turbulent (fluctuating) velocity of the solid particles and compare it to that of the flow tracers. We use Eulerian ensemble averaging and directly measured Lagrangian properties to demonstrate main similarities/differences of the finite size particles and flow tracers.

4 Experimental method
In this section there is a description of the experimental arrangement, and a brief explanation about the ways of how the experiment was done. The first set of experiments is designed to measure the settling velocity of the particles under investigation in order to obtain precise definition of the Stokes number and verify the values of the particle diameter and density. Following the precise definition of the particles used in the following study, the PTV experiments, which include tracers and particles experiments, were designed and performed as described below.

4.1 Preliminary experiment


The settling velocity experiment, which is represented in Figure 4 along with direct diameter measurements represented in Table 1, provides us with sufficient information about particles we used later in 3D-PTV experiments. Figure 4 represents a schematic view of the experimental setup, showing a glass beaker aligned with a camera and a combination of two images, where the particle is shown at the beginning and toward the end of measuring volume (103 frames between those two images).

12

Figure 4 Schematic view of the preliminary experiment setup along with combined images from the experiment. x2-represents measuring volume, x1- represents distance between measuring volume and water surface, xn- xn-1 distance between two adjusted mark lines.

Experimental procedure: Ten particles were released one by one into a tall beaker that has marks at constant and known distances of 2.7 0.1 mm, (small unnumbered marks that are too numerous to be counted in Figure 4). Particle release was made in the center of the vessel (this was done so influence of the walls would be minimal), and the settling was measured by a high-speed camera at a speed of 15 frames per second. We used a high-speed camera because its internal electronic shutter is very precise and we know that the time delay between the frames is 0.06667 seconds with a negligible error, therefore we know the exact time that it takes for a particle to fall when thrown a known distance. By filming the falling of the particle in the beaker and knowing the time passed between each frame where we see the particle passing the marks, we can manually extract (using Matlab, detecting manually particle location and calculating distances the particle made between frames) particle velocity according to the following equation: 12 Where is the distance between marking lines, - settling velocity. is the number of frames, frames-per-second (fps) rate and

The particles fall almost vertically, for example, in Figure 4, the upper particle x position and y position is 442 and 78 pixels respectively, x and y position of the same particle later on (shown as lower particle in Figure 5 ) is 440 and 871 pixels, indicating that it moved 2 pixels in horizontal position while moving 793 pixels in vertical position. 13

The observation volume is 5 cm high and located 20 cm beneath the water level. At this depth particles already reach constant velocity according to Stokes law. Using the settling velocity results, we can estimate particle density with high accuracy. If the particles are falling in the viscous fluid by their own weight, then a terminal velocity, also known as the settling velocity, is reached when the frictional force combined with the buoyant force exactly balances the gravitational force. The resulting settling velocity is given by: (13) where dp is the radius of the particle, fluid, respectively and gravity. Figure 5 presents electronic microscope images of two particles with measured diameters. It shows the rigid spherical shape of the particles. , - is the density of the particle and of the

is dynamic viscosity and g is the acceleration due to the

Figure 5 Electronic microscope images of two particles.

In the following Table 1 we summarize the results of the measurements of particle diameters. Ten randomly selected particles were measured using analog caliper (measurement error of 0.005 mm) and the results are given in millimeters.
Table 1 Diameter of 10 random particles used in preliminary experiment and the mean diameter

Particle

10

0.73 0.68 0.74 0.72 0.79 0.66 0.64 0.72 0.67 0.8

14

4.2 3D-PTV experimental setup


The main experimental method used in this work is 3D-PTV, providing particle and tracer location, velocity and acceleration, along with statistical information needed to fulfill the research objectives. Figure 6 shows a schematic view of the experimental arrangement, with the main components listed.

Figure 6 3D-PTV experimental setup

We use a glass cubic cavity of 80 x 80 x 85 mm. The cavity is located inside a large glass tank and positioned using a plastic frame such that its top is parallel to the top edge of the large glass tank. The main reason of using the small cavity in the larger tank is to have the lid submersed in water, preventing air entrainment and open surface disturbances. A flexible smooth plastic belt (total length of 103 cm, 2 mm thick) serves as a lid in our setup. The belt is driven by a servomotor (Dahaner Motion, Santa Barbara, CA USA) and the mechanical system designed by IT&ES Engineering (Haifa, IL) The control is done using Kollmorgen SepLink 2.0.7 software and a SERVOSTAR controller. The controller adjusts the angular velocity of the driving motor with the precision of 1 [rev/min]. The speed of the belt is measured manually using the time lapse for the 10 full cycles of the belt. The experiments shown in this work were performed using the belt speed 0.125 [m/s]. Laser light provided by the dual pulsed laser (527 nm, Quantronix, USA) was pumped at 1000 Hz, providing approximately 75 watt of light, illuminating the small cavity. We used four digital CMOS cameras (1280 x 1024 pixels, 8 bit, Mikrotron, Germany). The cameras are located as shown in Figure 7, while two tapirs of cameras were facing the cavity from two sides, pointing toward the center of it. The positions are given in

15

coordinate system attached to the calibration target as we explain below. The cameras simultaneously captured and transferred the video using a real-time image streaming and storage system. The four cameras are synchronized with a maximum possible time jitter of 1/1000 frame rate. The digital video recording system includes 48 hard drives and four special frame grabbers (CLFC, IO Industries), each allowing for a 700 Mb/sec data transfer rate. In addition to the large particles described above, the flow was seeded by the flow tracers that are hollow-glass spheres with the density of 1.03 g/cm3 and an average diameter of 11 microns.

Figure 7 Schematic view of cameras and cavity with axes origin and cavity location with respect to the origin.

4.2.1 Calibration
The calibration procedure included in the 3D-PTV software provides configuration of the cameras, consisting of 16 parameters such as focal length, X, Y, Z position of the camera imaging center, angles of the camera imaging axis in respect to the coordinate system in the cavity frame of reference, magnification factor and optical distortions. A detailed description is given in Dracos (1996). The results are the calibration files that allow direct and inverse projection of every pixel into the 3D space and every point in 3D space onto the 4 imaging planes of the cameras. The calibration body used in calibration is made of a non reflective black metal alloy with white light-reflective dots see Figure 8, central dot is defined as axes origin of the cavity and the whole experimental coordinate system.

16

Figure 8 Isometric view of calibration body of 0.060.060.006m, the body is 3D shaped and has channels of 1mm depth on each side with distance between white dots in x-y plane of 5 mm and of 1 mm in z plane.

The calibration procedure included the following steps: The calibration body is securely placed inside the cavity corresponding to the axis shown in (Figure 7). Pictures of the calibration body are taken in the cavity by all four cameras (Shown in Figure 9). Measurement of the approximate camera location and viewing angles with respect to the calibration body axes.

a)

b)

c)

d)

Figure 9 Four calibration images (taken at the same time by 4 cameras), a-d) are cameras 1-4, respectively, according to the camera positions.

4.3 Methods of analysis


We used 3D-PTV software developed by ETH Zurich (Swiss Federal Institute of Technology) and available under a free academic license from http://ptv.origo.ethz.ch. The output is a set of trajectories organized in ASCII files. The database is processed then using Matlab software that creates the necessary Lagrangian and Eulerian statistics.

4.3.1 3-D PTV program processing analysis

17

The software analyzes the images according to the following Figure 10, which schematically represents the 3D-PTV algorithm.

Figure 10 Flow chart of the 3D-PTV software procedure (Willneff, 2003).

Figure 11 presents an example of an image captured by one of the cameras. The image shows the laser illuminated cavity with tracers and particles (marked in red). The particles are clearly large compared to tracers. The software allows discriminating the particles by size and track separately the large particles and the flow tracers.

Figure 11 Example of an image (captured by camera 1) in tracers and particles experiment.

18

In this experiment we analyzed 1000 frames of tracers and 1617 frames of particles, captured at rate of 150 frame per second. The output of the software is positions of the particles in the cavity at every time instant, linked together into trajectories. Particle positions in time are used to estimate velocity and acceleration along the trajectory. During the post-processing stage, the particle positions are smoothed to fit into curve performing spline functions for parts of the trajectory (detailed explanation on the process and its errors are given in Lthi et al., 2005).

4.3.2 Matlab analysis


Data obtained in the experiment is in Lagrangian system of coordinates (meaning that we were following an individual fluid parcel as it moves through space and time). In Lagrangian sections, short trajectories are usually negligible (less than 10 frame trajectories) and long trajectories are the ones that provide us with useful data for a relatively long period of time and different locations in the cavity. In Eulerian section (at fixed locations) we use the entire data processed throw the analysis described below without filtration. The reason for this is that most of the data are in short trajectories (trajectories that have more than 10 frames are less than 5 percent of all the trajectories) and the accuracy of statistical processing that describe the Eulerian analysis depends directly on the amount of data. Figure 12 presents a sketch of Lagrangian to Eulerian coordinate system transformation (explained below), showing briefly how particles velocities are transformed to fixed system locations velocity.

Figure 12 Lagrangian (right) to Eulerian (left) transformation of system of coordinates.

Lagrangian analysis Lagrangian analysis is carried out using the information about the path of particles. It contains information about evolution of particles in flow. The analysis is pointed toward 19

particle information along its path, and is a direct continuation of 3-D PTV program processing, so that no further processing is necessary. Using Matlab functions and self made codes (like autocorrelation and structure functions) results are achieved in the form presented in the results section. Lagrangian analysis provides information about particle trajectories, locations, velocity and acceleration autocorrelations, velocity structure functions etc. There are flow parameters that cannot be evaluated using Lagrangian analysis, including mean and fluctuating velocity analysis, spatial distributions of Reynolds stresses. Therefore the Lagrangian data has to be interpolated onto a fixed grid to proceed with the Eulerian analysis. Eulerian analysis The Eulerian results we obtain by the interpolation of Lagrangian (sparse and unevenly distributed in space) to Eulerian (uniformly distributed in space) data. The procedure is performed in Matlab, using linear interpolation in three-dimensional space. The cavity is divided into voxels (3D spaces) of 555 mm with special attention devoted to the cavity boundaries. In some cases the reflected light masked the data in near wall regions. Thus, for example, at the top part we had to remove from the analysis the top 5 mm. Therefore we do not have a direct measure of the lid velocity by the 3D-PTV. Due to sparse distribution of the particles, there are plenty of empty voxels (usually near the boundaries of the cavity) at different time steps. This causes an undesired effect of regions with minor statistics and in result we have windows where the resultant data is not reliable, for 5202 (181717) voxels we usually have 274 tracers and a quarter of that amount of particles, therefore although we use the data from these regions in our Lagrangian analysis part we cannot do that in the Eulerian part. Error of the interpolation procedure: Eulerian analysis uses velocity components in order to calculate average and fluctuation velocity as well as Reynolds stresses, which are derived from fluctuating velocity components. Therefore it is important to perform error analysis of the interpolation procedure, in which the main component of uncertainty is velocity due to an error related to the position of the interpolated point in respect to the sparse data points. The first issue is related to the number of voxels (3D boxes the center of which we are the Eulerian grid nodes). In order to evaluate the average analysis error we must estimate the amount of particles and tracers been in voxel and participated in the 20

averaging procedure. The average number of flow tracers per voxel is evaluated using location analysis presented in results section (5.2.2.2), from which we get an average of 2000 tracers and 1000 particles per 18 voxels, which gives us around 111 tracers and 55 particles per voxel; average velocity is generated from this amount of tracers/particles velocity data. Instant velocity maps are used for averaging and fluctuation analysis. In assembly average analysis, 3D-PTV error (calculated in appendix) is divided by square root of average number of tracers/particles that appeared in voxel as a result of average operation. In order to calculate root mean-square-maps, fluctuation maps are calculated by subtraction of average velocity map from instantaneous velocity maps. The uncertainty of the process is the same value as 3D-PTV error. A root-mean-square map is made by composing averages of squared fluctuation maps and taking square root of the resulting map. The uncertainty brought by this process is twice as big as the averaging analysis error. This leads us to the second issuethe evaluation of the error in instant velocity voxel, for which we established the aforementioned procedures. The second issue is that in transformation processes there is a significant chance in high visited areas of the cavity (middle areas) as shown above that there are several tracers/particles in one voxel. In this case the linear interpolation is made between velocity components of all the particles/tracers that are in the voxel at the same time. The error of mentioned interpolation can be estimated according to the linear interpolation error. This is the algorithm used in the Matlab subroutine (griddata). Linear interpolation is often used to approximate a value of some function f using two known values of that function at other points. The error of this approximation is defined as R(t)=f(x)-p(x) where p denotes the linear interpolation polynomial defined below. It can be proven using Rolle's theorem that if f has a continuous second derivative, the error is bounded by| |

The approximation between two points on a given function becomes even less accurate with the second derivative of the function that is approximated. That means that in our case the approximation in high strain regions like near downstream wall is less accurate then in low strain regions like center of the cavity. We analyzed squared spatial derivatives (5.2.2.4) and assumed that they are of the same size with second spatial velocity derivative. Since the typical values of the derivatives

21

are about the same for tracers and particles and equal to 5 [1/s2] we take this value as a representative value in error analysis. The maximum distance between particle and center of the voxel to which it is interpolated is half the size of the voxel and equal to 0.0025m. This provides us with an error of 0.410-5 [m2/s2]. Therefore for typical velocity in voxel of 1 cm/s, we have an error of square of the error calculated above of 1.1 mm/s. The resulting error in calculations of instant velocity in a single voxel is 10% of the measured velocity, both for tracers and for particles. The error of the averaging procedure of instant velocity maps (as described above) for tracers is equal to (
% %

=1%) for tracers and (

=1.35% ) for particles.

The error of root mean square analysis is twice the error of the averaging procedure (since the average operation in rms procedure is done over squared numbers the overcool error growth is twice larger then in ensemble average operation), and in it the particles error stands at 2.7% with tracers error of 2% of the rms velocity values. The error analysis presented above is an evaluation for most of the cavity regions, whereas the error in high strain regions (near the corners of the cavity and downstream wall) is much higher than the evaluated error.

5 Results and Discussion


5.1 Preliminary experiment
In this section, results of the preliminary experiment which setup is explained in the experimental method section (4.1) are presented. Analysis performed to the time data gathered in the experiment, data measured from 10 particles, is presented as a distance between marked locations of the particle (see Figure 4 for the sketch of the experimental setup). The result is the settling velocity shown in Figure 13. This figure presents

calculated velocities of 10 particles as function of the distance from the first volume mark line in the control volume (as can be seen as a function of settling time). As we can see from this figure, the particles have a constant settling velocity, with variation of about 0.001 m/s, due to the uncertainty of the experiment. The main sources of the uncertainty are the error in number of frames (maximum error is 1 frame); other error sources such as fps error (1/1000 of frame rate) and distance measuring errors (0.01mm) provide negligible errors that do not affect total velocity error.

22

Figure 13 Settling velocities of the particles as function of the depth.

The average settling velocity is 6.37

0.95 mm/s.

The uncertainty in velocity

represents the possible deviation of particle settling velocity values from average settling velocity value (shown as error bars in figure above); the difference in particle diameters (15% uncertainty, section 4.1) are the main reason for that deviation. Dependence between settling velocity and density of the settling particle is according to this equation: V

from which we estimate the particle density, density is 1.024 0.004

0.024

. Thus our particles

. The density difference is small, therefore the Stokes

number (settling) is also small and cannot explain the differences we observe in large particle behavior in the results of particle tracking experiments, given in the following section 5.2.

5.2 3D-PTV experiment


This section presents results of the 3D-PTV experiment data analysis done using 3DPTV techniques described above. The experiments were carried out according to the experimental method shown in section 4.2. Using 3D-PTV techniques two experimental sets were performed: The first one used flow tracers, and the second one uses large solid particles. 23

The second set contains 2 experimental runs of about ten seconds each. These experimental runs were performed for ten minutes, one after another in order to verify that our ensemble is taken in a steady-state flow. Both experiments were performed in the cavity under steady conditions of a constant lid speed (U particles was 150 fps. The Reynolds number, Re 0.125 . The camera frame rates in both sequences of tracers and 10 10000, (where L is the side ) in both

length of the cavity at 0.08 m and is the fluid viscosity, 1

experimental sets. The root-mean-square velocity of the velocity fluctuations is estimated as 0.01 m/s (the distribution is shown below) and the Taylor micro-scale is estimated as 3.8 mm. Reynolds based Taylor micro-scale is Re 38.

5.2.1 Eulerian Results


The results presented in this section should be interpreted with care, being interpolated from the Lagrangian results as described in section 4.3.2. The Eulerian results are important to understand the overall flow pattern and to enable estimation of fluctuations in respect to the average flow. The Eulerian voxel maps of instantaneous velocity are further processed in Matlab in order to get the results we present below. We present the mean velocity using the cross-sections through the cavity mid-planes from orthogonal views, distributions of fluctuating velocity and the Reynolds stress contours. 5.2.1.1 Average velocity maps The average maps presented below are a product of the ensemble averaging operation whose procedure and errors are described in section 4.3.2, according to formula ( is , , component of instant velocity). The maps present cavity mid-

planes from orthogonal views showing each time two average velocity components. Figure 14 (a) - shows Ux ,Uy (Capital letters denote average) vector field (the scale of arrows is shown in the figure) in z mid-plane, Figure 14 (b) - shows Ux, Uz vector field in the y mid plane and Figure 14 (c) - shows Uz, Uy vector field in the x mid plane. Blue color arrows are tracer average velocity components and red arrows are particle average velocity.

24

(a)

(b)

(c)
Figure 14 Velocity vector maps. (a) z mid plane map of Ux , Uy , (b) y mid plane map Ux , Uz , (c) x mid plane map of Uz , Uy . Blue color represents tracer velocity and red color represents particle velocity, top arrow represent velocity scale.

As one can see in Figure 14, the main difference between average velocity of particles and tracers are in the high shear (near the lid and cavity walls) regions. In the low shear regions (center of the cavity) the difference is insignificant. The high shear regions are also regions with fewer particles (as shown below) and therefore the uncertainty is higher. Also be seen that the main vortex is not exactly in the middle of the cavity but is located slightly in downstream direction (positive x) of it; this is the reason why we dont see zero Uy values in Figure 14. The normalized axes presented in Figure 15 are from -0.5 to 0.5 of the whole cavity length, but for reasons weve discussed above, we did not present data on the regions near the walls.

25

In the following, average velocity field in Eulerian setting is compared to the results available in the literature of average profiles in lid driven cavity flows with the same geometric and flow conditions. We compare the mean velocity profile at lines of x= -0.01m and y=0 m for z middle plane to the velocity profiles presented by PK89 (Prasad and Koseff 1989). The comparison has to be taken with care since the precise horizontal/vertical location of the profile by PK89 is not known. In Figure 15 (a) we present the mean velocity distributions (velocity profiles) obtained from the Eulerian results and compare it to the results of Prasad and Koseff (1989), using the data from flow tracers and particles, and in Figure 16 (b) we made an Eulerian analysis for the whole cavity height with two different voxel sizes the common used in the rest of the analysis 555 mm voxels and a smaller size of 222 mm voxels, this analysis is done only to particles x axes velocity in order to see if there is any significant change in average velocity values between two voxel sizes especially in the region where our results are different from PK89 data (bottom of the cavity). There are two pairs of axes in the Figure 16 (a); the inner scale represents the distance normalized by using the lid velocity, U . the cavity side length D and the outer scale represents average velocity normalized

(a)

(b)

Figure 15 Average velocity profiles comparison: (a)- Prasad and Koseff (1989) results ( SAR 1:1 Re=10000), tracers velocity profile using the Eulerian analysis, particles velocity profile using the Eulerian analysis. (b)- particles average profile for two different voxels sizes.

From Figure 15 (a) we see that the average velocity profiles are similar for particles and tracers, though our results are higher than the results of PK89. It is worthy to mention that in PK89 the two velocity profiles meet in the middle of the cavity presenting the center of the main vortex. Our expected results do not support this finding. In Figure 14 26

(a) you can see that the center of the vortex is not located in the middle of the cavity. From Figure 16 (b) you can see that there is no significant change between two profiles made using different voxel sizes (222, and common 555 mm3). Even through interpolation the whole cavity length we do not get the correct measurements near the top and bottom regions of the cavity. Therefore we exclude these regions from the analysis.

5.2.1.2 Root-mean-square velocity contours Velocity fluctuation maps u are estimated by decreasing from instant velocity maps ui calculated from fluctuation velocity as u their average velocity maps Ui u u
. .

U . Root-mean-square (rms) velocity is u


.

Fields of rms values estimated from fluctuation velocity fields after the Lagrangian to Eulerian interpolation are described in section 4.3.2. Each figure presented below shows rms of the velocity components calculated in mid-planes of the cavity in the same manner as the mean velocity fields. Contours of ux_r.m.s are shown in Figure 16, uy_r.m.s in Figure 17 and uz_r.m.s in Figure 18 in the corresponding mid-planes. In addition to the standard method of estimating fluctuating velocity we estimate the fluctuating particle field as in a one-way coupling mannerby subtracting the mean flow fields estimated from the tracer motion. In the figures below we demonstrate an additional contour map of rms velocity component where fluctuating velocity of particles is calculated by decreasing from instant particle velocity maps tracer average velocity map:
. .

, this was done in

order to see the fluctuation of particles compared to the flow average velocity.

27

(a)

(b)

(c)
Figure 16 ux r.m.s in z mid-plane (a) r.m.s velocity of tracers, (b) r.m.s velocity of particles, (c) r.m.s velocity of particles calculated using average velocity of tracers, (all axes are normalized by cavity width D).

(a)

(b)

28

(c)
Figure 17 uy r.m.s in z mid-plane (a) r.m.s velocity of tracers, (b) r.m.s velocity of particles, (c) r.m.s velocity of particles calculated using average velocity of tracers (all axes are normalized by cavity width D).

(a)

(b)

(c)
Figure 18 uz r.m.s in x mid-plane (a) - r.m.s velocity of tracers, (b) - r.m.s velocity of particles, (c) r.m.s velocity of particles calculated using average velocity of tracers (all axes are normalized by cavity width D).

29

We see that the structure of r.m.s contours is nearly identical for both particles and tracers in all velocity components. The r.m.s values of particles are slightly higher at several regions. This might be due to the fact that we measure particle fluctuations using its average profile and this does not necessary provide us a true view of the flow pattern in that region since the real particle fluctuation is from the flow field average. In addition values received in that region are compiled from fewer data sources than in other regions. Therefore, when we perform an r.m.s velocity analysis of the particles using tracers average velocity ((c) in all three figures), we observe smaller fluctuation values. This strengthens the view that the usual way of estimating particle velocity fluctuations is more sensitive to the low density of the solid particles. The alternative type of estimating fluctuations of the solid particles seems to be more accurate in this case. Similarly to the mean velocity profiles we compare the r.m.s results to those of PK89. Figure 19 (a) and (b) present profiles ux_r.m.s and uy_r.m.s respectively. The velocity profiles shown in the figures are of tracers, particles and PK89 results

(1)

(2)

Figure 19 Profiles of r.m.s velocity for middle z plane of the cavity. (1) uy_r.m.s in cross-section of z and y middle planes, (2) ux_r.m.s in cross-section of z and x middle planes. (a) Prasad and Koseff (1989) results ( SAR 1:1 Re=10000), (b) tracers velocity profile, (c) particles velocity profile. Velocity r.m.s values are presented in scale of 10:1 for PK89 results and 20:1 to our results, rescaling was done in order to save the PK89 original scales and to present the results in one figure.

As in section 5.2.1.1, we compare the results of average velocity we calculated in Eulerian analysis to that of the PK89. We see that although we have the same shape, we

30

have twice as big values as theirs. The explanation to that may be in differences between measuring and processing techniques and in understanding the definition of r.m.s velocity. 5.2.1.3 Reynolds stresses Reynolds stresses u u are representation of apparent stress arising from the

fluctuation velocity field and stems from momentum transfer by the fluctuating velocity field (Pope 2000). Contours of the Reynolds stress fields, represented in Figure 20, ensemble average of multiplications of the fluctuation velocity component ux by component uy, calculated in z mid-plane of the cavity for tracers and particles.

(a)

(b)

Figure 20 Reynolds stress contour maps at the z mid-plane (a) tracers, (b) particles.

High Reynolds stress values (both positive and negative, since sign is according to axes definition) show us the high shear regions in the cavity for both particles (a) and tracers (b). We see the same values and shapes, and as we expected the highest fluctuation area is near the downstream wall, where x velocity from maximum drives to zero, and y velocity gets its maximum negative values.

5.2.1.4 Summary of Eulerian analysis Eulerian analysis provided us with information about the general pattern of tracer and particle flow in the cavity. The information includes: location of central vortex, presence and location of high shear regions, fluctuating and average velocity values, and enabled approval of our result by comparing them to the work of Prasad and Koseff (1989).

31

Eulerian analysis cannot provide us with a close look into the flow, because our experimental and data processing methods provide us with indirect Eulerian data and involve additional processing. Besides the fact that particle fluctuating velocity values in general are higher than those of the tracers, it does not teach us of the differences between particles and tracers, which is not surprising, since all the differences are expected to be in small scales that our Eulerian analyses could not reach. Therefore, we must proceed to the second part of analyses: the Lagrangian analysis that can provide us the closer look at the smallest scales of the flow and teach us the differences between particles and tracers.

5.2.2

Lagrangian results

Lagrangian results are derived directly from Lagrangian analysis processing performed by 3D-PTV software on experiments raw data after the processing described in section 4.3.1. The results presented below include location analysis, settling estimations, spatial derivatives, autocorrelations of velocity and acceleration, acceleration analysis, velocity increments and structure functions. All analysis in the Lagrangian results section is trajectory-based. Results like correlations, acceleration statistics, velocity increments, structure functions etc. are made from trajectories that last for 10 frames and more. An example of these trajectories is presented in Figure 21 which shows long particle trajectories (more than 100 frames and 11 longest trajectories), their shape and spatial location.

(a)

(b)

32

Figure 21 Particles trajectories in 3D cavity (a) trajectories Longer than 100 frames, (b) longest particle trajectories. Red circle represent the end point of trajectory. Cube denotes the limits of the cavity.

From figures above we see that we have plenty of long trajectories, which give us enough data for our analysis. In addition we see that long trajectories are located near the center of the cavity. This could be explained by the fact that near the center of the cavity velocity values are relatively small, therefore particles move slower and are easily located compared to high velocity and strain regions which are located near the boundaries of the cavity.

5.2.2.1 Scales evaluation

In Lagrangian analysis, absolute values often do not provide us with right view and understanding of sizes of phenomena observed; therefore most of the values in this part are presented in non-dimensional. This part presents a set of calculations performed in order to evaluate several lengths, time and other scales, which describe the turbulent flow and are being used in figures in the Lagrangian section for normalization and scaling purposes. The major value on which most other values evaluations are based is the r.m.s velocity of the flow. Our method of calculating the r.m.s velocity value is to take it from contour maps of tracer r.m.s velocity (which represent the flow) in the middle regions of the cavity ( Figure 16, 18, and 19). The value that was taken represents of the r.m.s velocity common to most parts of the cavity so it probably could represent the flow. The value used is 0.01 m/s. The r.m.s velocity values represent the first velocity scale of the flow (besides the geometric scales which include belt velocity and geometric properties of the cavity), the scale which represents the length scale is called turbulent integral length scale. A theoretical method of estimating turbulent length scale is to take it as something like 1/6 of geometry scale (0.08m) (Pope, 2000); we take integral length scale l as 0.01m.

33

After evaluating the basic velocity and length scales, we can proceed to the calculation of one of the energy dissipation rate, on which all other scales (time, velocity and length) are based. The definition of energy dissipation rate is

where

,(Tennekes and Lumley 1972). Since we can't measure sij directly we use approximation formulas in order to estimate , due to the following equations
. .

, where is integral length scale. That way we receive an evaluation of energy


. .

dissipation rate: the experiment is not.

1 10

. It is important to note that this estimate is

not correct to our case since it was developed for homogeneous flows and the flow in Another length scale which can be estimated is Taylor length scale , along with Reynolds number based on Taylor length scale 15 Re ; Rel=100. Time scale used widely in analysis of turbulent flow is Kolmogorov time scale which is the smallest time scale of the flow. Using values evaluated below are 1 10 l=0.01m and
. . .

, to be 3.8 mm. Both Reynolds numbers based on Taylor length scale

. We estimate it as:

and on integral length scale make use of velocity r.m.s value and are equal to: Re = 38

0.01

we can evaluate Kolmogorov time scale () defined as:

0.1 s.
.

The smallest length scale of the flow is the Kolmogorov length scale defined as 0.31mm. Also, adopting the Gibert et al. (2010) definition of Stokes /

number for solid particles which densities vary between 1 (knowing that particle diameter is

8, and particle

diameters of order of Kolmogorov length scale as most appropriate to our particles twice as big as Kolmogorov length scale, we calculate Stokes number defined in eq. 4 in section 2.1 as 0.34. 5.2.2.2 Distribution analysis Particle distribution analysis provides us with information about particle and tracer 3D scatter inside the cavity, and therefore teaches us about the amount of data in every region. This analysis provides us with estimation of the reliability of our statistical analysis (it is known that the amount of data processed is the most important factor in the accuracy of results). This analysis was made using position data disregarding the

34

sampling time. It is important to know how the homogeneous particle/tracer distribution, and to find the regions of high/low density. Figure 22 shows the histogram of particle/tracers distributions in the cavity. In ideal experimental conditions one could expect a homogeneous, uniform distribution of particles/tracers across the cavity.

(a)

(b)

(c)
Figure 22 Particles/tracers histogram of spatial distribution, along (a) x, (b) y, (c) z, directions, respectively.

Figure 22 reflects the larger concentration of tracers as compared to particles, used in the experiments. On average, there are about 2000 tracers and 1000 particles at every location in the cavity, with the maximum in the center and slight reduction towards the edges. Clearly x and y distributions are different from the distribution along z-axis, due to the optical restrictions of the stereoscopic imaging. (Not uniform Laser beam diverging by lenses in z direction as well as growing distance from cameras focus. (see Dracos 1996)). Although the shapes of the histograms of the particles and of the tracers are similar, in order to understand if the particles probe the same flow as the tracers we present normalized location distributions in the following Figure 23. It is important to see if

35

there are some regions that particles sample differently from tracers. Probability density functions (PDF) allow for the comparative analysis of the distributions.

(a)

(b)

(c)
Figure 23 PDF of spatial particle/tracer location. (a) x location, (b) y location, (c) z location.

Figure 23 shows that particles and tracers are distributed in a similar manner throughout the cavity with lower probability near the walls. The main reasons are laser light reflection from glass walls and optical restrictions described above which caused a severe reduction in detection in those regions. 5.2.2.3 Estimation of the settling velocity effect It is important to estimate whether or not the particles we use in our experiment experience significant influence of gravity forces in the cavity. This is carried out in addition to the preliminary experiment, comparing vertical components of particle velocity to that of the tracers. In Figure 24 we present the probability density function of the particle and tracer U velocity component.

36

Figure 24 PDF of vertical velocity component Uy. (a) tracers (average value is 0.012 m/s), (b) particles (average value is 0.007 m/s).

As we see from the Figure 24, the shape is almost the same for tracers and particles. Positive average value of both particles and tracers is caused by detection problem due to light shading effect in high negative velocity regions near high downstream wall, this effect is more dominant for tracers than for particles. Although the solid particles experience lower (on average) vertical velocity due to their settling (negative) velocity, they do not settle during the experimental run. We show that the number of particles in the cavity (see Figure 25) per frame is almost constant.

37

Figure 25 Number of detected particles as function of frame number, with average number line (black) of 274 detected particles.

5.2.2.4 Spatial derivatives A spatial derivative is one of the flow parameters, which indicates velocity transport in the flow. We cannot estimate it directly from knowledge of velocity in a constant location, since our Eulerian analysis does not gives us enough precision. Therefore we measure particle velocity derivatives along the trajectory. Normal spatial derivative is calculated for long trajectories (longer than 10 frames). Derivation tracks the change in fluctuation velocity as function of distance between two derivate points on the trajectory. The problem is that average and fluctuation velocity in Lagrangian system of coordinates is not defined. This process, which is quite simple in Eulerian coordinates systems, becomes very complicated in Lagrangian ones, since in the definition of fluctuating velocity presents an average velocity value which is subtracted from instantaneous velocity. In a system of coordinates where the origin follows the particle and all that is known is the particle spatial location and its velocity and acceleration in that location at a specific time, there is no simple definition of local average velocity that can be used to find particle fluctuations in that location. Our way of dealing with this problem (since we need to estimate fluctuating velocity in most sections below) is to define our average velocity in the Lagrangian system. 38

The definition is as follows: knowing particle trajectory, we define the average velocity component of a particle as the linear trend of its instantaneous component along the trajectory. Therefore the fluctuation of particle velocity in time is its components deviation from the linear trend as shown in Figure 26. A straight line presents linear trend of velocity component, and a curve represents velocity component evolution along trajectory in time. The difference of velocity component from trend line is its fluctuation.

Figure 26 (Left) Plot of x component of particle velocity for several insulated trajectories in time, each blue line represent single trajectory. (Right) Velocity fluctuation v in respect to the linear fit representing average velocity.

Velocity components for trajectories shorter than 100 frames look like almost straight lines, without breaking its trend, while velocity of longer trajectories usually experience more complicated evolution and do not looks like a straight line, and thus linear trend is not a quite good estimation of their average velocity. That is why statistical data out of long trajectories for long times (more than 100 frames) received this way should be trusted carefully. Statistically, these long trajectories do not play a significant role, since they make up less than 5% out of all trajectories. Figure 27 (a) shows the derivation of instantaneous particle/tracer velocity and Figure 27 (b) shows the derivation of fluctuating velocity calculated as described above. The derivation is performed following the presented formula between each set of two adjusted points on the trajectory : , , . (14)

( is vector tangential to trajectory connecting each set of adjusted points). 39

(a)

(b)

Figure 27 Spatial derivatives. (1) Spatial derivative of full velocity, (a) tracer (average value is 25.5 1/s2 ), (b) particle (average value is 25.8 1/s2). (2) Spatial derivatives of velocity fluctuations. (a) tracer (average value is 23.9 1/s2 ), (b) particle (average value is 24.4 1/s2 ).

From Figure 27 we see that, as expected, there is no difference between fluctuating and full velocity component derivatives, neither in values nor in shape, for both tracers and particles. It shows that the subtraction of linear fit for each component of velocity does not introduce new errors to the spatial derivatives. Using the spatial derivatives shown in Figure 27 we can provide another estimate of the dissipation rate. Based on the known estimate for the isotropic case u 15 , u

(Tennekes and Lumley, 1971), and correcting it for the spatial , we obtain = 4.25 10-5 m2/s3 .

derivatives

5.2.2.5 Lagrangian correlation of velocity Lagrangian velocity autocorrelation function is for turbulent diffusion study and can be used to characterize the time and length scales of the flow or of the moving particles. Figure 28 presents a schematic view of a particle trajectory, showing two points on it with the velocity components, which is used in the definition of Lagrangian autocorrelation functions.

40

Figure 28 Particle location and velocity at two different times on its trajectory.

Figure 29 (a) presents Lagrangian velocity autocorrelation of tracers (solid lines) and particles (symbol lines). Correlation coefficient of velocity vector is obtained according to equation R

(15)

where v is the Lagrangian velocity vector and in time between two correlated points).

is the lag time (represents a difference

The correlation coefficient of Lagrangian velocity vector projections on orthogonal axes is calculated according to formula where v is the i component of velocity vector (i = x, y, z). Figure 29 (b) presents the Lagrangian fluctuating velocity autocorrelation of particles (symbols lines) calculated according to formula R v

R v

(16)

where v is the i component of velocity vector obtained with the process described in section 5.2.2.4 and full present autocorrelation of vector combined from fluctuation components.

(17)

41

(a)

(b)

Figure 29 Lagrangian velocity autocorrelation coefficient (a) tracer (represented by solid lines curves) and particle (represented by symbols) autocorrelation coefficient of Lagrangian velocity vector and its components, (b) tracer and particle autocorrelation coefficient of fluctuating velocity vector and its components. (full-Rvv , Rxx , Ryy , Rzz time is normalized by Kolmogorov time scale which is equal to 0.1s ) .

We observe in Figure 29(b) that the correlation time scale of fluctuating velocity components is shorter if compared to the full velocity autocorrelation, shown in Figure 29(a). This analysis demonstrates that most of the correlation is due to the non-local effects and large-scale velocity patterns, imposed on the local variations of the Lagrangian velocity. This is to emphasize the inability of the undecomposed Lagrangian correlation analysis to predict dispersion of particles (continuously growing distance between the neighbor particles on relatively short time scales). The presented autocorrelation analysis has to be taken with caution, since the flow is not homogeneous and there is dependence between the values of the autocorrelation and the region in the flow where the trajectory was detected. Nevertheless, we conclude that in general, particles have stronger and longer velocity correlations as compared to tracers (almost twice as long). This can be seen clearly in the example of the correlation of the v component In Figure 29 (a) and (b). 5.2.2.6 Autocorrelation of acceleration The acceleration correlation is defined in the same way as the velocity correlation. The correlation in time of the acceleration vector and its components may provide us with some understanding of dominant small scale forcing terms, like Bourgoin et al (2009), who suggest that the dominant small scale forcing term acting on particles is pressure 42

gradient. From the acceleration correlation we can get parameters such as time scale based on integration of acceleration autocorrelation coefficient from zero to zero crossing time. R 0. R s ds , where is the first zero crossing time, i.e.

Bourgoin et al (2009) claim that

remains of the Kolmogorov dissipation time of

the carrier flow for a variety of particle densities and diameters; for which Stokes number changes by several orders of magnitude from the smallest neutrally buoyant

particles to the largest and heaviest particles, and thus claiming that the Stokes number effect alone cannot properly characterize and model such inertial particles. The correlation was calculated for the full acceleration vector and for the tangential to trajectory component of the acceleration vector (sees Figure 30).

Figure 30 Sketch of acceleration vector with its components of two points on particles trajectory.

is the full acceleration vector correlation R


(18)

a - acceleration vector, trajectory. R

- lag time - a time interval starting from time between two

frames and ending with time interval passed between beginning and the end of the

is the tangential components correlation of acceleration vector R


(19)

n - unity vector tangential to trajectory (see Figure 30).

43

(a)

(b)

Figure 31 Acceleration autocorrelation function. (a) Autocorrelation function of full acceleration vector, Raa (b) autocorrelation function of tangential to trajectory component of acceleration vector RaTaT. Time is normalized by Kolmogorov time scale which is equal to 0.1s.

From Figure 31 (a) we see that there is minor difference in shape and value between tracers and particles. Also the autocorrelation acceleration of tangential projection on trajectory did not change significantly in shape and value but there are some vibrations of particle projection autocorrelations which are not presented in tracer projection autocorrelation ( see Figure 31 (b)), these changes are supposed to be due to different statistics of long trajectories. By calculation, component acceleration correlation. for full acceleration

correlation is 0.35 for both tracers and particles, and 0.25 and 0.3 for tangential

5.2.2.7 Acceleration analysis Some say that acceleration statistics can clearly resolve the transition from tracer particle to large particle behavior claimed by Brown et al. (2009), while others disagree claiming that for fixed carrier flow conditions, acceleration statistics of finite size inertial particles are very robust to size and density variations (Bourgoin et al. 2009). In particular, the shape of normalized acceleration probability density function remains unchanged. We checked our particle and tracer acceleration statistics to see which approach is more correct for our case. Acceleration values for this analysis are also taken only for those particles/tracers for which trajectories last for at least 10 frames, since data from these trajectories have been used in other analyses (see 5.2.2). Figure 32,Figure 33 and Figure 34 present probability density functions (defined as before) of acceleration components of particles and tracers 44

of the x, y, z components of acceleration vectors, respectively. In these figures the left panel (a) represents PDF of the dimensional acceleration, and the right panel (b) represents the same distribution but of the values normalized with the root mean square of acceleration component a / a
.

(a)
<ax2>0.5=0.17 m/s2 .

(b)

Figure 32 Acceleration in direction of x axis (ax), tracers r.m.s <ax2>0.5=0.16 m/s2, particles r.m.s

(a)
<ay2>0.5=0.17 m/s2 .

(b)

Figure 33 Acceleration in direction of y axis (ay), tracers r.m.s <ay2>0.5=0.14 m/s2, particles r.m.s

45

(a)
<az2>0.5=0.3 m/s2 .

(b)

Figure 34 Acceleration in direction of z axis (az), tracers r.m.s <az2>0.5=0.3 m/s2, particles r.m.s

Our results confirm the robustness of the acceleration probability density functions (Bourgoin et al., 2009) it is similar for the particles and tracers. Therefore, the PDF cannot be used as a probe for the different behavior of the particles in respect to the tracers. The single point analysis that combines the statistics that are obtained disregarding the time history of the trajectory is useless for describing the evolution (dispersion, diffusion) of particles. Estimates of the dissipation rate using r.m.s values of acceleration can be done using the formula a
.

a .

where a0 is between

1 and 10. We have acceleration r.m.s values of long (more than 10) tracer trajectories around 0.15 m/s2. Therefore the values of dissipation are in the range of 10-4 m2/s3. 5.2.2.8 Velocity increments Velocity differences (increments) measured along Lagrangian trajectories at varying time gaps provide a signal closely related to acceleration, and the change in the shape of the PDF is typically related to the intermittency of the turbulence (e.g. Toschi and Bodenschatz 2009). In a view of the previous result it is important that this PDF does include some history effect, as it is taken for two-point statistics at different time separations. In Figure 35 we present the probability density function of the velocity increments function defined as v v t v t , which is calculated for the same

data as the acceleration statistics. Time lags are dependent of the frame rate and are calculated for increments of 1, 2, 4, 20 and 50 frames, which are in our experimental sets (I and II): 0.0067, 0.013, 0.026, 0.133 and 0.333 seconds or 0.07, 0.15, 0.3, 1.5 and 3.5 (in Kolmogorov time scale).

46

(a)

(b)

Figure 35 Velocity increments. (a) velocity increments of tracers, (b) velocity increments of particles. Time lags are 0.07 0.15 0.3 1.5 3.5 , and displayed with a vertical shift from 0.07 (up) to 3.5 (down) for clarity.

We observe a continuous change in PDF shape from lines representing the short-timelags to lines representing long-time-lags. The shape of short-time-lags curve looks sharp and thick compared to long time lags curves. This shape teaches us of a weak correlation between the velocity values in adjusted times therefore indicating the strength of the intermittency in our flow. There is a little difference between Figure 35 (a) and (b) which cannot emphasize some major difference between particles and tracers as we see also from the acceleration statistics. Therefore, there is a need to present velocity increment statistics in other more specific forms to see the dissimilarity between them.

5.2.2.9 Structure function The velocity increments are shown as probability density functions and we can also average the differences at different time lags to construct the structure functions. The Lagrangian structure functions are not directly related to the Eulerian structure functions, which are used in the core of Kolmogorov theory of turbulence (Kolmogorov 1941). The Lagrangian velocity structure functions are estimated according to: S v t v t (20)

where i is the orthogonal component of velocity. In Figure 36 we demonstrate structure functions of the second order with time lags normalized by the Kolmogorov time scale

47

. The convenience of structure functions is that they use the Lagrangian trajectories. Similar to the Eulerian structure functions, we expect to observe structure functions that show some asymptotic behavior if normalized with average dissipation. Therefore, we expect to see a horizontal portion of S2 when plotted versus time lag.

Figure 36 Lagrangian second order structure function of particles and tracers x and y velocity components, lag is lag time from minor time interval between adjusted point to maximum interval between start and the end point of the trajectory, is Kolmogorov time scale of the flow.

From Figure 36 we a plateau between the values of

=1 to 7 which reminds us the

asymptotic behavior which present in Eulerian structure functions; it is clearly seen for tracers and less clear for particles. Also, the velocity structure functions, which are sort of "correlation of increments", for particles are stronger than for tracers as we can see from the values in Figure 37 (particles reach 0.8-0.9 and tracers 0.6). The shape is similar for both tracers and particles, with the same change in dominating component correlation from x to y, when normalized time lag value is approximately equal to one. . 5.2.2.10 Logarithmic derivatives Logarithmic derivatives of structure functions are used usually in order to emphasize the intermittency of the turbulence. In our case these, are used to verify the intermittency as it is shown by the flow tracers and particles. These functions are defined as , when the results represent some sort of local (at scale ) scaling exponents. 48

Figure 37 and Figure 38 show logarithmic derivatives of order 4 (p=4) and 6 (p=6) , respectively, as a function of time lag lag normalized by the Kolmogorov time scale , calculated from structure functions defined and evaluated as in the previous section (section 5.2.2.9).

(a)

(b)

Figure 37 Logarithmic derivatives of order 4. (a)-tracers, (b)-particles, lag is lag time and is Kolmogorov time scale of the flow.

(a)
Kolmogorov time scale of the flow.

(b)

Figure 38 Logarithmic derivatives of order 6. (a)-tracers, (b)-particles, lag is lag time and is

In Figures 37 and 38 above we observe a dip near the scales where the inertial range we found in the previous section (5.2.2.10, 5.2.2.9). The dip represents the border between dissipative and larger scales of the turbulence (e.g. Toschi and Bodenschatz, 2009), and we clearly see that the dip in the tracer result is steeper than in the particle data in both exponents (p=4 and p=6). To the best of our understanding, this is because particles

49

tend to leave high shear areas faster than the tracers, thus showing less intermittent behavior, or in other words, low-pass filtering of the underlying turbulent flow. 5.2.2.11 Mixed velocity-acceleration structure function density on mixed velocity acceleration structure function v During the analysis of the dynamics of heavy particles, the strong effect of particle a was observed

(Gibert et al. 2010) that can point on preferential alignment between velocity and acceleration increments. We are interested to test if the mixed structure functions in time are also sensitive to the size variations rather than density variations of the v a v a v t v t a t

particles. In Figure 39 we present mixed structure function obtained for the particles and tracers a t a t where: v t v t a t a t v t v t

a t .

Figure 39 Lagrangian time mixed structure function, lag is a lag time and is Kolmogorov time scale of the flow.

In Figure 39 we observe that for longer time lags (

4) the particles show lower

values, pointing towards the lower dissipation values, or, in other words, weaker intermittency and less variations than the tracers. Gibert et al. (2010) showed that for particles with various densities and the same average diameter, the mixed structure function equals a constant (approximately -2) in the inertial range. The constant value decreases with density and equals one for neutrally buoyant tracers. If we try to 50

transform our resulting time scale to length scale, we must expect some constant values and the same as Gibert et al. (2010) effects in range of 4, but in fact we see a

different result than what we expected: tracers with a small diameter compared to particles show smaller values than particles, and so only for particles we can assign some sort of constant value for 5 / 1.

These results could be explained by differences in experimental setups and statistical processing, or this could be a real difference due to a diameter variation. Both reasons may be proven through additional experiments with larger variety of particle diameters and densities or Reynolds and Stokes numbers that are used to define the parameter space. It is worthy of mention that there is a clear difference between particles and tracers and thus the mixed time structure function could be used as an indicator of the different Lagrangian motion of the particles of varying size is similar to the case of particles of with different density. In order to see if the effect is seen in the dimensionless units, we normalize the mixed structure function with the absolute values of the velocity and acceleration increments and obtain the so-called cosine of the angle between the velocity-acceleration increments( cos
| || |

). 1,

The result is shown in Figure 40. It is shown here that on average, the correlation between velocity and acceleration is lower for the large particles at extremes, turbulent flow.

enforcing the filtering behavior of larger than Kolmogorov length scale particles in

51

Figure 40 Probability density function (F) of the cosine of the angle between v and a

From Figure 40 we see that there is a clear tendency of both particles and tracers toward negative values, which point out that the two vectors (velocity and acceleration), are more likely to point in opposite directions. Tracers demonstrate these phenomena significantly compared to particles, which demonstrate almost no change in angle. Again, this points out to the filtering effect of large particles, which are less sensitive to fast changes in velocity vector direction, i.e. strong accelerations.

6 Summary and conclusions


Weve presented the experimental study of particles and tracers motion in turbulent flow with emphasis on the difference between them. The experimental technique used is the three-dimensional particle tracking velocimetry (3D-PTV). Using this technique with additional processing enables us to obtain particle and tracer Lagrangian properties, i.e. trajectory along with velocity and acceleration. The experimental volume is a cubic lid-driven cavity (LDC) flow at Reynolds number 10,000, based on lid velocity and length of the cavity. Results are presented to provide an insight on (a) particles and tracers behavior in a turbulent LDC flow, (b) differences in behavior between large particles and small tracers, and (c) statistical properties that emphasize these differences. Summarizing the experimental results we can conclude that:

52

Eulerian

analysis

performed

on

data

transformed

from

Lagrangian

experimental system provides us with a general view on the flow pattern but does not emphasize the differences in particle-tracer behavior. Acceleration analysis is truly robust to size and density variations. Structure functions of velocity components along with mixed acceleration velocity structure functions best emphasize the difference between particles and tracers and could become the probe property in experiments dealing with particles. There is are extensive difficulties in working with Lagrangian data received in experiments with 3D turbulent flow without one main flow direction, as a result of lack of proper defined fluctuating and average velocity values in trajectory laden data. Evaluation of the dissipation rate value that was done in all three ways (r.m.s value of velocity components, r.m.s value of acceleration components and spatial derivation value) provided us with dissipation rate values of the same order. Therefore evaluation of this flow parameter for the presented experimental setup can be done with a decent accuracy by any of the three ways. Concluding our studies we suggest performing a set of experiments using particles with a variety of diameters and densities in order to see if the conclusions we made are valid. In addition, we need to extend the resolution of the experimental systems in order to get a more precise view on the smallest scales along with faster data capturing which will allow us to perform experiments with much higher velocities.

53

7 Bibliography
1. Bourgoin M., Qureshi N., Baudet C., Cartellier A and Gagne Y., Lagrangian statistics of inertial particles in turbulent flow, B.Eckhardt(ed.), Turbulence 12 , Springer Proceedings in Physics 132, 2009. 2. Brennen C.E. (ed.), Fundamentals of multiphase flows, Cambridge University Press, 2005. 3. Brown R.D., Warhaft Z. and Voth G.A., Acceleration Statistics of Neutrally Buoyant Spherical Particles in Intense Turbulence, Phys. Rev. Lett., 103:194501, 2009. 4. Calzavarini E., Volk R., Bourgoin M., Leveque E., Pinton J., and Toschi F.,., Acceleration statistics of finite-sized particles in turbulent flow: the role of Faxn forces, J. Fluid Mech. 630: 179189, 2009. 5. Clay Mathematical Institute, http://www.claymath.org/millennium/Navierstokes_Equations 6. Crowe C.T., Schwarzopf J.D., Sommerfeld M. and Tsuji Y., Multiphase Flows with Droplets and Particles, CRC Press, 1998. 7. Dracos Th. (ed.), Three-Dimensional Velocity and Vorticity Measurements and Image Analysis Techniques, Kluwer, 1996. 8. Elfassi R., Experimental study of two- and three-dimensional aspects of lid-driven cavity turbulent flows, M.Sc. thesis, Tel Aviv University, 2010. 9. Fallon T. and Rogers C.B., Turbulence-induced preferential concentration of solid particles in microgravity conditions, Exp. Fluids, 33: 233-241, 2002. 10. Gasteuil Y., Shew W.L., Gibert M., Chill G., Castaing B., and Pinton J.-F., Lagrangian Temperature, Velocity, and Local Heat Flux Measurement in RayleighBnard Convection, Phys. Rev. Lett. 99, 234302, 2007. 11. Gibert M., Xu H. and Bodenschatz E., Where do heavy particles go in a turbulent flow, submitted to Phys. Rev. Lett., 2010. 12. Gore R.A. and Crowe C.T., Effect of particle size on modulating turbulent intensity, Int. J. Multiphase Flow, 15: 279-285, 1989. 13. Guala M., Liberzon A., Hoyer K., Tsinober A. and Kinzelbach W., Experimental study on clustering of large particles in homogeneous turbulent flow, Journal of Turbulence, 9: 34 1-20, 2008. 14. Hoyer K., Holzner M., Lthi B., Guala M., Liberzon A. and Kinzelbach W., scanning particle tracing velocimetry, Exp. Fluids, 39(5) 923-934, 2005. Advances in

54

15. Lthi B., Tsinober A. and Kinzelbach W. Lagrangian measurements of vorticity dynamics in turbulent flow, J. Fluid Mech, 528: 87-118, 2005. 16. La Porta A.,Voth G.A.,Crawford A.M., Alexander J. and Bodenschatz E., Fluid particle accelerations in fully developed turbulence, Nature 409, 1017-1019, 2001. 17. Maxey M.R. and Riley J.J., Equation of motion for a small rigid sphere in a non uniform flow, Phys. Fluids 26(4):883-888, 1983. 18. Mordant N, Crawford A.M., and Bodenschatz E., Three-Dimensional structure of the Lagrangian acceleration in turbulent flows, Phys. Rev. Lett. 93, 214501, 2004. 19. Ott S. and Mann J., An experimental investigation of the relative diffusion of particle pairs in three-dimensional turbulent flow, J. Fluid Mech, 422:207-223, 2000. 20. Ouellette N.T., O'Malley, P.J.J., Gollub J.P., Transport of finite-sized particles in chaotic flow, Phys. Rev. Lett., 101, 174504, 2008. 21. Pope S.B. (ed.), Turbulent Flows, Cambridge University Press, 2000. 22. Prasad A.K. and Koseff J.R., Reynolds number and end-wall effects on a lid-driven cavity flow, Phys. Fluids, 1(2): 208-218, 1989. 23. Qureshi N.M., Bourgoin M., Baudet C., Cartellier A. and Gagne Y., Turbulent Transport of Material Particles: An Experimental Study of Finite Size Effects, Phys. Rev. Lett., 99:184502, 2007. 24. Sato Y. and Yamamoto K., Lagrangian measurement of fluid-particle motion in an isotropic turbulent field, J. Fluid Mech., 176: 183-199, 1987. 25. Shirolkar J.S. Coimbra C.F.M. and McQuay Q., Fundamental aspects of modeling turbulent particle dispersion in dilute flows, Prog. Energy Combust. Sci., 22: 363399, 1996. 26. Shlien D.J. and Corrsin S., Measurement of Lagrangian velocity autocorrelation in approximately isotropic turbulence, J. Fluid Mech., 62(2) :255-271, 1974. 27. Snyder W.H. and Lumley J.L., Some measurements of particle velocity autocorrelation functions in a turbulent flow, J. Fluid Mech. 48, 41, 1971. 28. Tanaka T. and Eaton J.K., Classification of turbulence modification by dispersed spheres using a novel dimensionless number, Phys. Rev. Lett., 101: 114502, 2008. 29. Tennekes H. and Lumley J.L. (ed.), A First Course in Turbulence, MIT Press, Boston, 1972. 30. Toschi F. and Bodenschatz E., Lagrangian Properties of Particles in Turbulence, Annu. Rev. Fluid Mech., 41:375-404, 2009.

55

31. Virant M. and Dracos T., 3D PTV and its application on Lagrangian motion, Meas. Sci. Technol., 8:15391552, 1997. 32. Voth G.A., Bigger B., Buckley M. R., Losert1 W., Brenner M. P., Stone H. A., and Gollub J. P., Ordered Clusters and Dynamical States of Particles in a Vibrated Fluid, Phys. Rev. Lett. 88:234301, 2002. 33. Yeung P.K. and Pope S.B., Lagrangian statistics from direct numerical simulations of isotropic turbulence, J. Fluid Mech., 207: 531-586, 1989.

56

APPENDIX A: UNCERTAINTY ANALYSIS We present here the uncertainty analysis of 3D-PTV experiment following the method presented in Elfassi (2010), who used the same experimental setup, the same calibration files, and investigated similar Reynolds number flows. Introduction The aim of any experiment is to measure some quantities as accurately as possible. Therefore, the error in the measurement must be estimated to arrive at the conclusion about correctness of the results obtained. Uncertainty analysis is the procedure used for quantity data validity and accuracy. Uncertainty NR of the function R is due to the experimental measurements). It may be obtained by: N n n .. n
.

combined effect of uncertainty intervals

is the independent parameter in the

( i=1..k ) are the relative uncertainties of individual follow, R x ; n ;n .

. The effect of error in measuring

on R may be estimated by analogy to the derivative of a function as

In the following we estimate uncertainty of the velocity and acceleration in 3D-PTV experiments. The maximum instantaneous velocity that one expects to capture is equivalent to the belt velocity 0.125 m/s. Uncertainty analysis of the 3D-PTV measurements: In the following we review the main errors that affect the 3D-PTV measurements. The errors are combined from the optical setup, calibration, experimental errors and data processing. In each experiment we acquire a sequence of images in the LDC setup and process it in order to get the velocity and acceleration field. The main error comes from the calibration process, and it affects the determination of the particle position in time (Lagrangian data). The measurements of the calibration target play an important role in the calibration process. The errors in calibration procedure are due to the error in particle position and image processing. We can estimate the image processing errors (estimation of the particle center after the high-pass filter) to be of the order of pixel. The conversion of pixels to millimeters depends on the magnification factor, which in our setup was 0.1 (mm/pixel). In addition, there are errors from imprecision in the calibration target production (distance between the points is nominally 5mm0.1mm) 57

and errors due to the optical aberrations of the lenses. In the present 3D-PTV system we use 4 cameras and the particle position in 3D space is estimated using 4 particle positions in images (x, y in pixels). The stereo-photogrammetric matching of the particles and reconstruction of the 3D position is the main source of position error. The software provides the r.m.s estimate of the error using known 3D calibration target (with multiple dots) which in our case was 10m for x and y position and 30m for z position. The time difference between the frames is t (1/frames-per-second) which is determined by the synchronization unit with very high precision (the error is prescribed to be 3 orders of magnitude smaller than the interval). In general, the velocity vector is estimated using the displacement of a given particle between two frames, divided by the time interval: u t a t x t t t x t

And the acceleration is calculated in the similar manner: u t t t u t

In the next table we summarize the calculations of the errors, and the uncertainties of the different parameters,
Table A1: The different errors and uncertainties in 3D-PTV process

Parameter x y z

xi units 0.8 0.8 0.8 mm mm mm s -150 150 150 9000

0.0125 0.0125 0.0375 0.01 0.01 0.03

units mm mm mm s

t 0.006667

0.0003 0.000002

The error propagation is estimated according to the following equation: u x t u u u x t


x t

u u x t t

t u u u t

u t

x t

t t

x t

58

Table A2: Velocity and acceleration values

Parameter u v w |

xi 120 120 120 120 18000

units mm/s mm/s mm/s mm/s mm/s2

| | get of only one component maximum values since there is no place in the cavity where two velocity components get maximum values.

In summary we can estimate the uncertainty of the velocity and acceleration measurements (both components and the magnitude of the vector) as:
Table A3: summary of uncertainties of velocity and acceleration

0.01768

0.01768

0.05303

0.05

0.1

And all the other parameters are: X=0.8 mm 1.25% Y=0.8 mm 1.25% Z=0.8 mm 3.75% t = 0.00667 0.03% U = 120 mm/s 5% a = 31176 mm/s2 10%

59


, . - ) (. . , , . . )3891( Maxey & Riley . Stokes . Kolmogorov . , . particle tracking velocimetry . , . 01 . LDC 000,01 08 " . ), ( second order structure function .

'
" "

"

0102

'
" "

0102

You might also like