• No results found

Investigations on the local entrainment velocity in a turbulent jet

N/A
N/A
Protected

Academic year: 2021

Share "Investigations on the local entrainment velocity in a turbulent jet"

Copied!
20
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

Phys. Fluids 24, 105110 (2012); https://doi.org/10.1063/1.4761837 24, 105110

© 2012 American Institute of Physics.

Investigations on the local entrainment

velocity in a turbulent jet

Cite as: Phys. Fluids 24, 105110 (2012); https://doi.org/10.1063/1.4761837

Submitted: 01 June 2012 . Accepted: 27 September 2012 . Published Online: 23 October 2012 M. Wolf, B. Lüthi, M. Holzner, D. Krug, W. Kinzelbach, and A. Tsinober

ARTICLES YOU MAY BE INTERESTED IN

Invariants of the velocity-gradient, rate-of-strain, and rate-of-rotation tensors across the turbulent/nonturbulent interface in jets

Physics of Fluids 20, 055101 (2008); https://doi.org/10.1063/1.2912513

Turbulent mixing of passive scalar near turbulent and non-turbulent interface in mixing layers

Physics of Fluids 27, 085109 (2015); https://doi.org/10.1063/1.4928199 Large-scale eddies and their role in entrainment in turbulent jets and wakes

(2)

Investigations on the local entrainment velocity

in a turbulent jet

M. Wolf,1,a)B. L ¨uthi,1M. Holzner,1D. Krug,1 W. Kinzelbach,1and A. Tsinober2

1Institute of Environmental Engineering, ETH Zurich, 8093 Zurich, Switzerland 2School of Mechanical Engineering, Faculty of Engineering,

Tel Aviv University, Tel Aviv 69978, Israel

(Received 1 June 2012; accepted 27 September 2012; published online 23 October 2012)

We report an experimental analysis of the local entrainment velocity in the self-similar region of a turbulent jet. Particle tracking velocimetry is performed to determine the position of the convoluted, instantaneous turbulent/non-turbulent interface and to compute velocity and velocity derivatives in the proximity of the interface. We find that the local entrainment velocity is mostly governed by a viscous component and that its magnitude depends on the local shape of the interface. It is illustrated that local entrainment is faster for surface elements concave towards the turbulent region. A closer analysis of the plane spanned by mean and Gaussian curvature reveals that de-pending on the surface shape, different small-scale mechanisms are dominant for the local entrainment process, namely, viscous diffusion for concave shapes and vortex stretching for convex shapes. Key quantities influencing viscous diffusion and vortex stretching in the entrainment process are identified. It is illustrated that the viscous advancement of the interface into the non-turbulent region mostly depends on the shape of the enstrophy profile normal to the interface. The inviscid contribution is in-timately related to the alignment of vorticity with the eigenvectors of the rate of strain tensor. Finally, the analysis substantiates that the convolution of the instantaneous in-terface is driven by the advection of the underlying fluid together with a contribution from the local entrainment velocity, with the advection velocity being the governing part.C 2012 American Institute of Physics. [http://dx.doi.org/10.1063/1.4761837]

I. INTRODUCTION

Many flows in nature and technology occur as partly turbulent flows with co-existing turbulent and non-turbulent flow regions separated by a sharp interface. Examples for such flows are smoke plumes from chimneys, clouds, effluent jets from waste-water outlets, atmospheric boundary layers, or wakes of aircrafts. An important process taking place at the turbulence boundary of those flows is the continuous incorporation of surrounding, irrotational fluid into the turbulent flow region. This process, known as turbulent entrainment (TE), has a direct impact on dynamics and mixing, thereby controlling transport rates of momentum, energy, heat, and mass.1Although the mechanism of the entrainment process has been studied for a long time, there are still many open questions that significantly hinder fundamental understanding, proper parametrization, and accurate modelling.2,3 An early experimental and theoretical study by Corrsin and Kistler4suggested that entrainment is governed by small-scale eddy motions, acting to diffuse vorticity outwards from the turbulent into the irrotational flow region. Recent investigations have shown supporting findings, see, e.g., Refs.5–7, and it is therefore widely accepted now that the main mechanism by which nonturbulent fluid becomes turbulent as it crosses the interface involves viscous diffusion of vorticity. Nevertheless, it is also well-known that the global entrainment rate is imposed by large-scale eddy motions,2,8–10

a)wolf@ifu.baug.ethz.ch.

(3)

reflecting that small and large scales of motion must be coupled in some way.1,11The two different ideas can be reconciled by arguing that the global entrainment flux Q determined by the mean entrainment velocity ue and a projected area A0, i.e., Q = ueA0 is also characterized by a small scale local entrainment velocityvnin combination with a considerably larger, strongly convoluted, instantaneous interface area, so that Q = ueA0=



vnd A,12,13wherevnis the velocity of an interface element relative to the fluid.14 However, so far, little is known in which way the local entrainment velocity and the strongly convoluted interface area adjust to account for the viscosity independent entrainment flux. In a similar spirit as the mathematical description of an iso-scalar surface shown in Refs.15and16, Holzner and L¨uthi17derived an expression relating the local entrainment velocity,

vn, to enstrophy production,ωiωjsij, the viscous term,νωi∇2ωi, and the gradient of enstrophy,∇ω2, so thatvncan directly be linked to the small scale mechanisms influencing the entrainment process. They experimentally and numerically investigated the turbulence boundary in a zero mean shear flow and showed that the local entrainment velocity is determined by the interface curvature and the concavity of the vorticity profile normal to the interface. Furthermore, Bisset et al.5investigated a direct numerical simulation (DNS) of a turbulent wake behind a flat plate showing that conditional averages of vorticity and velocity at the turbulent/non-turbulent interface vary depending on the local geometry of the interface. They studied the influence of surface shapes on vorticity generation and pointed out that the curvature of the interface is an important parameter for the overall understanding of the entrainment process.

The main objective of the presented study is a systematic analysis of the local entrainment velocity vn in a flow with mean shear. We used three-dimensional particle tracking velocimetry (3D-PTV) to determine experimentally the fluctuating, instantaneous turbulence boundary as well as velocity and velocity derivatives in the proximity of the interface. Two different methods for the determination ofvnare used for the evaluation of an enstrophy threshold suitable for the detection of the turbulent/non-turbulent interface (TNTI). Furthermore, the local entrainment velocity is related to small scale mechanisms as well as to the shape of the interface. Our results show that viscous effects mainly contribute to local entrainment. However, for strong curvatures one can single out different small scale mechanisms that dominate the outward spreading of the interface, i.e., vortex stretching for convex shapes and viscous diffusion for concave shapes.

The paper is organized as follows. In Sec.IIwe assess the newly designed experimental facility as well as the quality of the performed particle tracking velocimetry (PTV) measurements. SectionIII

contains the main results. We describe the determination of the local entrainment velocity and discuss the dependency ofvnon the shape of the turbulence boundary. Finally, in Sec.IVthe findings of the study are summarized.

II. EXPERIMENTAL APPROACH

Particle tracking velocimetry was performed to investigate turbulent entrainment in a submerged water jet. PTV is a non-intrusive flow measurement technique using neutrally buoyant tracer particles to determine velocity and velocity gradients along particle trajectories. Two types of measurements were carried out focusing on different aspects of the flow. First, two-dimensional measurements were conducted in the center plane of the axisymmetric jet to characterize the general flow field of the newly designed experimental facility. The investigated area is indicated by the large blue rectangle in Figure1. The main goal of the 2D-PTV measurements was to get a large scale picture of the flow field, and to check, within the achievable experimental accuracy, whether the jet flow exhibits self-similarity in the region where the local entrainment analysis should be conducted. Second, 3D-PTV measurements for the turbulent entrainment analysis were performed in a small interrogation volume at the boundary between turbulent jet and irrotational ambient fluid. The interrogation region is indicated by the small red square on the right in Figure1. The interrogation volume was restricted to a limited region around the interface in order to achieve the necessary spatial resolution to determine velocity and acceleration derivatives.

A sketch of the closed-loop water jet facility is shown on the left in Figure1. The setup consists of a vertical jet emerging from the end of a pipe (diameter d= 3 mm, length l = 240 mm) into a transparent cylinder of diameter D= 300 mm and length L = 2000 mm. The ratio of pipe length to

(4)

112.40 mm 31.02 mm 300 mm 2000 mm 240 mm cylindrical glass tank valve constant-head reservoir constant-head reservoir overflow tank pump flowmeter jet pipe diameter: 3mm rectangular glass box stream wise position x /d 2 crosswise position x /d1

FIG. 1. (Left) Sketch of the experimental setup (not to scale). (Right) Streak visualization of the axisymmetric jet. Two recordings are fitted together in one image in order to capture the flow field from the nozzle exit up to 133 x/d downstream (nozzle diameter d= 3 mm). The image shows the superposition of 1500 images corresponding to 0.5 s of recording. The red square indicates the position of the 3D interrogation volume of the 3D-PTV measurements.

pipe diameter is 80. As is shown by Zagarola and Smits,18this ratio is considered sufficient to obtain a fully developed turbulent velocity profile at the nozzle exit. In choosing the diameter of the test section D= 300 mm, a trade-off is made between a tank that is wide enough not to confine the jet too much, yet small enough to reach the measurement domain with our optical tools. Following the experimental design of Billant et al.,19the jet axial velocity is generated by the hydrostatic pressure difference between two constant-head reservoirs. In our measurements, the mean velocity of the flow at the jet exit was 1.67 m/s, implying a jet Reynolds number (Re) of 5000. A valve and a flow meter in front of the jet pipe are used to control the flow rate ( ˆ=0.71 ± 0.017 l/min). The water circuit is finally completed by an overflow tank and a pump. A rectangular glass box, filled with the same fluid as the test section, is fitted around the cylindrical water tank to avoid image distortion due to the difference in index of refraction between the in- and outside of the circular tank.

2D-PTV measurements were conducted to characterize the general flow field of the experimental facility. The flow was seeded with neutrally buoyant polystyrene tracer particles with an average diameter of 45μm. Illumination of the particles was provided by a continuous 15 W argon-ion laser. The beam was expanded through a cylindrical lens and formed a thin light sheet of about 1 mm thickness passing through the streamwise jet axis. The flow was recorded at a frame rate of 3000 Hz using a Photron SA 5 camera (1024× 1024 pixels2, object lens of 60 mm) focusing on a field of view of 200× 200 mm2. Sequential measurements were performed starting from the nozzle exit of the jet pipe up to 400 mm downstream ( ˆ=133 x/d), tracking on average 1800 particles in the field of view. A streak visualization of the flow is shown on the right in Figure1.

3D-PTV measurements were performed aiming at the region in proximity of the turbulent/non-turbulent interface. For the measurements we used the same camera as before, equipped with a 200 mm lens in combination with an image splitter to mimic a four-camera setup, as described in Hoyer et al.20 The camera system was focused on an observation volume of 11× 11 × 4 mm3, which is indicated by a red square in the streak visualization in Figure1. The light beam of the laser was expanded to illuminate the observation volume. Choosing a non-cubic observation volume has the advantage that the number of ambiguities in the stereoscopic detection procedure is reduced, allowing for a more reliable establishment of correspondences. Thus, it is possible to seed and determine the position of a larger number of particles resulting in a higher spatial resolution. In our

(5)

(a) 0.004 0.01 0.014 −0.006 −0.002 0 0.002 0.004 0.006 0.008 0.01 x, azim3 uthal direction x 1, radial direction x2 , axial direction 0 0.01 0.02 0.03 [m/s] (b) 0.004 0.01 0.014 −0.006 −0.002 0 0.002 0.004 0.006 0.008 0.01 x3, azim uthal direction x 1, radial direction x2 , axial direction (c)

FIG. 2. Processing steps of the three-dimensional measurements (red square, Figure1): (a) 2D trajectories as obtained from one camera view, (b) 3D trajectories color coded with velocity magnitude, (c) three-dimensional vorticity field interpolated onto a Eulerian grid for one time step.

measurements we were able to track the movement and to compute the complete tensor of velocity as well as acceleration derivatives for about 200 particles simultaneously. The particle position accuracy is 0.1 mm for the x1 and x2components and 0.3 mm for the x3component, respectively. In this paper x1, x2, and x3correspond to the radial, axial, and azimuthal jet axes. The Kolmogorov length scale,η, and time scale, τη, are estimated from the average dissipation rate in the turbulent region, = 2νsijsij ≈ 100 mm2/s3, as η = (ν3/)1/4 = 0.3 mm and τη = (ν/)1/2 = 0.1 s. The average particle distance in our measurements was 1.3 mm yielding a spatial resolution of about 4η. Recordings were taken at a frame rate of 500 Hz corresponding to a temporal resolution of 1/50τη. The measurements were performed from 92 x/d to 96 x/d downstream of the nozzle exit.

Figure2depicts relevant processing steps of the three-dimensional measurements. First, a spatio-temporal tracking algorithm20determines trajectories of fluid particles in image space. An example of the 2D trajectories in image space for one camera view is shown in Figure 2(a). Subsequently, three-dimensional trajectories in object space are constructed through stereoscopic matching of different camera views and Lagrangian quantities, e.g., velocity and acceleration, are calculated along the trajectory. Figure 2(b)shows an example of three-dimensional trajectories color-coded with its corresponding velocity magnitude. Given that the smallest length scales of the flow lie within the average inter-particle-distance of tracked particles, spatial derivatives of velocity and of acceleration can be obtained.21In a last step, we interpolate our Lagrangian data onto a grid, which is further discussed in Sec.II C. Figure2(c)depicts the interpolated three-dimensional vorticity field for one time step.

A. Large scale flow statistics

Large scale flow statistics for the characterization of the mean flow field are presented in this section. We used an ensemble average of 400 statistically independent 2D-PTV measurement sets to determine the stream-wise development of mean centerline velocity, Uc, and jet half width, b0.5. Every measurement set consists of 450 images corresponding to 0.15 s of recording. Hence, in total we obtained 60 s of recording for the investigation of the mean velocity profiles from the nozzle exit up to 133 diameters downstream. The spatial resolution of the measurements was 2 mm. The rather coarse resolution is a result of the large field of view that was investigated (200× 200 mm2). However, as stated above, the main goal of these measurements was to get a large scale picture of the flow field and to check, within the achievable experimental accuracy, whether the statistically axisymmetric turbulent jet exhibits self-similarity in the region where we want to perform the local entrainment analysis. Figures3(a)and3(b)show the decay of the mean centerline velocity, Uc, and the development of the velocity half width, b0.5, along the stream-wise jet axis, normalized by the mean exit velocity at the nozzle, Unozzle, and the diameter, d, of the jet pipe. The results clearly show a x−1 decay for the centerline velocity and a linear development of the half width, which is

(6)

−3 −2 −1 0 1 2 3 −0.2 0 0.2 0.4 0.6 0.8 1 1.2 r / b0.5 U / U c x/d = 110 x/d = 100 x/d = 90 x/d = 80 x/d = 70 x/d = 60 x/d = 50 x/d = 40 20 40 60 80 100 120 140 2 3 4 5 6 7 8 9 10 11 12 x/d b0.5 /d b0.5 / d (x) = C2 (x / d − x02) C2 = 0.089912 x02 = 4.3582 20 40 60 80 100 120 140 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2 x/d Uc / U nozzle Uc / Unozzle(x) = C1 / (x / d − x0) C1 = 6.1164 x0 = 3.2365 (a) (b) (c)

FIG. 3. Variation of normalized centerline mean velocity Uc(a) and normalized half width b0.5 (b) as a function of the distance from the nozzle exit x/d. (c) Profiles of the streamwise velocity component U normalized by the centerline velocity

Ucand the velocity half width b0.5.

consistent with the literature, e.g., Refs.22and23. The measured velocity decay constant C1= 6.1 as well as the spreading rate C2= 0.09 lie in the range reported by others, see Refs.24–26. Additionally, we present in Figure3(c)profiles of the streamwise velocity component for different downstream positions normalized by the centerline velocity and the velocity half width. As it is expected for a self-preserving flow, the profiles collapse into a single curve. We therefore conclude that a self-preserving turbulent jet is established in the experiment.

B. Accuracy of 3D-PTV measurements

A brief discussion of the quality of the 3D-PTV measurements is presented in this section. Three kinematic relations are used to assess the accuracy of computed velocity and acceleration gradients. Following L¨uthi et al.,21 the first accuracy test we use is a check on the divergence of the velocity vector. Due to the incompressibility of water, the trace of∂ui

∂xj for each data point should be equal to zero, i.e., −∂ui ∂xi = ∂uj ∂xj +∂uk ∂xk , (1)

with no summation over i, j, and k applied. Plotting joint probability density functions (joint PDFs) of the expression and calculating correlation factors gives a first indication about the accuracy of the performed measurements. However, it is explained in Hoyer et al.20that in a spatially under-resolved flow field, underestimated velocity gradients might still result in an acceptable correlation factor for expression(1). In order to rule out the possibility of underestimated velocity derivatives, we apply additional checks providing a more rigorous assessment of the accuracy of velocity gradients. The second check uses the relation between Lagrangian acceleration, ai = DuDti, and the sum of local acceleration, al,i = ∂u∂ti, and convective acceleration, ac,i = uj∂u∂xij, i.e.,

Dui Dt = ∂ui ∂t + uj ∂ui ∂xj . (2)

The different components of Eq.(2)are derived in different manners. The left hand side of the expression is directly derived from the trajectory of a given particle, whereas the components of the right hand side use information of surrounding particles. Equation(2)is therefore an excellent check for the accuracy of the measurements because each of the quantities is measured independently. The third accuracy test is based on a relation between the divergence of acceleration and the second invariant of the velocity gradient tensor, Q= 1/4(ω2− 2s2), whereω2is enstrophy and 2s2is total strain, i.e.,

(7)

This check is even more difficult to pass because the divergence of a and Q are both obtained from spatial derivatives, but of different quantities.∇ · a is derived from derivatives of the Lagrangian acceleration, whereas Q is obtained from spatial derivatives of velocities.

For the following data processing as well as the analysis of the entrainment velocity, we only use Lagrangian data points that pass a certain quality criterion. The quality criterion is a weighted sum of the three accuracy checks shown above. It consists of a relative divergence error,δu, a relative acceleration error, δa, and a relative divergence of acceleration error,δdiva, see Eqs.(4)–(6). The relative errors can range from zero to one, where values close to zero signify a small error. For our measurements, the weighted relative error, , is defined as 0.2δu+ 0.4δa+ 0.4δdiva, putting a higher emphasis on the more rigorous checks. Only data points that have a relative error of less than 0.3 are accepted for further analysis. The inset of Figure4(a)shows the probability density function of the relative quality error. Ninety percent of the data have a quality error of less than 0.3, so that only a small amount of data is discarded, while at the same time a large improvement is achieved for the correlation factors. Note that the threshold of 0.3 serves only as an upper boundary in order to discard the strongest outliers of the measurements. The average relative error of our measurements is 0.08, as depicted in the inset of Figure4(a),

δu = |∂u1 ∂x1+ ∂u2 ∂x2 + ∂u3 ∂x3| |∂u1 ∂x1| + | ∂u2 ∂x2| + | ∂u3 ∂x3| , (4) δa= 1 3 3  i=1 |ai− al,i− ac,i| |ai| + |al,i| + |ac, i| , (5) δdiva= 2Q+ ∇ · a |1 4ω2| + | 1 2s2| + |∂a 1 ∂x1| + | ∂a2 ∂x2| + | ∂a3 ∂x3| . (6)

Figure4shows the joint PDFs of expressions(1)and(2)(only component x1), and expression

(3). All three joint PDFs show a distribution along the diagonal, hence, no systematic error in the computed velocity and acceleration derivatives can be seen. In theory, all data points should lie on the dashed diagonal, resulting in a correlation factor of 1. The rank correlation factors of expression

(1)are 0.89, 0.87, and 0.90 for the radial, axial, and azimuthal components, respectively. Expression

(2) has correlation factors of 0.90, 0.92, and 0.87 for the respective components. Expression (3)

shows a correlation of 0.91. Note that we used the Spearman rank correlation factor because it is less sensitive to outliers. Although these results imply a random error of about 10% for the determination of velocity and acceleration derivatives, studies performed by L¨uthi et al.21 and Liberzon et al.27 have shown that correlation factors in this order are sufficiently high to perform detailed analysis of quantities related to velocity and acceleration gradients.

For further explanations about the accuracy of the experimental technique and the calculation of spatial derivatives from 3D-PTV data, we refer the reader to previous studies, see Refs. 21

and 11. In Holzner et al.,11 we produced a “synthetic” experimental data set with the use of Lagrangian data from DNS and performed the same data processing as applied in the investigated experiment. Different spatial and temporal resolutions were realized and gradients of velocity as well as acceleration were compared pointwise to the accurate, i.e., well resolved, values. The relative error was in the order of 10%, which is in agreement with the average error obtained for the turbulent jet. Alternative validation procedures to determine accuracies of spatial resolution and differential methods conducted by other groups can be found in, e.g., Refs.28and29.

C. Interpolation of Lagrangian data onto a grid

In preparation for the subsequent enstrophy iso-surface analysis, Lagrangian data of every time step were linearly interpolated onto a regular, equidistant grid with a spacing of 1 mm. In this way the determination of the iso-surface position could be refined by a first order linear approximation instead of just relating it to the random distribution of Lagrangian particle information, which would

(8)

−15 −10 −5 0 5 10 15 −15 −10 −5 0 5 10 15 δu1 δx1[s−1] δu2 δx2 + δu3 δx3 [s 1] 2 2.5 3 3.5 4 4.5 5 5.5 0 0.5 1 0 3 6 Δ PDF (a) −0.5 0 0.5 −0.5 0 0.5 Du1 Dt[m/s2] δu1 δt + ui δu1 δxi [m/s 2] 2 2.5 3 3.5 4 4.5 5 5.5 (b) −200 −100 0 100 200 −200 −150 −100 −50 0 50 100 150 200 ∇ · a[1/s2] 2Q [1 /s 2] 2 2.5 3 3.5 4 4.5 5 5.5 (c) FIG. 4. (a) Joint PDF of∂u1

∂x1 versus ∂u2 ∂x2+

∂u3

∂x3; inset: PDF of quality check of the complete data set. (b) Joint PDF of a1

versus al,1+ ac,1. (c) Joint PDF of divergence of a versus the second invariant Q of the velocity gradient tensor (color code

is in logarithmic scale).

be equivalent to a zero order approximation of the iso-surface position. After interpolating the data onto the grid, signals for every grid point were additionally filtered over time to account for noise introduced by the interpolation process. As a sample for this filtering process, Figure5shows the azimuthal vorticity componentω3 over time for one grid point. Due to the mean shear,ω3 is positive in the mean. The sample illustrates the intermittent nature of the instantaneous flow field. The blue curve illustrates the grid data derived from the interpolated Lagrangian data and the dashed red curve shows the filtered vorticity signal discarding high frequency fluctuations. The applied filter is a Savitzky-Golay filter, which is a moving polynomial fit. See Savitzky and Golay30 for further explanations about the method. We used a cubic polynomial filter with a span of 101 time steps, which corresponds to 2 Kolmogorov time scales. The rank correlation factors of the grid data for the divergence of velocity check shown in Eq.(4) are 0.92, 0.92, and 0.93 for the radial, axial, and azimuthal directions, respectively. We therefore conclude that the interpolation and smoothing procedure did not have a negative influence on the overall data quality, but even slightly improved the accuracy of the divergence of velocity check.

Finally, the filtered field of velocity derivatives is used to determine the enstrophy field for every time step, from which the enstrophy iso-surface at the interfacial region of turbulent jet and irrotational ambient fluid is detected. An example of the interpolated vorticity field for one time step is shown in Figure2(c).

(9)

0 0.5 1 1.5 2 2.5 3 3.5 4 −5 0 5 10 15 20 time [s] ω 3 [1/s] smoothed data interpolated grid data

FIG. 5. Smoothing of grid data for one grid point over time. The azimuthal vorticity component,ω3, is shown. Position of the grid point in the measurement volume is [0.011 mm/0.009 mm/−0.005 mm] in radial, axial, and azimuthal directions, see, e.g., Figure2(c).

III. RESULTS AND DISCUSSION

This section deals with the analysis of an enstrophy iso-surface representing the turbu-lent/nonturbulent interface of the jet. Similar to, e.g., Refs.5,31, and32, the interface is detected in our experiment by using a specific threshold on enstrophy,ω2= ω · ω, where ω is the vorticity vector. Our investigations primarily focus on the local entrainment velocity,vn, defined as the prop-agation velocity of an iso-surface area relative to the fluid, with the positive direction being defined by the enstrophy gradient. Note that this implies that on averagevnof a spreading surface is negative. In particular, we look at the dependency ofvn on the local shape of the iso-surface and identify governing small scale mechanisms of the entrainment process. Different enstrophy thresholds were tested to determine an enstrophy value that reliably detects the turbulence boundary.

For the calculation of the iso-surface corresponding to a specificω2threshold, we used Matlab’s custom iso-surface algorithm, which is based on the marching cubes algorithm by Lorensen and Cline.33Figure6shows a snapshot of the resulting iso-surface for an enstrophy threshold of 4s−2, which is about 5% of the mean enstrophy in the turbulent region. Details about the determination of the threshold will be given subsequently. The turbulent region is indicated by caps color coded with ω2. The surface normal of the iso-surface element points in the direction of the positive enstrophy gradient. In the following text, different areas of the iso-surface will be classified by either concave or convex areas, which is the observed shape looking from the turbulent region towards the iso-surface. Examples of the respective surface types are indicated in Figure6.

A. Determination of the local entrainment velocity

The present section focuses on findings related to the local entrainment velocity,vn. The section is structured in two parts. First, two approaches for calculation ofvnare presented, compared, and are used to assess an appropriate enstrophy threshold for the detection of the interfacial region. Second, small-scale dynamics related to the local entrainment velocity are investigated.

As explained in Refs. 14 and 11, the evolution of the turbulent/non-turbulent interface is governed by advection due to the underlying flow field and a movement relative to the fluid caused by turbulent entrainment. From this it follows that the velocity of an iso-surface element, us, can be written as the sum of fluid velocity, u, and velocity of the area element relative to the fluid, V = vnn,

(10)

FIG. 6. Snapshot of enstrophy iso-surface. The turbulent region is indicated by end-caps color coded with theω2value.

that is, us= u + V , where vn is the local entrainment velocity and n= ∇ω2/|∇ω2| the surface normal. One can derive an expression for the relative velocity by looking at the movement of the enstrophy iso-surface in a Lagrangian frame of reference, i.e., “sitting” on the iso-surface, see, e.g., Refs.34and14. When moving with an enstrophy iso-surface element, the total change ofω2is by definition equal to zero, i.e.,

Dsω2 Dst = δω2 δt + u s j δω2 δxj = δω2 δt + (uj+ Vj)δω 2 δxj = 0, (7)

Inserting V = vnn into Eq.(7)leads to an expression forvn, as shown in Holzner and L¨uthi,17

vn = − Dω2 Dt |∇ω2| = − δω2 δt + ujδω 2 δxj |∇ω2| . (8)

Equation (8) presents a first possibility to determine vn. Another way to derive the local entrainment velocity is to use position information of theω2iso-surface for consecutive time steps. As illustrated in Figure7,vn can also be derived from the displacement of the iso-surface in time. By subtracting the advective component from the overall displacement,vn t is directly obtained. Hence,vncan also be calculated without information on

2

Dt or the gradient of enstrophy,∇ω2, which very often are difficult to determine or are even not at all accessible in experimental measurements. Both methods were applied for the computation of vn in order to investigate two different aspects. First of all, the two approaches should be validated by comparing the results ofvn for the two methods. In particular, we wanted to check whether the determination ofvn based on Eq.(8)is reliable, sincevn was further decomposed into an inviscid and viscous contribution for the subsequent analysis. Second, the two methods were applied to evaluate an appropriate enstrophy threshold for the detection of the turbulence boundary.

Figures8(a)and8(b)show the respective probability density functions of the local entrainment velocity, once computed using Eq.(8)and a second time determined by the “graphical” approach illustrated in Figure7. The data used for the statistics comprises about 107 data points for each method. Negative values ofvn mean that the front propagates from the turbulent into the non-turbulent region. Values are normalized by the average Kolmogorov velocityuk in the turbulent region. Comparing the curves of the two figures, it can be seen that both methods show similar distributions in terms of peak values and standard deviations for different enstrophy thresholds.

(11)

iso-ω2(t) iso-ω2(t+ t)Δ iso-ω2(t)+vΔt n v (x,t)n -u(x,t+ t) tΔ Δ

FIG. 7. Graphical determination of the local entrainment velocity.

Average entrainment values vary about 0.1vn/uk between the two methods. Nevertheless, similar trends are clearly visible and the difference in magnitude is not significant. Both figures display a shift of the average towards zero for increasingω2values. Based on these observations, we conclude that both approaches are capable to determine distribution and order of the local entrainment velocity. Different enstrophy thresholds are plotted in Figures 8(a)and8(b)in order to investigate the influence of enstrophy iso-surface levels on the local entrainment velocity. Due to noise related to PTV measurements and post-processing, small but non-zero ω2 values are also found in the irrotational ambient flow field, hence, an appropriate threshold value has to be defined for a reliable detection of the interfacial region. Chosen thresholds usually lie in the range of 5%–10% of a characteristic value in the turbulent region, such as the mean vorticity value or the product of vorticity magnitude and mean centerline velocity over half-width, see, e.g., Refs.5,11, and35. For our analysis, we use an enstrophy threshold of 4 s−2, which corresponds to 5% of the mean enstrophy value in the turbulent jet. Figure8(a)gives a justification for the choice of the threshold. As can be seen,vn depends on the chosen enstrophy threshold. The average entrainment velocity shifts from about 0.7uk for thresholds 0.4s−2 and 4s−2to zero for a threshold of 400s−2. In our experiment, the mean enstrophy value in the turbulent jet region is about 80s−2. For a threshold much higher than the mean, the PDF ofvnis distributed around zero. There is of course no average entrainment

−7 −6 −5 −4 −3 −2 −1 0 1 2 3 4 5 6 70 0.05 0.1 0.15 0.2 0.25 vn/<uk> PDF th = 0.4 th = 4 th = 40 th = 400 (a) −7 −6 −5 −4 −3 −2 −1 0 1 2 3 4 5 6 70 0.05 0.1 0.15 0.2 0.25 vn/<uk> PDF th = 0.4 th = 4 th = 40 th = 400 (b)

FIG. 8. PDF of the local entrainment velocity for different enstrophy thresholds (th) [s−2] (a) computed by Eq.(8)and (b) using the “graphical” approach.

(12)

−7 −6 −5 −4 −3 −2 −1 0 1 2 3 4 5 6 7 10−6 10−4 10−2 100 102 v n/<uk> PDF v n vninv v n vis

FIG. 9. PDFs of the local entrainment velocityvnand its inviscidvni nvviscousvnviscomponent (semi-logarithmic scale).

for this threshold since the selected iso-surface lies basically inside the turbulent jet and not in the interfacial region. The enstrophy value of 4s−2shows the highest average entrainment velocity and lies above the measurement noise level, which is why it is chosen for the subsequent analysis.

Inspired by the analysis of Refs. 15 and 16, Holzner and L¨uthi17 derived a mathematical description relating the local entrainment velocity to small scale quantities, such as enstrophy production,ωiωjsij, and the viscous term,νωi∇2ωi. They explain that with the use of the enstrophy transport equation, δω2/2 δt + ujδω 2/2 δxj = ωiωjsi j+ νωi∇2ωi, (9)

in combination with Eq.(8), one can further decomposevninto the sum of an inviscid and a viscous contribution, vn= − 2ωiωjsi j |∇ω2| − 2νωi∇2ωi |∇ω2| = v i nv n + vnvis. (10)

With our measurements, we were able to determine velocity and acceleration derivatives. There-fore, the local entrainment velocity as well as its inviscid component could directly be calculated. Since higher order derivatives were not accessible in our PTV measurements, the viscous component vvis

n was determined using expression(10). Figure9shows the PDFs ofvn,vni nv, andvnvis, normalized byuk. It can be seen that all three quantities are skewed towards negative values. A negative entrain-ment velocity implies that the velocity vector is oriented towards the irrotational flow region, i.e., the turbulence boundary advances into the ambient fluid. Mean value as well as standard deviation ofvn lie in the order ofuk. However, the scale separation for the current Reynolds number is not sufficiently large to give a firm determination about the correct scaling parameter of the entrainment velocity. More importantly the figure illustrates thatvnappears to be mainly governed by the viscous term and that the inviscid contribution is small, which is consistent with findings in a zero-mean shear turbulent flow17and was postulated in the work of Corrsin and Kistler.4

B. Entrainment velocity conditioned on surface shapes

This section deals with the dependency of the local entrainment velocity on the geometrical shape of the turbulence boundary. Bisset et al.5 showed that conditional averages of vorticity and velocity at the turbulent/non-turbulent interface vary depending on the local geometry of the interface. Furthermore, they pointed out that small scale mechanisms such as vortex stretching

(13)

-1000-6 -500 0 500 1000 -4 -2 0 2 4 6x 10 5 H K complex curvature region elliptic concave elliptic convex K=H2 saddle saddle concave concave saddle convex convex

FIG. 10. Basic surface types depending on Gaussian and mean curvatures K and H, adapted from Dopazo et al.34

increase or decrease for different shapes of the interface. Taking this aspect into account, we investigate subsequently in which way local entrainment velocity and shape of the enstrophy iso-surface are related. Therefore, vn as well as its components,vi nnv andvnvis, are conditioned on the curvature of the local iso-surface element. In particular, they are conditioned on mean curvature, H= ∇ · n/2 = 12(κ1+ κ2), and Gaussian curvature, K= κ1κ2, where n is the iso-surface normal vector andκ1 as well asκ2are the principal curvatures. As mentioned in the previous chapter, for the detection of the entrainment interface we use an enstrophy threshold of 4 s−2, which corresponds to 5% of the mean enstrophy value in the turbulent jet.

Gaussian and mean curvatures of the triangular mesh of the enstrophy iso-surface were estimated by a paraboloid fitting procedure. Paraboloid fitting is a common approach in the field of computer science to estimate intrinsic geometric properties of a surface derived from a polygonal mesh. Gaussian and mean curvature of every vertex on the triangulated surface are estimated by a least square fitting of an osculating paraboloid to the vertex and its surrounding neighbors. Curvatures of the surface and the paraboloid are considered identical. Further details about the paraboloid fitting procedure are depicted in the Appendix and in Magid et al.36 Figure10shows the different iso-surface geometries in the H-K plane. The iso-iso-surface normal vector is directed in the direction of the increasing enstrophy gradient, i.e., the shapes mentioned in Figure10display the view from the turbulent region towards the interfacial region. The zone K> H2in the H-K plane implies complex curvatures, which do not occur for the investigated iso-surface.

The joint PDF of mean and Gaussian curvatures H and K is displayed on the left of Figure11. Iso-probability contours surround the origin and decrease in magnitude for increasing curvature values. All samples in the joint PDF are below the parabola K = H2, which separates real and complex curvature regions. It can be seen that the joint distribution is skewed towards negative values indicating a slightly higher probability of concave shapes. This feature is also depicted in the PDFs of the single components shown on the right of Figure11. H as well as K are negatively skewed, which reflects the concave shape of the mean jet flow.

Figure12depicts the conditional average of local entrainment velocityvnas well as its inviscid

vi nv

n and viscousvnviscomponents with respect to the surface curvatures H and K. It is shown thatvn changes considerably depending on the local shape of the interface. It tends towards values around zero for large positive mean curvatures (H 0) and decreases gradually to higher negative values in areas of concave shape. As mentioned above, a negativevnmeans that the direction of the velocity is opposite to the normal vector, i.e., the front propagates from the turbulent into the non-turbulent region. From this it follows that in concave areas local entrainment of irrotational fluid into the

(14)

−1000−6 −500 0 500 1000 −4 −2 0 2 4 6x 10 5 H K 4 4.5 5 5.5 6 6.5 7 7.5 8 −1000 0 1000 10−4 10−2 H PDF −5 0 5 x 105 10−8 10−6 K PDF

FIG. 11. (Left) Joint PDF of mean and Gaussian curvatures H and K. (Right) PDFs of single components.

turbulent flow region is higher than for convex shapes. Furthermore, this finding is an indication thatvn enhances the convolution of the iso-surface since concave shapes advance into the ambient fluid with a higher speed than convex shapes. The main parameter for the change ofvn appears to be the mean curvature H. The influence of the Gaussian curvature K indicating saddle points or extreme curvature points seems to be less pronounced. Figure12(b)shows the inviscid velocity componentvi nv

n conditioned on H and K. The characteristic of the distribution differs from the one seen forvn. The inviscid component increases in magnitude for areas of convex shape, i.e., for iso-surface elements with a positive mean curvature. Figure12(c)illustrates the distribution of the viscous componentvnvisin the H-K plane. This quantity displays a similar characteristic asvnhaving large negative values in areas of negative mean curvature. However, in regions of positive H,vvisn assumes even positive values counteracting on average the advancement of the iso-surface into the irrotational flow region. In summary, it appears that depending on the curvature of the iso-surface elementvi nv

n andvnvis are either enhancing or counteracting each other. In concave surface regions

vnis governed by the viscous component related to viscous diffusion,νωi∇2ωi, andvni nv enhances the process, whereas in convex areas the inviscid term related to vortex stretching,ωiωjsij, appears to be the dominant part for the relative advancement of the iso-surface into the irrotational region. Hence, we find that depending on the surface shape, different small-scale mechanisms are dominant for the local entrainment process, i.e., viscous diffusion for concave shapes and vortex stretching for convex shapes.

The components of the local entrainment velocity can be further decomposed in order to determine governing terms for the advancement of the interface into the irrotational flow region related to viscous diffusion and vortex stretching. As explained in detail in Holzner and L¨uthi,17the viscous part ofvn can be split into three parts, resulting in

vvis n = −ν(∇ · n) − ν(δ2ω2/δx2 n) δω2/δx n +2ν∇ωi:∇ωi |∇ω2| , (11)

where xndenotes the coordinate normal to the enstrophy iso-surface. The terms of the equation are not available from our measured data, however, the discussion of the terms together with the results shown in Figure12(c)give some indications about the governing contribution ofvnvis. From the terms on the right-hand side, the last one is always positive and does not contribute to the entrainment of irrotational fluid. The first one is the velocity induced by the mean curvature H= ∇ · n/2. As can be seen in Figure12(c), the conditional average ofvnvis tends towards positive values for positive H values and negative for negative H values. If the mean curvature term was the governing factor,

(15)

FIG. 12. Conditional average of local entrainment velocity,vn, and its components,vi nnvandvvisn , conditioned on mean and

Gaussian curvatures H and K.

the effect should be the opposite. From this we conclude that the first term of Eq. (11)is not the dominant term for the investigated flow. This further leads to the conclusion that the second term of the equation must be the governing contribution for the viscous advancement of the interface, which is similar to findings in a zero-mean shear flow.17 The second term involves the one-dimensional local profile ofω2(x

n) normal to the iso-surface. Therefore, it seems that the shape of the enstrophy profile, in particular, gradient and curvature ofω2(x

n), across the interfacial region has the strongest impact on the viscous component of the local entrainment velocity. A positive value for gradient and curvature of theω2profile will contribute to the local viscous spreading of the iso-surface into the irrotational region. This is the case for the first increase of enstrophy from the irrotational zone to the interfacial region, hence, at the very edge of the interfacial region, known as the “laminar superlayer.”4,17

For a closer analysis of the inviscid component of the local entrainment velocity, we use the relation of enstrophy production with eigenvalues, i, and eigenvectors, λi, of the rate of strain tensor, sij,

(16)

−10001.5 −500 0 500 1000 1.6 1.7 1.8 1.9 2 2.1 2.2 2.3 2.4 H ω i ω j s ij (a) −1000−3 −500 0 500 1000 −2 −1 0 1 2 H Λ i /Λ 2 Λ12 Λ32 (b) −1000 −500 0 500 1000 0.2 0.25 0.3 0.35 0.4 0.45 0.5 H cos 2 (ω ,λi ) cos2(ω,λ1) cos2(ω,λ 2) cos2(ω,λ3) (c)

FIG. 13. ωiωjsij and its components as shown in Eq.(12), conditioned on mean curvature H. Ratios are shown for

eigenvalues i.

Since ω2 is constant for the iso-surface, only

i and cos2(ω, λi) can have an influence on

ωiωjsij. Figure13(a)shows the conditional average of enstrophy production with respect to the mean curvature of the iso-surface. As was already shown in Figure12(b),ωiωjsij increases for convex surface areas, i.e., positive values of H. Figure13(b)shows the ratios of 1and 3with respect to 2. Both ratios show only little dependency on the mean curvature and are quasi constant, hence, it seems that the eigenvalues have only little influence on the change in enstrophy production. From this it follows that the alignment of vorticity and the eigenvectorsλishould be the governing component, which is depicted in Figure13(c). The three alignments change considerably with respect to the mean curvature. Surprisingly, unlike in homogenous turbulence, vorticity is not only strongly aligned with λ2, but also withλ1. In fact, the alignment ofω and λ1 seems to be the driving mechanism for the change ofωiωjsijwith respect to the mean curvature, since its distribution in Figure13(c)reflects the change of enstrophy production illustrated in Figure13(a).

It was mentioned earlier that the local entrainment velocityvnenhances the convolution of the iso-surface since concave shapes advance into the ambient fluid with a higher speed than convex shapes. However, it is known that the main contribution for the convolution of the interface stems from the large scale eddies in the flow,1,8 which are related to the underlying velocity field of the

(17)

FIG. 14. Magnitude of advection velocity|u| at the interface with respect to H and K, normalized by the mean centerline velocity Uc.

flow. Figure14displays the magnitude of the advection velocity at the iso-surface conditioned on Gaussian and mean curvature. Similar to Figure12(a), the velocity magnitude is significantly larger in areas of concave shape in comparison with convex areas, increasing on average the convolution of the iso-surface. The magnitude is at least one order larger than the local entrainment velocity, so that the dominating part for the convolution of the interface is of course the advection velocity. Nevertheless, we can summarize that both components of the propagation velocity of the enstrophy iso-surface, namely, advection and local entrainment velocity, contribute to the convolution of the interface.

According to Turner,37 the global (or average) propagation velocity E

b of the boundary inter-face between turbulent and non-turbulent flow is given by Eb= −d xdbUm, where dbd x is the average spreading rate of the jet and Um the jet velocity averaged over one cross section. From Figure3, we obtain dbd x = 0.09. Using the axial velocity profiles determined from the 2D-PTV measurements, we get Um≈ 0.048 m/s for the streamwise positions x/d = 92 ÷ 96, which is the section where the 3D-PTV measurements were performed. From this it follows that the average propagation velocity in the investigated region is Eb= −0.0043 m/s. Taking into account the information from advection and entrainment velocity of iso-surface elements in radial flow direction, we can also determine the average propagation velocity of the boundary interface from Eb= u + vnnr adi al. In this way we obtain Eb= −0.004 m/s, which is in close agreement to the theoretical value.

IV. CONCLUSION

In this paper, we experimentally investigated the local entrainment velocity in the self-similar region of a turbulent jet flow. Particle tracking velocimetry measurements were performed to deter-mine the position of the fluctuating, instantaneous turbulence boundary and to compute velocity and velocity derivatives at the interface. A quality analysis of the measurements showed that the accuracy of the obtained data is sufficiently high to study small scale mechanisms of the entrainment process, which are related to velocity and acceleration gradients. Two different methods for the determination of local entrainment velocity were compared to check the reliability of both approaches. They were also used to assess an appropriate enstrophy threshold used for the detection of the interfacial region. Decomposing the local entrainment velocity into an inviscid and a viscous part, we found that vn appears to be mainly governed by the viscous term.

(18)

One of the main results of this study is that the local entrainment velocity changes considerably depending on the local shape of the interface. It is illustrated that local entrainment increases for concave surface elements, looking from the turbulent region towards the convoluted boundary. Furthermore, depending on the surface shape, different small-scale mechanisms are dominant for the local entrainment process, i.e., viscous diffusion for concave shapes and vortex stretching for convex shapes. By further decomposing viscous and inviscid terms ofvn, it was illustrated that the viscous advancement of the interface into the non-turbulent region mostly depends on the enstrophy profile normal to the interface. In addition, we showed that the inviscid contribution is mainly governed by the alignment of vorticity with the eigenvectors of the rate of strain tensor. Finally, it was found that the convolution of the instantaneous interface is driven by the advection of the underlying fluid together with a contribution of the local entrainment velocity, with advection velocity being the governing part.

ACKNOWLEDGMENTS

We gratefully acknowledge the support of this work by the Georg Fischer Fund under Grant No. ETH-14 10-2.

APPENDIX: PARABOLOID FITTING METHOD

In the following, we briefly explain the paraboloid fitting method. The description is mainly derived from explanations by Magid et al.36 Additional information can be found in, e.g., Refs.38and39.

Paraboloid fitting is a common approach in the field of computer science to estimate Gaussian and mean curvature for triangular meshes. The algorithm approximates a small area of the triangulated mesh by an oscillating paraboloid. The principal curvatures of the surface area are estimated at every vertex by a least square fitting of the osculating paraboloid to a vertexv and its surrounding neighbors.36 Curvatures of the triangulated surface element and the paraboloid are considered identical.

The first step in the procedure is to determine the immediate neighbors of vertex v. The neighboring verticesvi are found by looking for edges containing vertexv, i.e., searching for edges

ei = vvi. Principal curvatures are estimated for the surface element comprised of triangles Tiformed by vertexv and its neighboring vertices vi. Figure15shows the notations used for vertices, edges,

Nv

v

1

v

2

v

3

v

4

v

5

v

0

e

0

v

e

1

e

2

e

3

e

4

e

5

T

0

T

1

T

2

T

3

T

4

T

5

(19)

and triangles. Subsequently, the normal vector Nvof the specific surface element is estimated. Let Nibe the unit normal of the triangle Ti, with

− →N i = (vi− v) × (v(i+1)modn− v) (vi− v) × (v(i+1)modn− v) . (A1)

Then, Nvcan be estimated as an average of normals −→Ni,

Nv= 1 n n−1  i=0 − →N i; −→Nv= Nv Nv. (A2)

The third step is to transform vertexv along with its neighbors viinto a new coordinate system; withv being the origin and −Nvcoalescing with the z axis. Then, the osculating paraboloid can be computed by

z= ax2+ bxy + cy2. (A3)

The coefficients a, b, and c are found by solving a least squares fit for the neighboring vertices vi. Finally, Gaussian and mean curvatures of the surface element are determined as

K = 4ac − b2, H = a + c. (A4)

1A. A. Townsend, The Structure of Turbulent Shear Flow, 2nd ed. (Cambridge University Press, 1976).

2J. C. R. Hunt, I. Eames, and J. Westerweel, “Mechanics of inhomogeneous turbulence and interfacial layers,”J. Fluid

Mech.554, 449–519 (2006).

3C. B. DaSilva, “The behaviour of subgrid-scale models near the turbulent/nonturbulent interface in jets,”Phys. Fluids21, 081702 (2009).

4S. Corrsin and A. Kistler, “The free-stream boundaries of turbulent flows,” Technical Reports TN-3133 and TR-1244 (NACA, 1954), pp. 1033–1064.

5D. K. Bisset, J. C. R. Hunt, and M. M. Rogers, “The turbulent/non-turbulent interface bounding a far wake,”J. Fluid Mech. 451, 383–410 (2002).

6W. Dahm and P. Dimotakis, “Mixing at large schmidt number in the self-similar far field of turbulent jets,”J. Fluid Mech. 217, 299–330 (1990).

7J. Westerweel, C. Fukushima, J. Pedersen, and J. Hunt, “Momentum and scalar transport at the turbulent/non-turbulent interface of a jet,”J. Fluid Mech.631, 199–230 (2009).

8D. Coles, “Interfaces and intermittency in turbulent shear flow,” Mec. Turbul., C.N.R.S. 108, 229–248 (1962); in A. Favre, editor, “M´ecanique de la turbulence,” Proceedings of the Colloques Internationaux du CNRS, 108, Marseille, 28 Aug.– 2 Sept. 1961 (CNRS, Paris, 1962), pp. 229–248.

9D. J. Tritton, Physical Fluid Dynamics, 2nd ed. (Clarendon, Oxford, 1988).

10A. Tsinober, An Informal Conceptual Introduction to Turbulence, 2nd ed. (Springer, Berlin, 2009).

11M. Holzner, B. L¨uthi, A. Tsinober, and W. Kinzelbach, “Acceleration, pressure and related quantities in the proximity of the turbulent/non-turbulent interface,”J. Fluid Mech.639, 153–165 (2009).

12A. Liberzon, M. Holzner, B. L¨uthi, M. Guala, and W. Kinzelbach, “On turbulent entrainment and dissipation in dilute polymer solutions,”Phys. Fluids21, 1–6 (2009).

13K. R. Sreenivasan, R. Ramshankar, and C. Meneveau, “Mixing, entrainment and fractal dimensions of surfaces in turbulent flows,”Proc. R. Soc. London., Ser. A421(1860), 79–108 (1989).

14S. B. Pope, “The evolution of surfaces in turbulence,”Int. J. Eng. Sci.26(5), 445–469 (1988).

15C. Dopazo, “Iso-scalar surfaces, mixing and reaction in turbulent flows,”C. R. Mec.334, 483–492 (2006).

16C. Dopazo, J. Martin, and J. Hierro, “On conditioned averages for intermittent turbulent flows,”J. Fluid Mech.81(3), 483–492 (1977).

17M. Holzner and B. L¨uthi, “Laminar superlayer at the turbulence boundary,”Phys. Rev. Lett.106(13), 134503 (2011). 18M. V. Zagarola and A. J. Smits, “Mean-flow scaling of turbulent pipe flow,”J. Fluid Mech.373, 33–79 (1998).

19P. Billant, J. M. Chomaz, and P. Huerre, “Experimental study of vortex breakdown in swirling jets,” J. Fluid Mech. 376(183), 183–219 (1998).

20K. Hoyer, M. Holzner, B. L¨uthi, M. Guala, A. Liberzon, and W. Kinzelbach, “3D scanning particle tracking velocimetry,”

Exp. Fluids39(5), 923–934 (2005).

21B. L¨uthi, A. Tsinober, and W. Kinzelbach, “Lagrangian measurement of vorticity dynamics in turbulent flow,”J. Fluid

Mech.528, 87–118 (2005).

22J. O. Hinze, Turbulence, 2nd ed. (McGraw-Hill, New York, 1975). 23S. B. Pope, Turbulent Flows (Cambridge University Press, 2000).

(20)

25N. J. C. Hussein, S. P. Capp, and W. K. George, “Velocity measurements in a high-reynolds-number, momentum-conserving, axisymmetric, turbulent jet,”J. Fluid Mech.258, 31–75 (1994).

26G. Xu and R. A. Antonia, “Effect of different initial conditions on a turbulent round free jet,”Exp. Fluids33, 677–683 (2002).

27A. Liberzon, B. L¨uthi, M. Holzner, A. Tsinober, S. Ott, J. Berg, and J. Mann, “On the structure of acceleration in turbulence,”

Physica D241, 208–215 (2012).

28G. Gulitski, M. Kholmyansky, W. Kinzelbach, B. L¨uthi, A. Tsinober, and S. Yorish, “Velocity and temperature derivatives in high Reynolds number turbulent flows in the atmospheric surface layer. Part II. Accelerations and related matters,”

J. Fluid Mech.589, 83–102 (2007).

29S. I. Chernyshenko, G. M. Di Cicca, A. Iollo, A. V. Smirnov, N. D. Sandham, and Z. W. Hu, “Analysis of data on the relation between eddies and streaky structures in turbulent flows using the placebo method,”Fluid Dyn.41(5), 772–783

(2006).

30A. Savitzky and M. J. E. Golay, “Smoothing and differentiation of data by simplified least squares procedures,”Anal.

Chem.36(8), 1627–1639 (1964).

31J. Westerweel, “Mechanics of the turbulent-nonturbulent interface of a jet,”Phys. Rev. Lett.95, 174501 (2005). 32M. Holzner, A. Liberzon, N. Nikitin, W. Kinzelbach, and A. Tsinober, “Small scale aspects of flows in proximity of the

turbulent/non-turbulent interface,”Phys. Fluids19, 071702 (2007).

33W. E. Lorensen and H. E. Cline, “Marching cubes: A high resolution 3d surface construction algorithm,”Comput. Graphics 21(4), 163–169 (1987).

34C. Dopazo, J. Martin, and J. Hierro, “Local geometry of isoscalar surfaces,”Phys. Rev. E76(5), 056316 (2007). 35R. K. Anand, B. J. Boersma, and A. Agrawal, “Detection of turbulent/non-turbulent interface for an axisymmetric turbulent

jet: Evaluation of known criteria and proposal of a new criterion,”Exp. Fluids47, 995–1007 (2009).

36E. Magid, O. Soldea, and E. Rivlin, “A comparison of gaussian and mean curvature estimation methods on triangular meshes of range image data,”Comput. Vis. Image Underst.107, 139–159 (2007).

37J. S. Turner, “Turbulent entrainment: The development of the entrainment assumption, and its application to geophysical flows,”J. Fluid Mech.173, 431–471 (1986).

38B. Hamann, “Curvature approximation for triangulated surfaces,”Comput. Suppl.8, 139–153 (1993); in G. Farin, H. Hagen, and H. Noltemeier, eds., “Geometric Modelling,” Computing Suppl. 8 (Springer-Verlag, UK, 1993), pp. 139–153. 39E. Stokely and S. Y. Wu, “Surface parametrization and curvature measurement of arbitrary 3-D objects: five practical

Referenties

GERELATEERDE DOCUMENTEN

In dit gebiedsplan moeten aanvullende landschappelijke, ruimtelijke en energetische maatregelen en voorzieningen worden geïntegreerd, die proportioneel zijn in verhouding tot de

De toenmalige registratiehouder van cysteamine met vertraagde afgifte (Procysbi®) heeft de aanvraag voor opname in het GVS teruggetrokken na het uitbrengen van ons advies, vanwege

There is no minimum of 33% caused by words with only one or two forms (like we see in the.. lexicostatistical results); there is no conceivable maximum like you’d have in a

In this study, the operational water footprints of the products included water incorporated into the product as an ingredient, water consumed during the production process, and

Het zou mogelijk kunnen zijn dat personen die op de voormeting een hoge capaciteit voor negatieve toekomstige ingebeelde gebeurtenissen lieten zien mede hierdoor meer angst konden

Difaqane het dit in die meeste gevalle gegaan oor oorlewing - met ander woorde die verkryging van voedsel waar 'n gevestigde starn of sibbe uit sy woonplek

Complexities regarding the user characteristics of elderly, conversational mechanisms (i.e. the processes of turn­taking, grounding and conver­ sational repairs), dialogue structure

move through the membrane the total energy requirement of the per- vaporation-distillation scheme would be 15.7 MJ kg −1 butanol, which is still less than half (44%) of the