Dynamics of quasi-two-dimensional turbulent jets Julien Re´my Dominique Ge´rard Landel Churchill College A dissertation submitted for the degree of Doctor of Philosophy Department of Applied Mathematics and Theoretical Physics and BP Institute The University of Cambridge May 2012 Dynamics of quasi-two-dimensional turbulent jets Julien Re´my Dominique Ge´rard Landel Abstract The study of quasi-two-dimensional turbulent jets is relevant to chemical reactors, the coking process in oil refinement, as well as rivers flowing into lakes or oceans. In the event of a spillage of pollutants into a river, it is critical to understand how these agents disperse with the flow in order to assess damage to the environment. For such flows, characteristic streamwise and cross-stream dimensions can be much larger than the fluid-layer thickness, and so the flow develops in a confined environment. When the distance away from the discharge location is larger than ten times the fluid- layer thickness, the flow is referred to as a quasi-two-dimensional jet. From experimental observations using dyed jets and particle image velocimetry, we find that the structure of a quasi-two-dimensional jet consists of a high-speed mean- dering core with large counter-rotating eddies developing on alternate sides of the core. The core and eddy structure is self-similar with distance from the discharge location. The Gaussianity of the cross-stream distribution of the time-averaged velocity is due, in part, to the sinuous instability of the core. To understand the transport and dispersion properties of quasi-two-dimensional jets we use a time-dependent advection–diffusion equation, with a mixing length hypothesis accounting for the turbulent eddy diffusivity. The model is supported by experimental releases of dye in jets or numerical releases of virtual passive tracers in experimentally- measured jet velocity fields. We consider the statistical properties of this flow by releasing and then tracking large clusters of virtual particles in the jet velocity field. The probability distributions of two- point properties (such as the distance between two particles) reveal large streamwise dispersion. Owing to this streamwise dispersive effect, a significant amount of tracers can be transported faster than the speed predicted by a simple advection model. Using potential theory, we determine the flow induced by a quasi-two-dimensional jet confined in a rectangular domain. The streamlines of the induced flow predicted by the theory agree with experimental measurements away from the jet boundary. Finally, we investigate the case of a quasi-two-dimensional particle-laden jet. De- pending on the bulk concentration of dense particles, we identify different flow regimes. At low concentrations, the jet features the same core and eddy structure observed with- out the particles, and thus quasi-two-dimensional jet theory can apply to some extent. At larger concentrations, we observe an oscillating instability of the particle-laden jet. Preface This thesis, which is submitted for the degree of Doctor of Philosophy at Churchill College, University of Cambridge, describes work carried out from January 2009 to May 2012 in the Department of Applied Mathematics and Theoretical Physics and the BP Institute, University of Cambridge. This dissertation is the result of my own work and includes nothing which is the outcome of work done in collaboration, except where specifically indicated in the text. No part of this work has been, or is being submitted for any other qualification at this or any other university. Acknowledgements First of all, I would like to express my deep appreciation to my supervisor Dr Colm Caulfield and my advisor Prof. Andy Woods for giving me the opportunity to pursue my PhD research on this fascinating project, here at the University of Cambridge. I am particularly grateful for the constant support and useful advice I have received from Dr Colm Caulfield. He has taught me the rigour of mathematical modelling, and how it can be applied to physical problems. I wish to thank Prof. Andy Woods for all the fruitful and enthusiastic discussions. His unique way of simplifying any complex problem into its essential components will continue to serve me as a guide throughout my professional career. I am very grateful to Dr Stuart Dalziel for his invaluable help in experimental and technical matters, and in the use of DigiFlow. He has always made himself available at any time during my PhD. His knowledge and expertise in experimental Fluid Dynamics is a wonderful source of inspiration. I would like to extend my gratitude to my friend Samuel Rabin for helping me resolve numerous mathematical problems and equations throughout my PhD. I would like to thank my friends Megan Davies Wykes, Alan Jamieson, Antoine Julia, Hugh Lund and Dr Ben Maurer for proofreading parts of this document. I am also grateful to all my colleagues and friends at the G.K. Batchelor lab and at the BPI lab for useful chats as well as many memorable events and dinners together. Thanks are due to the technician of the BPI, Andrew Pluck, who built my experimental apparatus. I am also grateful to the DAMTP technicians, David Page-Croft, Colin Hitch, John Milton and Neil Price, for their technical support. I would like to thank BP, the Engineering and Physical Sciences Research Coun- cil, the Cambridge Philosophical Society, DAMTP, the BPI and Churchill College for their financial support. I would like to express my love and gratitude to my parents, sister, brothers and other family members. This work would not have been possible without their continuous encouragement, support and love. Finally, I am eternally grateful to the Creator of all things. I am simply fasci- nated by the wonders of His work. Contents 1 Overview 1 2 Meandering due to large eddies and the statistically self-similar dynamics of quasi-two-dimensional jets 5 2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 2.2 Experimental procedure . . . . . . . . . . . . . . . . . . . . . . . 7 2.2.1 Dye tracking experiments . . . . . . . . . . . . . . . . . . 8 2.2.2 Particle image velocimetry experiments . . . . . . . . . . . 9 2.3 Qualitative observations . . . . . . . . . . . . . . . . . . . . . . . 10 2.4 Time-averaged mean flow field . . . . . . . . . . . . . . . . . . . . 12 2.5 Quantitative analysis of the time-dependent core and eddy structure 20 2.5.1 Time-dependent eddy structure . . . . . . . . . . . . . . . 20 2.5.2 Time-dependent core structure . . . . . . . . . . . . . . . 26 2.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 3 Advection–diffusion model for the streamwise transport, disper- sion and mixing in quasi-two-dimensional jets 33 3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33 3.2 Turbulent model hypothesis . . . . . . . . . . . . . . . . . . . . . 37 i Contents 3.3 Mathematical model . . . . . . . . . . . . . . . . . . . . . . . . . 41 3.3.1 Similarity transformation . . . . . . . . . . . . . . . . . . . 42 3.3.2 Constant-flux release: concentration . . . . . . . . . . . . . 43 3.3.3 Constant-flux release: concentration flux . . . . . . . . . . 49 3.3.4 Finite-volume release: instantaneous release fundamental solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55 3.3.5 Finite-volume release: time-dependent release general solu- tion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61 4 Streamwise transport, dispersion and mixing in quasi-two-dimen- sional jets: experimental results 67 4.1 Experimental procedure . . . . . . . . . . . . . . . . . . . . . . . 67 4.1.1 Constant-flux releases of dye . . . . . . . . . . . . . . . . . 68 4.1.2 Instantaneous finite-volume releases of clusters of virtual particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70 4.1.3 Finite-volume releases of dye . . . . . . . . . . . . . . . . . 71 4.2 Experimental results . . . . . . . . . . . . . . . . . . . . . . . . . 73 4.2.1 Qualitative assessment . . . . . . . . . . . . . . . . . . . . 74 4.2.2 Constant-flux releases of dye . . . . . . . . . . . . . . . . . 80 4.2.3 Instantaneous finite-volume releases of clusters of virtual particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88 4.2.4 Finite-volume releases of dye . . . . . . . . . . . . . . . . . 93 4.3 Statistical significance of the experimental results . . . . . . . . . 99 4.3.1 Constant-flux release of dye . . . . . . . . . . . . . . . . . 100 4.3.2 Instantaneous finite-volume release of virtual particles . . . 102 4.3.3 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . 104 4.4 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106 5 Two-point statistics for turbulent relative dispersion in quasi- two-dimensional jets 111 5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111 5.2 Mathematical definitions of two-point probability distributions . . 115 5.2.1 Continuous formulation . . . . . . . . . . . . . . . . . . . . 115 5.2.2 Discrete formulation . . . . . . . . . . . . . . . . . . . . . 117 ii Contents 5.3 Test studies in diverging velocity fields . . . . . . . . . . . . . . . 117 5.3.1 Circular domain in an axisymmetric diverging velocity field 118 5.3.2 Elliptical domain in a non-axisymmetric diverging velocity field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123 5.3.3 Square cluster of virtual particles in a diffusive velocity field 126 5.3.4 Conclusion of the test studies . . . . . . . . . . . . . . . . 129 5.4 Analysis of the virtual particles in the jet structures . . . . . . . . 131 5.4.1 Virtual particles: time-dependent versus time-averaged ve- locity fields . . . . . . . . . . . . . . . . . . . . . . . . . . 131 5.4.2 Two-point statistics: time-dependent versus time-averaged velocity fields . . . . . . . . . . . . . . . . . . . . . . . . . 132 5.5 Discussion and Conclusion . . . . . . . . . . . . . . . . . . . . . . 139 6 Flow induced by a quasi-two-dimensional jet in a confined rect- angular domain 145 6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145 6.2 Potential flow model . . . . . . . . . . . . . . . . . . . . . . . . . 149 6.2.1 Description of the entrainment problem . . . . . . . . . . . 149 6.2.2 Decomposition of the problem . . . . . . . . . . . . . . . . 151 6.2.3 Solution to the uniform problem ϕ˜u . . . . . . . . . . . . . 152 6.2.4 Solution to the perturbation problem ϕ˜p . . . . . . . . . . 155 6.2.5 Solution to the entrainment problem . . . . . . . . . . . . 161 6.3 Experimental results . . . . . . . . . . . . . . . . . . . . . . . . . 165 6.3.1 Experimental procedure . . . . . . . . . . . . . . . . . . . 165 6.3.2 Qualitative observations . . . . . . . . . . . . . . . . . . . 166 6.3.3 Quantitative results . . . . . . . . . . . . . . . . . . . . . . 168 6.4 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174 7 Dynamics of particle-laden jets in quasi-two-dimensional envi- ronments 179 7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179 7.2 Experimental procedure . . . . . . . . . . . . . . . . . . . . . . . 181 7.3 Phenomenological description . . . . . . . . . . . . . . . . . . . . 182 7.3.1 Regime diagram . . . . . . . . . . . . . . . . . . . . . . . . 182 7.3.2 Regime I: fluidized bed . . . . . . . . . . . . . . . . . . . . 184 iii Contents 7.3.3 Regime II: oscillatory flow . . . . . . . . . . . . . . . . . . 184 7.3.4 Regime III: core and eddy flow . . . . . . . . . . . . . . . 187 7.4 Core and eddy flow model . . . . . . . . . . . . . . . . . . . . . . 187 7.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190 7.5.1 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . 190 7.5.2 Future work . . . . . . . . . . . . . . . . . . . . . . . . . . 192 8 Conclusion and future work 197 8.1 Review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197 8.2 Future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201 A Advection–diffusion model for quasi-two-dimensional jets 203 A.1 Proof of equation (3.89) . . . . . . . . . . . . . . . . . . . . . . . 203 B Two-point statistics in circular distributions 205 B.1 Conditional probability for the x-distance between two points in a disc . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205 B.2 Value at the origin for the p.d.f. of the lateral distance between two points in a disc . . . . . . . . . . . . . . . . . . . . . . . . . . 207 B.3 Value at 2R(t) for the p.d.f. of the lateral distance between two points in a disc . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208 B.4 P.d.f of the x-distance between two points in a square domain . . 208 B.5 Conditional probability for the Euclidean distance between two points in a disc . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209 B.6 Conditional probability for the ratio of the lateral distance to the streamwise distance between two points in a disc . . . . . . . . . . 211 Bibliography 212 iv List of Figures 1.1 Photographs of quasi-two-dimensional jets in the laboratory and in nature. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3 2.1 Schematic diagram of the experimental apparatus. . . . . . . . . . 8 2.2 Sequence of grey-scale pictures of a dyed jet. . . . . . . . . . . . . 12 2.3 The eddy and core structure of quasi-two-dimensional jets. . . . . 13 2.4 Average dye edge and average velocity spread rate. . . . . . . . . 16 2.5 Maximum time-averaged streamwise velocity versus height. . . . . 17 2.6 Time-averaged momentum flux versus height. . . . . . . . . . . . 18 2.7 Time-averaged streamwise velocity at various heights. . . . . . . . 19 2.8 Identification of the eddy and core structures. . . . . . . . . . . . 22 2.9 Eddy locations in quasi-two-dimensional jets. . . . . . . . . . . . . 23 2.10 Eddy z-coordinate versus time. . . . . . . . . . . . . . . . . . . . 24 2.11 Eddy frequency and Strouhal number versus height. . . . . . . . . 25 2.12 Time-averaged mean core structure. . . . . . . . . . . . . . . . . . 28 2.13 Instantaneous streamwise velocity at different heights. . . . . . . . 29 2.14 Mean streamwise velocity and mean core–eddy structure. . . . . . 32 3.1 Concentration similarity solution for the constant-flux release. . . 47 v List of Figures 3.2 Centroid and standard deviation of the concentration solution for the constant-flux release. . . . . . . . . . . . . . . . . . . . . . . . 48 3.3 Volume of tracers ahead of the advective front for the constant-flux release. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50 3.4 Concentration-flux similarity solution for the constant-flux release. 52 3.5 Centroid and standard deviation of the concentration-flux solution for the constant-flux release. . . . . . . . . . . . . . . . . . . . . . 54 3.6 Concentration flux of tracers ahead of the advective front for the constant-flux release. . . . . . . . . . . . . . . . . . . . . . . . . . 55 3.7 Similarity solution for the instantaneous finite-volume release. . . 58 3.8 Centroid and standard deviation of the concentration solution for the instantaneous finite-volume release. . . . . . . . . . . . . . . . 60 3.9 Volume of tracers ahead of the advective front for the instantaneous finite-volume release. . . . . . . . . . . . . . . . . . . . . . . . . . 62 3.10 Deviation between the general solution and the fundamental solu- tion for the finite-volume release. . . . . . . . . . . . . . . . . . . 64 4.1 Schematic diagram of the experimental apparatus. . . . . . . . . . 68 4.2 Core and eddy structures in quasi-two-dimensional jets revealed by dye and passive tracers. . . . . . . . . . . . . . . . . . . . . . . . 75 4.3 Typical trajectories of single virtual particles in an eddy, in the core and at the interface between the two structures. . . . . . . . 78 4.4 Trajectories of virtual particle clusters seeded in an eddy, in the core and at the interface between the two structures. . . . . . . . 79 4.5 Constant-flux and finite-time finite-volume dye releases in quasi- two-dimensional jets. . . . . . . . . . . . . . . . . . . . . . . . . . 81 4.6 Time evolution of the experimental dye concentration for constant- flux releases. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84 4.7 Comparison between experimental data and theory for the concen- tration of constant-flux releases. . . . . . . . . . . . . . . . . . . . 85 4.8 Comparison between experimental data and theory for the concen- tration flux of constant-flux releases. . . . . . . . . . . . . . . . . 88 4.9 Instantaneous finite-volume releases of virtual particles. . . . . . . 90 4.10 Comparison between experimental data and theory for the concen- tration of instantaneous finite-volume releases. . . . . . . . . . . . 92 vi List of Figures 4.11 Experimental dye data for finite-time finite-volume releases. . . . 95 4.12 Comparison between experimental data and theory for the concen- tration of finite-time finite-volume releases. . . . . . . . . . . . . . 98 4.13 Evolution in time of the theoretical prediction of the concentration for a finite-time finite-volume release. . . . . . . . . . . . . . . . . 99 4.14 Probability density function and critical probability of the experi- mental dye concentration in the case of constant-flux releases. . . 101 4.15 Probability density function and critical probability of the concen- tration of virtual particles in the case of instantaneous finite-volume releases. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104 5.1 Probability distributions of two-point properties for a uniformly distributed disc. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121 5.2 Time evolution of a uniform distribution of particles in an ellipse. 125 5.3 Probability density functions of two-point properties for a uni- formly distributed ellipse. . . . . . . . . . . . . . . . . . . . . . . 127 5.4 Time evolution of an initially square distribution of particles fol- lowing random walks. . . . . . . . . . . . . . . . . . . . . . . . . . 128 5.5 Probability density functions of two-point properties for a diffusing cluster of virtual particles. . . . . . . . . . . . . . . . . . . . . . . 129 5.6 Time evolution of three virtual-particle clusters seeded in an eddy, in the core and at the interface between the two structures in both a time-averaged and a time-dependent velocity field. . . . . . . . . 133 5.7 Probability distributions of two-point properties for a cluster of virtual particles seeded at the location of an eddy for the time- dependent and the time-averaged velocity fields. . . . . . . . . . . 136 5.8 Probability distributions of two-point properties for a cluster of virtual particles seeded at the interface between an eddy and the core for the time-dependent and the time-averaged velocity fields. 138 5.9 Probability distributions of two-point properties for a cluster of virtual particles seeded in the core for the time-dependent and the time-averaged velocity fields. . . . . . . . . . . . . . . . . . . . . . 140 6.1 Description of the entrainment problem. . . . . . . . . . . . . . . 152 vii List of Figures 6.2 Decomposition of the entrainment problem into a uniform problem and a perturbation problem. . . . . . . . . . . . . . . . . . . . . . 153 6.3 Potential and stream function of the uniform problem. . . . . . . 154 6.4 Velocity field of the uniform problem. . . . . . . . . . . . . . . . . 154 6.5 Potential and stream function of the perturbation problem. . . . . 157 6.6 Velocity field of the perturbation problem. . . . . . . . . . . . . . 158 6.7 Distribution of the flux at the jet boundary for the perturbation problem. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160 6.8 Convergence of the numerical truncated series for the flux at the jet boundary. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162 6.9 Potential and stream function of the flow induced by a quasi-two- dimensional jet modelled as a varying line sink. . . . . . . . . . . 164 6.10 Velocity field of the flow induced by a quasi-two-dimensional jet modelled as a varying line sink. . . . . . . . . . . . . . . . . . . . 165 6.11 Schematic diagram of the experimental apparatus. . . . . . . . . . 166 6.12 Pictures of the flow induced by quasi-two-dimensional jets. . . . . 169 6.13 Experimental and theoretical distributions of the streamlines of the flow induced by quasi-two-dimensional jets. . . . . . . . . . . . . . 170 6.14 Experimental and theoretical distributions of the velocity field in- duced by quasi-two-dimensional jets. . . . . . . . . . . . . . . . . 171 6.15 Experimental and theoretical distributions of the volume flux and momentum flux of the flow induced by quasi-two-dimensional jets. 174 7.1 Experimental apparatus to study quasi-two-dimensional particle- laden jets (Q2DPL jets). . . . . . . . . . . . . . . . . . . . . . . . 182 7.2 Regime diagram. Three phenomenological regimes are observed during the Q2DPL jet experiment. . . . . . . . . . . . . . . . . . 183 7.3 Illustration of the fluidization regime. . . . . . . . . . . . . . . . . 185 7.4 Illustration of the oscillatory flow regime. . . . . . . . . . . . . . . 186 7.5 Illustration of a large vortical structure in the oscillatory flow regime.186 7.6 Illustration of the core and eddy flow regime. . . . . . . . . . . . . 188 7.7 Particle maximum height and bed thickness versus flow rate. . . . 191 viii Chapter 1 Overview In turbulent jets, fluid is driven by momentum from an orifice into an environ- ment filled with similar fluid. The complexity of this flow, which has been studied for more than 80 years (see e.g. List, 1982, for a detailed review), resides in its turbulent nature. Turbulence develops due to a shear instability at the boundary between the jet fluid and the ambient fluid. The transition of the flow from lami- nar to turbulent typically occurs at a Reynolds number Re = bU/ν (where b is the characteristic width of the jet, U is the jet characteristic streamwise velocity and ν is the kinematic viscosity of the fluid) of the order of 3000. From the equations of motion, the momentum flux is approximately conserved (see e.g. Kotsovinos, 1978, for a discussion on the conservation of momentum in turbulent jets), while its mean kinetic energy is dissipated by turbulence. Momentum spreads laterally due to entrainment of ambient fluid in the jet. The entrainment process is gov- erned by the large-scale turbulent structures in the flow and is self-similar in the streamwise direction. 1 1 Overview The capacity of turbulent jets to entrain ambient fluid and mix it efficiently with jet fluid accounts in large part for the attention this flow has received in both the scientific community and the industrial world. Also, behind the appar- ent simplicity of the jet mean motion lies the fascination for the elusive underlying physics of turbulence. Whether for their dilution properties, their efficient mixing properties or the thrust they can provide, jets have been used in various indus- trial applications, such as waste water disposal (Yannopoulos, 2006), chemical reactors (Jirka & Harleman, 1979), or as a means of propulsion (Stanley, Sarkar & Mellado, 2002). In geophysical flows, turbulent jets are, for instance, relevant to the study of explosive volcanic eruptions, where a mixture of gas, fluid lava and solid particles is initially driven by momentum out of the crater (Woods & Caulfield, 1992). In this study we are interested in a particular type of turbulent jet called a quasi- two-dimensional steady turbulent jet (which we refer to, hereafter, as a quasi-two- dimensional jet). Giger, Dracos & Jirka (1991) gave the first description of quasi- two-dimensional jets (earlier studies of bounded plane jets include Foss & Jones, 1968; Holdeman & Foss, 1975, who focused on the near field of the flow). They observed that, in the far field of a plane turbulent jet confined between two close boundaries separated by a gap widthW , the flow develops into a meandering core with large counter-rotating eddies growing on alternate sides of the core. A qua- si-two-dimensional jet designates the region of the flow (starting from z ≈ 10W , with z the streamwise distance from the source) where the meandering core and the large growing eddies appear. The sinuous instability of the jet is due to lateral transverse shear (Jirka & Uijttewaal, 2004). Dracos, Giger & Jirka (1992) found an inverse cascade of quasi-two-dimensional turbulence, which affects not only the structure of the flow but also transport, dispersion and mixing properties. The aim of this thesis is to investigate experimentally and theoretically the transport, dispersion and mixing properties of quasi-two-dimensional jets. In figure 1.1(a), we show a picture of a typical quasi-two-dimensional dyed jet (Re ≈ 4000) produced in our experimental apparatus (whose gap width is W = 1 cm). As we can see, for z > 10 cm, the jet meanders and large eddies form on alternate sides of the core. The same core and eddy structure has been observed in geophysical flows, such as rivers discharging into lakes or oceans. At the discharge location, the depth of a river is often much smaller than the other 2 1 Overview (a) 10 cm (b) (c) Figure 1.1: Meandering quasi-two-dimensional jets in the laboratory and in nature: (a) grey-scale picture of a dyed steady quasi-two-dimensional turbulent jet (Re ≈ 4000) rising in our experimental apparatus; (b) photograph of a channel (Re ≈ 107) discharg- ing from the Lower Mississippi River (near Baton Rouge, LA, USA) into an oxbow lake, Image Source: 1998 US Geological Survey Digital Ortho-Quarter Quadrangle; (c) pho- tograph of a river (Re ≈ 107) flowing into Balaton Lake, Hungary (Jirka & Uijttewaal, 2004). In (b) and (c), the meanders are made visible by the sediment transported by the flow. two characteristic dimensions of the environment in the horizontal plane. Thus, as depicted in figures 1.1(b) and 1.1(c), a river flow can develop into a quasi-two- dimensional jet flow. In figures 1.1(b) and 1.1(c), the core and eddy structures, displayed by the two rivers discharging into lakes (Re ≈ 107), are revealed by the sediment transported by the flow. The study of river flows is relevant to coastal engineering problems, such as sediment transport and coastal erosion (Joshi & Taylor, 1983), as well as environ- mental pollution. In the event of a spillage of pollutants in rivers, the prediction and monitoring of the transport and dispersion of the pollutants is crucial. Ac- curate models of the flow, tested against experimental evidence, are therefore needed to control this type of environmental pollution. The main objective of this thesis is to address these issues. We compare the flow of a river discharging into a large basin with a (laboratory) 3 1 Overview quasi-two-dimensional jet. In Chapter 2, we analyze and model the time-averaged velocity field of quasi-two-dimensional jets. We present a quantitative description of the characteristic core and eddy structure. We discuss the implications of this core and eddy structure on the velocity field and the entrainment mechanism of the flow. Based on this analysis, we propose a model, in Chapter 3, for the transport and dispersion of passive tracers in the flow. This model is derived from a general effective advection–diffusion equation, using a mixing length hypothesis to model the turbulent eddy diffusivity. The theoretical predictions are then compared with experimental data in Chapter 4. We also study the statistical significance of the experimental data, and describe a method, based on these data, to assess pollution risks in quasi-two-dimensional jet flows. In Chapter 5, we explore further the turbulent relative dispersion mechanisms of the core and eddy structures using two-point statistical analysis. Then, Chapter 6 presents a potential model for the flow induced by quasi-two-dimensional jets in a rectangular domain. We study the impact of the induced flow on the jet. In Chapter 7, we investigate particle-laden jets confined in a quasi-two-dimensional environment. We compare the case of a dilute particle-laden jet (i.e. a particle-laden jet with a small bulk concentration of particles) with a particle-free quasi-two-dimensio- nal jet. Finally, we summarize the main findings of this thesis in Chapter 8 and discuss future work. The results presented in Chapter 2 have been published in Landel, Caulfield & Woods (2012a). Most of the results described in Chapters 3 and 4 have been submitted for publication in Journal of Fluid Mechanics, in an article by Landel, Caulfield & Woods (2012b, sub judice). We adopt a similar structure in every chapter, except in Chapters 3 and 4 which have a combined structure. The prob- lem studied in the chapter is introduced in the first section, which also includes a detailed review of past studies, and the last section is a conclusion of the chapter. 4 Chapter 2 Meandering due to large eddies and the statistically self-similar dynamics of quasi-two-dimensional jets 2.1 Introduction The study of turbulent plane jets is relevant to a wide variety of problems where both qualitative and quantitative knowledge of the concentration in time and space of tracers transported by the jet is needed (Kotsovinos, 1975). In many industrial applications, effluents, waste or even pollutants are released into large basins such as lakes or oceans. The source of the discharge can be rivers (see e.g. Rowland, Stacey & Dietrich, 2009, and references therein) or multiport dif- fusers (for an extensive study, see Jirka, 2006). In both situations, characteristic horizontal dimensions are much larger than the fluid-layer thickness and the flow 5 2 Meandering and self-similarity of quasi-two-dimensional jets develops in a confined environment. Early experimental studies of bounded plane jets by Foss & Jones (1968) and Holdeman & Foss (1975) showed the influence of secondary flows on the mean flow. However, as Giger et al. (1991) and Dracos et al. (1992) pointed out, these secondary flows disappear beyond a distance of 10 flow thicknesses. The present work focuses on this far-field region (z/W ≥ 10, where z is in the flow direction and W is the fluid-layer thickness), where the jet has been observed to meander due to the development of large eddies that grow on its sides. In this far-field region, the initially planar two-dimensional jet is referred to as a quasi-two-dimensional jet because of the influence of the spanwise restriction on the flow. The key characteristic of quasi-two-dimensional jets is the development of an instability (see Chen & Jirka, 1998, for a linear stability anal- ysis of shallow-water jets) featuring large planar counter-rotating eddies. Dracos et al. (1992) noted that the spanwise distribution of the velocity was approxi- mately uniform. Moreover, they found that in the far field the mean properties of the jet remained unchanged and turbulent energy was transferred to large scales thus indicating an inverse cascade characteristic of quasi-two-dimensional turbulence. Dracos et al. (1992) observed and studied the significance of large co- herent eddy structures in the jet. However, using only point measurements, they could not provide a complete dynamical study of these structures. Recently, Shin- neeb, Bugg & Balachandar (2011) conducted a statistical analysis of large vortical structures in shallow-water jets using particle image velocimetry. However, their layer thickness (W ∼ 5–15 d) was such that the flow evolution was inherently three-dimensional (albeit confined), and they did not focus on the far-field region because their measurements were taken only up to z/W ≤ 16. Their study was also uncorrelated in time, and so they were unable to identify the inherent time dependence of the flow quantitatively. We believe that a study of quasi-two-dimensional jets in the regime identified by Dracos et al. (1992) is necessary to assess the impact of the characteristic flow structures on the mixing, dispersion and diffusion of tracers in shallow jets, as sug- gested by Jirka (2001). For instance, undiluted patches of pollutants carried by a river discharging into the ocean can be disastrous for the local ecology. Informa- tion about the size, speed and typical travel distances of these patches is therefore crucial. To address this problem, we analyse the far field of a confined plane jet using particle image velocimetry. With a fully resolved velocity field in time and 6 2.2 Experimental procedure space, we can characterize the jet structure phenomenologically. We are par- ticularly interested in understanding quantitatively the relationship between the large-scale, and inevitably transient, flow structures and the long-time-averaged mean properties of the plane jet. The rest of this chapter is organized as follows. In § 2.2 we describe the ex- perimental procedure. In § 2.3 we then provide a qualitative overview of the flow structures observed from dyed-jet experiments and instantaneous velocity fields, while in § 2.4 we compare measurements of the time-averaged velocity field with classical theories for two-dimensional plane jets. In § 2.5 we present a quantita- tive study of the flow structures, in particular by tracking the large eddies as they interact with the high-speed core. We discuss how the frequency of occurrence of the eddies changes with distance due to eddy merger. The study of the probabil- ity density function of the core shows that the time-averaged mean distribution of the velocity is due to the large-scale dynamics of the core and eddy structure. Finally in § 2.6 we draw our conclusions. 2.2 Experimental procedure The experimental apparatus is shown schematically in figure 2.1. Water jets were discharged vertically upwards in a 1 m (L) × 0.01 m (W ) × 1 m (H) tank made of 10 mm thick Perspex sheets. An aluminium structure, made of two vertical beams located 0.4 m apart on each side of the jet axis and one horizontal beam located 0.8m above the nozzle, was added on each side of the tank to increase the rigidity of the walls and ensure a uniform gap width. Two overflows on the side of the tank maintained a constant water depth at 0.915 m. The flow was driven by a constant-head tank and discharged via a 0.1 m circular rigid tube of aspect ratio 20, leading to a 5mm (d)×10mm (W )×20mm chamber and finally into the tank. The aspect ratio of the tube was deemed sufficient to suppress any swirl in the flow. The flow rate was controlled through a valve and measured with a precision balance and a stopwatch for each experiment. The flow rate was found to be consistent in time with an accuracy of approximately 1 %. We conducted two distinct sets of experiments using two qualitatively different techniques: dye tracking and particle image velocimetry (PIV). 7 2 Meandering and self-similarity of quasi-two-dimensional jets x z d = 5 mm Constant-head tank Overflow H = 1 m L = 1 m W = 0.01 m u w 2b(z) CCD camera Red filter (dyed jets) Study area 2 (PIV) Study area 1 (PIV) Figure 2.1: Schematic diagram of the experimental apparatus. The two PIV study areas are shown with overlapping dashed lines. 2.2.1 Dye tracking experiments For the dye tracking experiments, we filled the tank with fresh tap water. We injected dark blue food dye through a needle placed 0.2 m upstream of the noz- zle. Simultaneously, we pumped the same volume of fluid out to minimize the disturbance introduced into the flow. Also, we injected the dye after the flow reached a steady state in the tank. We used diffuse ambient lighting for these experiments. A red filter was placed between the objective of the camera and the tank, as shown in figure 2.1, to increase the contrast between the jets and the background. The flow motion was recorded with a high-speed 8 bit grey-scale camera (Fastcam SA1.1 – Photron), mounted with a 62mm focal-length lens. We analysed 40 dyed jets with jet Reynolds number 2280 ≤ Rej = dws/ν ≤ 4030, where ws is the source velocity and ν is the kinematic viscosity of water, using the software code DigiFlow (Sveen & Dalziel, 2005). We determined the location of the edge of each dyed jet through an intensity criterion. Since the contrast between the dyed surface and the background was very strong but not saturated, the edge of the jet was very sharp. 8 2.2 Experimental procedure 2.2.2 Particle image velocimetry experiments For PIV experiments the tank was filled with water mixed with Pliolite VTAC particles of average diameter 0.23mm, which served as passive fluid tracers for the PIV. Approximately 2 mL of rinsing agent (Finish R© rinse aid) was added to the mixture to prevent aggregation of Pliolite particles. The small change in surface tension had no influence on the measurements. The choice of this particle size depended on both hydrodynamic and optical criteria (see e.g. Drayton, 1993). We find that the particle diameter is of the order of the smallest Kolmogorov length scale found in the flow, ηK ≈ 0.2 mm. Although this size is not optimal to study small-scale turbulence, it was the minimum size that could be detected by the image software while also capturing the largest length scales in the flow. The particle Stokes number based on the Kolmogorov time scale was StkK ≤ 10−1 (see Xu & Bodenschatz, 2008), which guaranteed that these particles followed the fluid motion closely. The particle concentration was kept relatively uniform at approximately 1.7 × 10−5 by volume due to the turbulence in the tank. Since the particle concentration was smaller than 10−3 by volume, particle–particle interactions and any changes in the fluid viscosity were insignificant (see Fung, 1990, for more discussion). We adjusted the water density to match the particle density of 1.03 g cm−3 by adding 35g of salt per litre of water. At rest, the particle distribution remained unchanged over 18 hours, thus confirming that the particles were neutrally buoyant. The mixture of salt water and particles recirculated in the experimental set-up in order to have identical conditions (particle concentration, water density and water temperature) for each experiment. We performed the PIV experiments in a dark room. Two 1kW filament photo- graphic lamps, each mounted with a long focal-length spherical lens to focus the filament into a sheet, illuminated the tank from above through a 5mm slit centred on the mid-plane (y = 0). Every effort was made to keep the width of the light sheet constant and smaller than the gap width in order to attenuate reflection issues with the tank walls. This also meant that we could not make any measure- ments away from the mid-plane (y = 0) because as we moved the light sheet closer to the wall in the narrow gap, reflection at the wall perturbed the measurements. From image inspection, the number of particles that appeared much slower than the rest, probably because they were trapped in the boundary layers, was suffi- ciently small (of the order of 10%) not to affect the imaging analysis and corrupt 9 2 Meandering and self-similarity of quasi-two-dimensional jets the computation of the velocity field. We recorded the flow motion using the same high-speed camera as described above. The camera filmed two 0.4m×0.4m study areas centred on the jet axis (as shown in figure 2.1). The frequency of image acquisition was set at 500 frames per second for a duration of 10.9 s for study area 1 and at 250 frames per second for a duration of 21.8 s for study area 2. The acquisition frequency was much higher than the largest Kolmogorov frequency scale. Moreover, the length of the video was long enough to compute meaningful temporal averages. Study area 1 covered a height from z = 0 to 0.4m, while study area 2 covered a height from z = 0.2 to 0.6m. Hence, the jet was studied from its source up to a distance of 120 d. The width of the study area is larger than the length scale of the jet at every height. The 1024×1024pixel images were analysed using DigiFlow (see Sveen & Dalziel, 2005, and references therein for more detail about the PIV algorithm used by DigiFlow). The spatial velocity resolution was at 6.6 mm based on interrogation areas of 17 × 17 pixels with 75 % overlapping. This resolution proved to be sufficiently small from z = 20 d upwards. Six steady turbulent jets of flow rates 33.2, 37.0 and 40.3 cm3 s−1 were investigated in both study areas. The jet Reynolds number was in the range 3320 ≤ Rej ≤ 4030. 2.3 Qualitative observations A sequence of grey-scale pictures of a typical injection of dye in a steady-state jet with Rej = 3850 is presented in figure 2.2 as the dye front rises through the full depth of the quasi-two-dimensional tank. These pictures reveal many interesting features of quasi-two-dimensional jets. The saturated dye clearly shows the maximum lateral extent of the turbulent jet. The dye gradually fills a triangle (plotted in black lines) which suggests that entrainment is self-similar with height, at least when averaged over sufficiently long times. Before filling the full triangle width, we can observe (especially in figures 2.2d and 2.2e) an oscillation of the jet, as the dye path is clearly sinuous. Large round structures corresponding to eddies can also be identified on either side of the centreline. Dracos et al. (1992) observed similar structures for a range of distances 10 ≤ z/W ≤ 120. The curvy edge of the jet suggests a characteristic scale, typically half the width of the triangle (approximately 10 cm at mid-height). These eddies result from the instability of the shear layer at the border between the jet and the ambient fluid (Jirka, 10 2.3 Qualitative observations 2001). Furthermore, tongues of ambient fluid (in white or light grey) appear at the rear of the largest eddies (see arrow in figure 2.2e). This phenomenon was also observed by Dimotakis, Miake-Lye & Papantoniou (1983) in the far field of round turbulent jets, and by Thomas & Brehob (1986) for two-dimensional turbulent jets. The role played by the eddies in the entrainment, by means of engulfment mechanisms at their rear, was modelled by De Young (1997) in an attempt to determine quantitatively the mass inflow contribution of large-scale structures in two-dimensional mixing layers. Although the eddies observed in quasi-two-dimensional jets, such as the jet presented in figure 2.2, have some similarities with eddies in planar two-dimensio- nal jets, it is important to note that the latter are genuine three-dimensional eddies while the former should be referred to as quasi-two-dimensional eddies because of the restriction imposed on the flow in the spanwise direction. The growth dynamics of quasi-two-dimensional eddies is governed by an inverse cascade of turbulence, while three-dimensional eddies tend to grow with mean-flow length scales. On the other hand, quasi-two-dimensional eddies also differ from purely two-dimensional eddies because friction at the boundaries, although relatively weak, restrains the maximum size of the eddies (Jirka, 2001) and eventually leads to their disintegration (Dracos et al., 1992). Finally, it is worth noting that at the leading edge the dye concentration attenuates suggesting that diffusion occurs in a steady jet. Diffusion in quasi-two-dimensional jets is likely to be the result of a complex interaction between the eddies and the sinuous turbulent core of the jet. We return to detailed investigation of this issue in Chapters 3 and 4. The second batch of experiments involved quantitative measurements of the velocity field using the PIV technique. Typical results for a jet at Reynolds number 4030 analysed in study area 2 are depicted in figure 2.3. In figure 2.3(a), a superposition of 40 images of the filming of the experiment shows the tracers as streaks to help visualize Eulerian structures in the flow. The corresponding velocity field is presented in figure 2.3(b), and it is clear that the main structures of the jet have been captured by the PIV. A high-speed core undulates along the centreline and is bordered by alternating counter-rotating eddies on the sides. The eddies are responsible for the entrainment and detrainment of fluid to and from the central core in a time-dependent fashion. Owing to the particular geometry of the tank, the turbulence cannot develop isotropically and we observe rather 11 2 Meandering and self-similarity of quasi-two-dimensional jets (a) (b) (c) (d) (e) Figure 2.2: Sequence of grey-scale pictures of a dyed jet (Rej = 3850) rising in the tank, at: (a) 1 s; (b) 2 s; (c) 3 s; (d) 4 s; and (e) 5 s. The average dye edge is plotted with black lines. The arrow in (e) points at the engulfment mechanism occurring at the rear of an eddy. an inverse turbulent cascade in which quasi-two-dimensional eddies grow with height (De Young, 1997). This mechanism is confirmed in the experiment, as flow structures increase in size as they are advected upwards. The schematic cartoon displayed in figure 2.3(c) summarizes these ideas. The time-averaged mean picture of quasi-two-dimensional jets is associated with a triangular shape encapsulating all the flow structures, while the time-dependent picture shows a sinuous core flanked by large growing eddies. This two-part structure remains self- similar with height and its dynamics is responsible for the Gaussian distribution of the mean velocity, as we will discuss in § 2.5. 2.4 Time-averaged mean flow field To characterize the mean behaviour of quasi-two-dimensional jets, we consider the ideal model of a turbulent momentum jet in a two-dimensional semi-infinite environment. Adopting the same conventions as Jirka & Harleman (1979), the flow is considered incompressible and steady in the mean. The x-direction is the lateral, cross-jet direction, the y-direction is the spanwise direction and the z-direction is the streamwise, axial direction. The velocity components are desig- nated by (u, v, w) for the Cartesian system (x, y, z) with the origin at the nozzle exit. We assume a plane flow in the domain: the velocity field and any other 12 2.4 Time-averaged mean flow field (a) (b) (c) Figure 2.3: (a) Passive tracers (Pliolite particles) shown as streaks in a typical jet (Rej = 4030) filmed in study area 2. (b) Velocity field (arrows) and vorticity field (background) of the same jet. (c) Schematic diagram describing the structure of qua- si-two-dimensional jets. quantities are invariant with y, and v = 0 everywhere. This hypothesis can be justified in three ways: the velocity profile across the gap must be self-similar in the core and the influence of the boundary layers is of second order at high Reynolds number; the v-component of the velocity is negligible compared to the other two components; and ambient fluid can only be entrained from the sides of the jet, i.e. in the x-direction. We also use the common hypothesis of a Gaussian profile (see, for instance, List, 1982) for the time-averaged streamwise velocity, w(x, z) = wm(z) exp [ − ( x b(z) )2] , (2.1) where the over-bar represents an appropriate average in time, wm(z) is the max- imum streamwise velocity at distance z from the source and b(z) is a measure of the local lateral spread of the jet velocity. We derive briefly the governing equations for plane jets, based upon the conservation of volume and momentum (see, for instance, Kotsovinos & List, 1977, for more details). The time-averaged volume flux and the time-averaged momentum flux are expressed respectively as Q(z) = ∫ ∞ −∞ w(x, z) dx and M(z) = ∫ ∞ −∞ (w)2 (x, z) dx. (2.2a,b) Solving the first-order integrated equations of motions dM dz = 0, dQ dz = 2αwm, (2.3a,b) 13 2 Meandering and self-similarity of quasi-two-dimensional jets we find M = M0, Q = Q0 ( 4 √ 2αM0zQ02 + 1 )1/2 , (2.4a,b) where we assume in equation (2.3b) that the entrainment velocity is proportional to the maximum time-averaged streamwise velocity, with α the entrainment co- efficient (Morton, Taylor & Turner, 1956), and Q0 and M0 are values at the origin for the volume flux and momentum flux, in (2.4a) and (2.4b) respectively. The e-folding value of the maximum time-averaged streamwise velocity and the maximum time-averaged streamwise velocity are, respectively, b(z) = Q0 2 √ 2πM0 ( 4 √ 2αM0zQ02 + 1 ) and wm(z) = √ 2M0 Q0 ( 4 √ 2αM0zQ02 + 1 )−1/2 . (2.5a,b) We can infer the theoretical virtual origin of the jet z0 = −Q02/(4 √ 2αM0), (2.6) which results from the choice of the boundary conditions (i.e. the distributions of the volume flux and momentum flux at z = 0). Alternatively, solving the plane jet equations assuming momentum-flux conser- vation and similarity (see e.g. Pope, 2000) also leads to the same power laws for the e-folding value of the maximum time-averaged streamwise velocity, b ∝ (z − z0), and the maximum time-averaged streamwise velocity, wm ∝ (z − z0)−1/2. The constants of proportionality and the virtual origin can differ because of the as- sumptions we make for the x-distribution of wm (essentially due to ‘shape factors’) and for the boundary conditions. As a direct comparison with the ‘velocity spread rate’ S defined as dx1/2/dz = S (where x1/2 is the velocity half-width defined by wm(z)/2 ≡ w(x1/2, z)), we can remark that S = 4(ln 2/π)1/2α (see Pope, 2000, for further details about S). Equations (2.5a,b) suggest that the natural scalings for length and time scales in our problem are d, the source width, and τ = d2/Q0, respectively. Therefore, when considering our experimental data, we will always scale quantities with these scalings, i.e. z˜ = zd, x˜ = x d , b˜ = b d, t˜ = t τ , w˜ = τ dw, (2.7a–e) where tildes denote non-dimensional variables. 14 2.4 Time-averaged mean flow field For comparison with the theoretical model, we time-averaged the velocity field measured with PIV. We plot the lateral spread, the evolution with height and the lateral distribution of the time-averaged streamwise velocity. We also discuss the influence of the free surface at the top boundary, the impact of the lateral confinement and possible three-dimensional effects on the flow, such as friction at the walls constraining the flow. In figure 2.4, we show the ensemble average of the edges of 40 dyed jets (plotted with dots). The evolution of the dye edge with height clearly indicates that above z/d = 120 the influence of the free surface becomes non-negligible. This height serves as a lower bound for the ‘impingement region’ (see Jirka & Harleman, 1979, for a detailed study). The zone of established flow is found to start at approximately z/d = 20, a value at which the streamwise velocity becomes self- similar. This value is commonly reported in the literature (see e.g. Kuang, Hsu & Qiu, 2001). A linear fit of the non-dimensional average dye edge (plotted with a thin line in figure 2.4) calculated for 20 ≤ z/d ≤ 120 gives a slope of 0.22± 0.08 for the half-spreading angle. We can observe that the non-dimensional e-folding value of the maximum time-averaged streamwise velocity b˜ (plotted with crosses) is much narrower. We discuss this difference further below. We also compute the quantity b˜ from the ensemble average of the 12 jets studied with PIV. A linear fit (plotted with a thick line) calculated between 20 ≤ z/d ≤ 120 gives the rate of change, db/dz = 0.154±0.016, which is slightly above the value of 0.135 reported by Ramaprian & Chandrasekhara (1985) and very similar to the value reported by Albertson et al. (1950). Using (2.5a) the corresponding entrainment coefficient (determined to best-fit the streamwise variation of b) is αb = 0.068±0.007 (which is equivalent to Sb = 0.125± 0.015), and we find that αb is almost constant in the zone of established flow, thus confirming the entrainment assumption (Morton et al., 1956). In figure 2.5, we plot the non-dimensional maximum time-averaged streamwise velocity wm/(Q0/d) against height. The crosses are plots of an ensemble average over all the jets studied with PIV. Although the agreement is good, they lie slightly but systematically above the theoretical curve (plotted with a solid line) for z/d ≤ 100. We compute the theoretical curve from (2.5b) and using α = αb. The value of Q0 was measured for each jet as described in the experimental procedure. On the other hand, since M0 could not be measured directly, it was 15 2 Meandering and self-similarity of quasi-two-dimensional jets 0 5 10 15 20 25 30 x/d 0 20 40 60 80 100 120 140 160 z/ d Average dye edge Fit of dye edge Average b˜ Fit of b˜ b˜ from best fit of wm/(Q0/d) Figure 2.4: Non-dimensional average dye edge (dots) with a linear fit (thin line); non-dimensional e-folding value of the maximum time-averaged streamwise velocity b˜ (crosses) with a linear fit (thick line); and a non-dimensional average velocity spread rate (dashed line) using α = αwm computed from the best fit of wm/(Q0/d) (see figure 2.5). replaced by M by virtue of (2.4a). As shown in figure 2.6, the momentum flux M (plotted with pluses) computed from the time-averaged streamwise velocity field using (2.2b) (the boundaries of the integral are chosen as the positions where wm = 0) is found to be approximately constant for 34 ≤ z/d ≤ 110. For z/d < 34, the data do not seem accurate, probably because the frame rate is not high enough for the large velocity at that distance, and the resolution of the PIV could also not be optimal where the jet is very narrow. For z/d > 110, the influence of the impingement region as the jet approaches the free surface at the top seems to start affecting the momentum flux. The mean non-dimensional value of the momentum flux is < M > / ( Q02/d ) = 0.55 ± 0.03 (plotted with a solid line in figure 2.6). Giger et al. (1991) reported and discussed the wide range of values for the non- dimensional momentum flux measured in plane jets from the literature: from 0.52 (Cervantes de Gortari & Goldschmidt, 1981) to 1.77 (Antonia, Satyaprakash & Hussain, 1980). We analysed the influence of friction at the rigid boundaries on the momentum flux and found a Fanning friction factor of f ≈ 0.007 assuming 16 2.4 Time-averaged mean flow field 0 20 40 60 80 100 120 z/d 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 w m /( Q 0 /d ) Data Theory Best fit Figure 2.5: Non-dimensional maximum time-averaged streamwise velocity (pluses) versus height, theoretical curve (solid line, and using αb = 0.068) and best least-squares fit (dashed line) optimising with respect to α (using αwm = 0.052). a wall stress of the form τw = fρ < w >2 /2, where ρ is the water density and < w > is the spatial averaged velocity in the y-direction (Bird, Stewart & Lightfoot, 2007). The influence of friction is relatively small compared to the mean value of the momentum flux (of the order of 10 %) and therefore has not been included in our constant momentum-flux model (see Giger et al., 1991, for a detailed study). In figure 2.5, a least-squares fit of the data (plotted with a dashed line) assuming equation (2.5b) and optimising with respect to the entrainment coefficient yields an optimal choice for α from the z dependence of the maximum velocity αwm = 0.052 (which is equivalent to a velocity spread rate Swm = 0.098). The fact that αwm (also plotted with a dashed line in figure 2.4) is slightly smaller than αb means that some assumptions of the model underlying (2.5a,b) (which should yield identical estimates for α using b(z) and wm(z)) are not perfect. In particular, we believe that the Gaussian distribution hypothesis is not ideal, as slight deviations from Gaussianity could explain the mismatch. In figure 2.7(a), we show the lateral distribution of the normalized time-ave- raged streamwise velocity w/wm. The x-axis is centred on the position of the 17 2 Meandering and self-similarity of quasi-two-dimensional jets 0 20 40 60 80 100 120 z/d 0 0.2 0.4 0.6 0.8 1.0 M /( Q 0 2 /d ) Data Average (34 < z/d < 110) Figure 2.6: Non-dimensional time-averaged momentum flux (pluses) versus height and average value < M > / ( Q02/d ) = 0.55 (solid line). maximum time-averaged streamwise velocity. All the curves result from an en- semble average of six or 12 jets, depending on where the z position of the curve lies with respect to the two study areas for the PIV. The experimental data (plot- ted with different colours) are in very good agreement with the theoretical curve (plotted with a thick red line) computed from equation (2.1) using α = αb and neglecting any consideration of virtual origin (Kotsovinos, 1976). Nevertheless, the experimental curves are somewhat narrower than the theoretical Gaussian velocity profile. This discrepancy is consistent with a smaller entrainment coeffi- cient, as suggested by the best fit of wm/(Q0/d) in figure 2.5. The mismatch is probably caused by the return flow in the tank which is not accounted for by the model, where an infinitely wide domain is assumed. The problem of the return flow in a domain of finite lateral extent is more promi- nent in plane jets than in (fully unconfined non-planar) three-dimensional jets. In plane jets, the entrainment velocity remains constant outside the jet, whereas it decreases with distance in the three-dimensional case. As a consequence, we can observe a negative shift in the lateral distribution of the time-averaged streamwise velocity (see figure 2.7a), which denotes the presence of the return flow. The flux 18 2.4 Time-averaged mean flow field −0.6 −0.4 −0.2 0 0.2 0.4 0.6 x/z −0.2 0.0 0.2 0.4 0.6 0.8 1.0 w /w m z/d =20 z/d =40 z/d =60 z/d =80 z/d =100 z/d =120 Theory −0.6 −0.2 0.2 0.6x/z −0.2 0.2 0.6 1.0 (w + |w r |)/ (w m + |w r |)(a) (b) Figure 2.7: (a) Lateral distribution of the normalized time-averaged streamwise veloc- ity at various heights (plotted with different colours) and theoretical prediction (plotted with a thick red line). (b) Lateral distribution of the normalized sum of the time- averaged streamwise velocity and the absolute value of the estimated time-averaged streamwise velocity of the return flow (defined by equation 2.8) at the same heights as in (a) (plotted with different colours) and theoretical prediction (plotted with a thick red line). of the return flow, Qr, increases with height, as it matches the jet volume flux Q owing to conservation of volume at every height across the width of the tank. We can estimate the time-averaged streamwise velocity distribution of the return flow by applying volume conservation at each height. For all z, the total volume flux on both sides of the jet is Qr = Q − Q0. We assume that the return veloc- ity is distributed uniformly along −L/2 ≤ x ≤ −x0 and x0 ≤ x ≤ L/2, where x0 ≈ 0.25z is defined as the location where w = 0. Therefore, using equations (2.4b) and (2.5b) we find that the time-averaged return velocity is wr wm = − Q0 2 2 √ 2M0(L/2− x0) [( 4 √ 2αM0Q02 z + 1 ) − ( 4 √ 2αM0Q02 z + 1 )1/2] , (2.8) where we use α = αb = 0.068 and M0 =< M >= 0.55Q02/d to plot figure 2.7(b). 19 2 Meandering and self-similarity of quasi-two-dimensional jets As can be seen in figure 2.7(b), adding this simple estimate of the return-flow ve- locity wr to the jet velocity w has corrected the negative shift in the experimental data. At every height, except z/d = 20, the velocity tends to a zero asymptotic value for large |x/z|. From comparison with similar experiments that we conducted in a smaller do- main (0.5 m × 0.01 m × 0.5 m) and with experimental results reported in the literature and obtained in larger tanks of various aspect ratios and with porous or non-porous lateral boundaries (Giger et al. (1991); Dracos et al. (1992); Rowland et al. (2009)), we believe that the impact of the return flow is limited and affects principally the distribution of the time-averaged streamwise velocity in the man- ner described above. From direct measurements we also find that the momentum flux associated with the return flow is small compared with the momentum flux in the jet (from 0 to 15% for z/d = 0 to 110). We have not observed any qualitative or quantitative influence of the return flow on the time-dependent core and eddy structure described in § 2.3. We discuss the spatial structure of this return flow in more detail in Chapter 6. The experimentally measured streamwise velocity field follows closely the pre- dictions given by the derivation of the momentum and continuity equations for two-dimensional turbulent jets. The small difference due to the lateral confine- ment of the experimental jets leads us to the conclusion that the entrainment coefficient is within the range 0.052 ≤ α ≤ 0.068. The purpose of the study of the mean flow is not to understand all the details of this flow but rather to give us some insight about the flow field in this particular geometry. More refined models for the plane jet can be found in the literature (see e.g. Giger et al., 1991; Hussein et al., 1994; Wang & Law, 2002). 2.5 Quantitative analysis of the time-dependent core and eddy structure 2.5.1 Time-dependent eddy structure We now analyse the core and eddy structure of the flow using the experimental results given by the PIV. We identify large vortical structures or ‘eddies’ in in- dividual frames of the instantaneous velocity field using DigiFlow, as shown in 20 2.5 Quantitative analysis of the time-dependent core and eddy structure figure 2.8(a). Considering a specific frame, we find regions of the instantaneous flow field where streamlines form a complete loop. This technique is similar to the eddy identification method proposed by Robinson (1991). We plot the streamlines forming a complete loop in figure 2.8(a) with grey curves. We then analyse each patch, or eddy, to obtain statistical measurements such as the centroid (identified by the location of the black crosses) and the standard deviations in the lateral and streamwise directions (shown by the size of the crosses). The coordinates of the centroid (xc,k, zc,k)(t) of eddy ‘k’ at time t are computed numerically as (xc,k, zc,k)(t) = 1 Lx∑ x=0 Lz∑ z=0 ∆k(x, z, t) Lx∑ x=0 Lz∑ z=0 (x, z)(t)∆k(x, z, t), (2.9) where Lx and Lz are the lateral and streamwise dimensions of the velocity field and ∆k(x, z, t) is 1 if the point (x, z)(t) belongs to a streamline identified as part of eddy k at time t and 0 otherwise. Similarly, the lateral and streamwise standard deviations (xs,k, zs,k)(t) of eddy k at time t are computed numerically as (xs,k, zs,k)(t) =   1 Lx∑ x=0 Lz∑ z=0 ∆k(x, z, t) × Lx∑ x=0 Lz∑ z=0 ( (x, z)(t)− (xc,k, zc,k)(t) )2 ∆k(x, z, t)   1/2 . (2.10) We applied the algorithm every 10 frames to the six experimental velocity fields. The eddies are thus tracked in time at a frequency of 25 frames per second. As a quality control of the technique, we conducted a visual inspection of all the eddies identified by the algorithm. This showed that the algorithm was very robust. It did not appear to be subject to ‘false positives’, i.e. the misidentification of a non- eddy feature of the flow as an eddy. The algorithm only occasionally failed to detect eddies (i.e. there were very few ‘false negatives’) when the eddies partially 21 2 Meandering and self-similarity of quasi-two-dimensional jets appeared at the edges of the frame. (a) (b) Figure 2.8: (a) Identification of the vortical structures in the instantaneous velocity field (plotted with black arrows). The streamlines identified within an eddy are plotted with grey curves. The black crosses designate the centroid of the eddies, and the size of the crosses represents their standard deviations in the lateral and streamwise directions. (b) Identification of the core structure in the same instantaneous velocity field (plotted with black arrows). The streamlines identified as part of the core structure are plotted with grey curves. The trajectories of 48 eddies are shown in figure 2.9 (plotted with black dots). A linear fit (shown with a solid line) gives an average slope of 0.22 from the z-axis. We also plot the linear fits of the ensemble-averaged lateral standard deviation of the eddies, a measure of the average eddy width, with dashed lines. The lateral and streamwise standard deviations were found to be almost identical, showing that the eddies are close to circular in shape. They both have a linear trend increasing with height at a rate of 0.07. The non-dimensional location of each eddy in time has been plotted in fig- ure 2.10 with dots. We can see a general trend, which follows the power law z˜ ∝ t˜ 2/3 (plotted with a solid curve) derived from the maximum time-averaged streamwise velocity formula (2.5b). The large scatter is due to the complex dy- namics of individual eddies. We found that not all eddies travelled through the depth of the PIV window completely unperturbed. We observed merging of close successive eddies, with the first eddy slowing down considerably, sometimes even halting, and the following eddy accelerating substantially. Similarly to the ob- servations made by Dracos et al. (1992), we did not see eddies rotating around 22 2.5 Quantitative analysis of the time-dependent core and eddy structure −40 −30 −20 −10 0 10 20 30 40 x/d 50 60 70 80 90 100 110 120 130 140 z/ d Data Average eddy path Average eddy width Figure 2.9: Eddy locations in PIV study area 2 (dots), linear fits of eddy locations (solid lines) and average eddy lateral standard deviations (dashed lines). a common axis before merging. We also noticed some small eddies disappearing in the vicinity of the core. From the best fit of the eddy streamwise position in time, we find that the average eddy speed is 0.28 times the theoretical wm and 0.24 times the best fit of wm. The fastest identified eddy rises approximately at the same speed as the centreline time-averaged streamwise velocity, whereas the slowest eddy rises at less than 15 % of wm. To investigate the eddy frequency, we counted the number of eddies (identified by the algorithm described above) passing at a given height on either side of the core of the jet. We measured this number for the six PIV experiments performed in study area 2 and then divided it by the duration of the experiment, i.e. 21.8 s. The resulting non-dimensional eddy frequency f˜ = fd2/Q0 is plotted with thin lines in figure 2.11(a). Dracos et al. (1992) found an empirical law for the eddy frequency, f˜ = 176z˜−3/2 (plotted in non-dimensional form with a dashed curve), and explained that f decreased as a result of the decrease of the eddy trans- port velocity and merging mechanisms. However, if we assume that eddies form periodically, then the eddy frequency should remain constant with height, since eddies travel on average at the same velocity. The frequency can only decrease 23 2 Meandering and self-similarity of quasi-two-dimensional jets 0 200 400 600 800 1000 1200 1400 t/(d2/Q0) 50 60 70 80 90 100 110 120 z/ d Data Best fit Figure 2.10: Eddy z-coordinate versus time (dots) and best least-squares fit (solid curve) assuming z˜ ∝ t˜ 2/3. if eddies merge (or, to a lesser extent, disappear). Merging occurs when the dis- tance between two successive eddies is smaller than a critical value. The distance between two eddies decreases because their transport velocity decreases like z˜−1/2 and because eddies grow approximately linearly with height due to entrainment of ambient fluid. The punctuated decrease in frequency can actually be observed in figure 2.11(a) as we follow individual experiments (see the values of f for a typical experiment plotted with red crosses). The frequency f is constant over a certain distance and then drops by a discrete value in a stepwise way. This is also clearly shown by the evolution of the Strouhal number, St = fb/wm, plotted with dots in figure 2.11(b) and with red crosses for the same typical experiment. The Strouhal number increases like St ∝ z˜ 3/2 from a minimum value of St = 0.07 consistent with the value reported by Dracos et al. (1992) (plotted with a dashed curve) and then drops, somewhat chaotically but consistently, to this minimum value as merging occurs. Because merging becomes less frequent as z/d increases, the length of time over which f is constant (and hence St increases) increases with z/d. This leads to the increase in both typical values of the Strouhal number and 24 2.5 Quantitative analysis of the time-dependent core and eddy structure 50 60 70 80 90 100 110 120 z/d 0 0.05 0.10 0.15 0.20 0.25 0.30 S t Data Individual experiment Dracos et al. (1992) (b) 60 80 100 120 z/d 0 2 4 10 00 f d2 /Q 0 Data Individual experiment Dracos et al. (1992) (a) Figure 2.11: (a) Data for the non-dimensional eddy frequency fd2/Q0 versus height (thin lines) and best fit of Dracos et al. (1992) (dashed curve). The values of fd2/Q0 for a typical individual experiment have been highlighted with red crosses. (b) Data for the Strouhal number St = fb/wm versus height (dots) and Strouhal number reported by Dracos et al. (1992) (dashed line). The values of St for the same experiment have also been highlighted with red crosses. its variance, as is apparent in figure 2.11(b). The actual value of the minimum Strouhal number appears to depend on the eddy formation frequency, the travel speed of the eddies, the growth of the eddies due to entrainment and the dynamics of merging, in ways that are not as yet fully understood. To summarize, the eddies have on average a linear trajectory, a constant growth with height and a velocity similar to the time-averaged mean streamwise velocity of the jet. All these findings lead to the conclusion that the dynamics of the eddies is essentially self-similar with height, at least within the region of the flow that we have studied. From the analysis of the time evolution of the streamlines leading to the eddies, we can also attribute the growth of the eddies mainly to the entrainment of ambient flow. Eddy merging occurs irregularly and is responsible for the decrease of the long-time-averaged eddy frequency, with an apparently well-defined minimum Strouhal number St ≥ 0.07. 25 2 Meandering and self-similarity of quasi-two-dimensional jets 2.5.2 Time-dependent core structure Similarly, we identify the core of the jet by plotting all the streamlines that exit through the top of a specific velocity field. Effectively, the algorithm follows the streamlines backwards starting from the points at the top horizontal boundary of the velocity field. However, in the following discussion, we consider the stream- lines in the forward direction with their endpoint at the top of the velocity field. The identification of the streamlines of the core is repeated every 10 frames for each PIV velocity field, thus giving a dynamical picture of the core at a frequency of 25 frames per second. It can be seen in figure 2.8(b) that some streamlines (plotted with grey curves) start at the bottom boundary of the window while oth- ers come laterally inwards. The streamlines coming from the bottom of the frame reveal the volume flux brought by the jet itself into the frame. The streamlines coming from the sides of the jet show the entrained volume flux. They actually reveal how entrainment of ambient fluid occurs as they wrap around eddies and then are incorporated into the core. It is clear that eddies constitute an essen- tial entrainment mechanism by engulfing ambient fluid at their rear. The starting point of entrained streamlines (i.e. the location at which we consider them as part of the core) is chosen where the streamwise component of their gradient changes sign. This choice raises the more fundamental question about the boundaries of the core. The boundary between the core and the eddy is clearly defined since the algorithms used to identify both structures ensure mutual exclusion. However, at the top and bottom of the window, this boundary can be ambiguous if large eddies are not entirely seen in the image frame. At the top of the frame, the error zone is actually restricted to z > 118 d, which is approximately where the self-similarity region of the jet ends. At the bottom of the frame, the error zone is insignificant since the eddies are much smaller. Moreover, we found that the starting point chosen for entrained streamlines has no effect on the time-averaged distribution of the core and negligible impact on time-dependent distributions. Therefore, although somewhat arbitrary, we believe that our criterion determining the boundary between the core and the ambient flow reflects the diffusion of momentum from the jet to the ambient flow. We present the lateral (or x-) distribution of the probability Pcore(x, z) of be- ing in the core in time (plotted with thick solid curves) at different heights in 26 2.5 Quantitative analysis of the time-dependent core and eddy structure figure 2.12. The discrete formulation of the probability Pcore is Pcore(x, z) = 1 N N∑ n=0 ∆n(x, z), (2.11) where N is the total number of frames for a given experiment, n designates the nth frame and ∆n(x, z) is 1 if the point (x, z) belongs to a streamline identi- fied as part of the core of the jet in the nth frame and 0 otherwise. Its shape is Gaussian-like on the edges and flatter in the middle. The flat portion where Pcore(x, z) = 1 corresponds to the section of the jet always occupied by the core in time. The width of this section grows linearly with height on average, as shown by the standard deviation measurement xstd (plotted with thin solid curves for the experimental data and dashed lines for the linear fits), at a rate of 0.12. Fur- thermore, the momentum flux of this portion remains constant with height at a value of 78% of the total momentum of the jet. The edges of the probability Pcore correspond to the lateral excursions of the core through time. It is interesting to note a similarity between the distribution of the probability Pcore, as presented in figure 2.12, and a typical distribution of the intermittency function measured in quasi-two-dimensional jets (see e.g. Dracos et al., 1992). Both display a plateau equal to 1 in the interior of the jet and a Gaussian-like decrease tending towards 0 as |x/z| increases. Nevertheless, the intermittency function and the probabil- ity Pcore are different, both in the way they are computed and in their meaning. The probability Pcore is a measure of the likelihood of being in the core in time (which is identified by the algorithm described above). On the other hand, the intermittency criterion measures the probability of being in a turbulent region in time. The similarity observed between these two functions is probably due to the fact that the core is a region where the amplitude of the turbulent fluc- tuations increases towards the jet centreline. However, the lateral spreading of the two functions should differ because, contrary to the intermittency function, the probability Pcore excludes the eddies, which are also regions of large velocity fluctuations. A typical standard deviation xstd(t) of the distribution of the core streamlines at the time instant corresponding to the jet shown in figure 2.8(b) is plotted with dashed curves in figure 2.13. The undulations of the jet, which we already ob- served on dyed jet pictures, are primarily a feature of the edges of the core. The 27 2 Meandering and self-similarity of quasi-two-dimensional jets −40 −30 −20 −10 0 10 20 30 40 x/d 40 60 80 100 120 140 z/ d 0 1 P co r e Pcore xstd Fit of xstd Figure 2.12: Time-averaged mean core structure in PIV study area 2. Lateral distribu- tion of the probability Pcore of being in the core in time (thick solid curves) at different heights, and data (thin solid curves) and linear fits (dashed lines) of the time-averaged standard deviation xstd of the probability Pcore. distribution of the instantaneous streamwise velocity w(t) corresponding to the same time instant is plotted with solid curves at different heights in figure 2.13. We normalize w(t) with the maximum instantaneous streamwise velocity wn mea- sured at the lowest height in the frame, z˜n = 42.4. We can observe that the instantaneous velocity decreases with height and spreads laterally as expected from the self-similar theoretical model. Furthermore, the velocity distribution is not centred on the z-axis but follows the undulations of the core described by xstd(t). The velocity within the core is much larger than the velocity outside, thus underlying the presence of this high-speed core in the jet. It is also very interesting to note that the lateral decrease of the velocity is slower in the inte- riors of the undulations than in the exteriors. This is due to the presence of the eddies (shown as crosses, with the size of the crosses representing the lateral and streamwise eddy standard deviations) located in the curves of the core structure and which carry some upwards momentum flux (slightly less than a quarter of the total momentum flux on average). 28 2.5 Quantitative analysis of the time-dependent core and eddy structure −40 −30 −20 −10 0 10 20 30 40 50 x/d 40 50 60 70 80 90 100 110 120 130 z/ d 0 25 w (c m s− 1 ) w(t)/wn xstd(t) Typical eddies Figure 2.13: Distribution of the instantaneous normalized streamwise velocity w(t)/wn (solid curves) at different heights and corresponding to the jet presented in figure 2.8(b). Instantaneous standard deviation xstd(t) (dashed curves) of the core of the jet presented in figure 2.8(b) with its eddies (crosses, with the size representing the lateral and streamwise eddy standard deviations). The linear growth of the core shows that it is self-similar with height within the flow region studied, as we found for the eddies. The spatial statistical dis- tribution of the location of the core is due to its particular wave-like dynamics. The undulations along the centreline of this high-velocity core are characterized by an essentially self-similar spatial probability distribution Pcore. The standard deviation of the probability Pcore increases with height at a rate of 0.12, which is quite close to the rate of change with height of the mean velocity spread rate db/dz = 0.15. The spatial Gaussian distribution of the time-averaged mean streamwise velocity is therefore the result of the statistical spatial distribution of the undulating core. It is difficult to assess whether the eddies have a di- rect contribution to this statistical process. However, their role in the large-scale dynamics of the core is essential. 29 2 Meandering and self-similarity of quasi-two-dimensional jets 2.6 Conclusion In this experimental study of quasi-two-dimensional turbulent jets (and similarly to Giger et al. (1991) and Dracos et al. (1992)), we have observed that the flow organizes into a very interesting structure with a sinuous core of high streamwise velocity oscillating about the centreline and eddies rising and growing along the undulations. As predicted by the theoretical model, we find that: the mean ve- locity field measured with PIV is self-similar with height (see figure 2.14); the normalized time-averaged streamwise velocity profile w/wn (plotted with thick solid curves, where wn is the maximum time-averaged streamwise velocity at the lowest height, z˜n = 42.4) is close to a Gaussian distribution; and the velocity peak decreases as z˜−1/2 with height. The return flow due to the lateral confinement of the jet could explain the small mismatch between the theory and the experi- mental results. Friction at the bounding walls has only a second-order effect on the momentum flux (of the order of 10 % compared to the average value of the momentum flux) and thus on the velocity field. Within the flow region studied, we also find that both the eddies (average eddy paths plotted with dashed lines) and the core (time-averaged standard deviation plotted with thin solid lines) are on average self-similar with height, which is not described by the theory and is fundamentally different from either a (fully unconfined planar) two-dimensional jet or a (fully unconfined non-planar) three-dimensional jet, where the turbulence is unconfined and three-dimensional. The confinement of the jet in a narrow gap undoubtedly changes the struc- ture of the turbulence in the flow with a quasi-two-dimensional inverse cascade allowing large eddies to grow. This persistent growth of eddies is contrary to three-dimensional turbulence. The eddies form within the intense shear layer at the boundary between the jet and the ambient flow when the width b of the jet is larger than the thickness W of the flow (Dracos et al., 1992). Then, the eddy structures appear periodically at a given height. The eddy frequency decreases with height due to merging and we find a well-defined minimum Strouhal number St ≥ 0.07. The dynamics of these eddies is strongly coupled with the dynamics of the core. The core, which moves on average four times faster and carries approx- imately 75 % of the momentum flux, flows round the eddies. The consequence of these lateral excursions is seen in the mean velocity field. We believe that the un- stable dynamics of the core characterized by its probability density distribution 30 2.6 Conclusion is principally responsible for the Gaussian profile of the time-averaged stream- wise velocity. In this flow, it is the two-dimensional macrostructure and not the three-dimensional small-scale turbulence that produces the Gaussian distribution. Therefore, analysing the instantaneous flow field is key to understanding how entrainment, mixing and dispersion occur in the jet. The eddies play a leading role in the entrainment by engulfing ambient fluid at their rear, as we noticed from the study of the streamlines in the core and eddy structure. This entrainment mechanism ensures the linear growth of both the core and the eddies, therefore explaining the self-similarity of these structures. The exchange of fluid between the core and the eddies is permanent and in both directions as streamlines evolve in time from being closed within an eddy to being open and stretched in the core. It is perhaps surprising that the entrainment assumption of Morton et al. (1956), modelling entrainment due to three-dimensional turbulent mechanisms, can also describe the fundamentally different two-dimensional case. We find that the en- trainment coefficient is 0.052 ≤ α ≤ 0.068, depending on how it is calculated. This range of values for the entrainment coefficient is very similar to the values reported in the literature for (fully unconfined planar) two-dimensional jets: for example, 0.060 (Ramaprian & Chandrasekhara, 1985), 0.069 (Albertson et al., 1950). The dyed jet experiments revealed the vigorous mixing effect of the ed- dies. It is also worth noting that the average dye edge (shown with dotted lines in figure 2.14) coincides with the average outer boundaries of the eddies, which is the physical maximum lateral extent of the jet. Mixing is apparently not as strong in the core, but intense stretching leading to large streamwise dispersion occurs at the interface with the eddies. This region is delimited between the thin solid lines and the dashed lines shown in figure 2.14. In conclusion, a probabilistic description of the core–eddy structure of quasi- two-dimensional jets leads to a self-similar Gaussian description of the time- averaged flow. The instantaneous flow has a very different character from either (fully unconfined planar) two-dimensional flows or (fully unconfined non-planar) three-dimensional flows. Bulk long-time-averaged properties are consistent with conventional theoretical models, but the mixing and dispersion cannot be ac- counted for by these time-averaged models. We present a model for this mixing and dispersion in the next chapter. 31 2 Meandering and self-similarity of quasi-two-dimensional jets −40 −30 −20 −10 0 10 20 30 40 x/d 40 50 60 70 80 90 100 110 120 130 z/ d w/wn Core xstd Average eddy path Average dye edge Figure 2.14: Distribution at different heights of: normalized time-averaged streamwise velocity w/wn (thick solid curves); time-averaged standard deviation of the mean core xstd (thin solid lines); ensemble-averaged mean trajectory of eddies (dashed lines); and average dye edge (dotted lines). 32 Chapter 3 Advection–diffusion model for the streamwise transport, dispersion and mixing in quasi-two-dimensional jets 3.1 Introduction In the event of a spill of pollutants, waste or any other tracers into a river, it is crucial to predict how the tracers are advected and dispersed by the flow after they reach a relatively shallow basin, such as a lake or the sea shelf. Such pre- dictions can be used to monitor the spread of the tracers, control their impact on the environment and assess any potential damage. One of the most impor- tant aspects of these shallow river flows, and one which has raised the interest of scientists for more than 20 years, is the emergence of large-scale eddy structures and meanders at some distance away from the river mouth. These eddies and 33 3 Model for the streamwise transport, dispersion and mixing meanders have been visualized in nature on several occasions due to sediments transported by the flow (see e.g. Giger et al., 1991; Jirka & Uijttewaal, 2004; Rowland et al., 2009). Giger et al. (1991) were interested in the entrainment and mixing in shallow water flows, whose characteristic streamwise dimensions were much larger than the fluid-layer thickness and where the flow developed in a con- fined environment. They showed that these geophysical flows could be reproduced in laboratory experiments by confining plane turbulent jets in the spanwise direc- tion (i.e. the direction parallel to the line source of the jet). Giger et al. (1991) observed that in the far field, or for z/W > 10 where z is the spatial coordinate in the streamwise direction and W is the fluid-layer thickness in the spanwise direction (i.e. W corresponded to the depth of the basin), the jet produced sim- ilar large eddies and meanders as observed in shallow river flows. In Chapter 2, we referred to turbulent plane jets in such a confined geometry in the far field as quasi-two-dimensional jets and considered in detail the meandering flow due to the large-scale eddy structures. The present chapter focuses on the advection and dispersion properties of such quasi-two-dimensional jets, particularly when considering the transport of a passive scalar. The essential characteristics of quasi-two-dimensional jets have been described previously. Dracos et al. (1992) showed that the large planar counter-rotating ed- dies observed in quasi-two-dimensional jets developed due to an inverse cascade of quasi-two-dimensional turbulence. Chen & Jirka (1998) proved through linear stability analysis that the meanders of the jet were the result of a sinuous insta- bility. According to Jirka & Uijttewaal (2004) the sinuous instability of the jet originated from internal transverse shear across the jet. In Chapter 2, we showed that the time-averaged velocity field of quasi-two-dimensional jets could be mod- elled using two-dimensional plane jet theory. We also studied the instantaneous velocity field and revealed the interactions between the high-speed meandering core of the jet and the large eddies alternating on its sides. We showed that these core and eddy structures were self-similar with distance and continuously exchanged fluid between themselves, as well as with the ambient fluid surrounding the jet. In particular, the eddies played a key role in the entrainment of ambient fluid by means of engulfment at their rear. Entrained fluid could either be trapped for a brief period in an eddy, where it experienced strong mixing, or be directly incorporated in the core of the jet, where it was advected downstream much more 34 3.1 Introduction rapidly. We further hypothesized that because of the difference in advection speed between the core and the eddies (we measured that on average eddies travelled at approximately 1/4 of the speed of the core), initially relatively close fluid parcels entrained by the jet should experience large streamwise dispersion depending on whether they were drawn into the eddies or the core. In order to study and model the transport, mixing and dispersion of tracers in shallow river flows, we investigate in this chapter the temporal and spatial evo- lution of the concentration of tracers released in quasi-two-dimensional jets. The mixing properties of turbulent jets have been studied experimentally many times. Uberoi & Singh (1975) measured instantaneous temperature profiles in plane jets and found that they differed from typical time-averaged Gaussian profiles. They reported a relatively well-mixed interior while most of the mixing was performed at the turbulent–non-turbulent interface of the jet. Schefer et al. (1994) also noted a difference between the instantaneous distribution and the time-averaged distribution of tracers in the case of three-dimensional round turbulent jets. They attributed this discrepancy to the development of large-scale vortical structures. Arguably, the dynamics of large-scale vortical structures is different in quasi-two- dimensional jets from the case of three-dimensional round or plane jets due to the confinement of the flow in one direction (see Jirka, 2001, for a discussion on large- scale flow structures in shallow flows, or Chapter 2 for quasi-two-dimensional jets specifically). Nevertheless, large-scale vortical structures do have an influ- ence on the mixing and dilution properties of quasi-two-dimensional jets. Giger et al. (1991) reported that mixing efficiency and dilution in quasi-two-dimensio- nal jets tended to diminish with distance. From turbulence spectral analysis and intermittency analysis, Dracos et al. (1992) argued that the decrease of mixing efficiency was due to the development of quasi-two-dimensional turbulence. Us- ing laser-induced fluorescence in quasi-two-dimensional jets, Chen & Jirka (1999) showed that quasi-two-dimensional turbulence induced patchiness in the time- dependent distribution of the tracer concentration. They found distinct regions of large concentration which corresponded to the large-scale turbulent structures. Jirka (2001) reflected upon the impact of large vortical structures in shallow river flows and emphasized their ability to transport relatively unmixed fluid over large distances. Despite the large number of experimental studies, there appear to have been 35 3 Model for the streamwise transport, dispersion and mixing relatively few attempts to provide a comprehensive model of the advection and dispersion processes in quasi-two-dimensional jets. Moreover, most models as- sume a steady state. Paranthoe¨n et al. (1988) suggested a limited model for the initial phase of the dispersion process in a turbulent plane jet. Then, from con- servation of mass in a classical plane jet, Chen & Jirka (1999) showed that the decay of the time-averaged concentration of passive tracers C along the centreline of quasi-two-dimensional jets followed C ∝ z−1/2. Using conservation of mass and the Reynolds-averaged Navier–Stokes equation with the boundary-layer ap- proximation for three-dimensional round and plane jets, Law (2006) proposed an analytic solution for the time-averaged concentration distribution across the jet. To close the problem, he used the common assumption that the turbulent dif- fusive term was proportional to the gradient of the time-averaged concentration across the jet. He also assumed that the coefficient of proportionality between these two quantities (i.e. the turbulent diffusivity) was constant across the jet and depended only on the eddy diffusivity and the turbulent Schmidt number (see e.g. Mathieu & Scott, 2000, for more details). Previous models often assume purely lateral entrainment, and then simple time- averaged streamwise motion. Owing to the cross-stream variation in along-stream velocity (due to the time-dependent core–eddy interaction and the time-averaged Gaussian streamwise velocity distribution) quasi-two-dimensional jets inevitably have significant along-stream dispersion. We want to investigate the implications of this along-stream dispersion for tracer transport and how it affects advection in quasi-two-dimensional jets. In this chapter, we propose a new one-dimensional model solving the time- dependent effective advection–diffusion equation along the direction of the flow, based on mixing-length theory. Mixing-length theory is appropriate because of the central role of large eddies (scaling with the local jet width) on the dis- persion within the flow. We find analytical solutions in similarity form for the case of a constant-flux release and the case of a finite-volume release of trac- ers, which appear to describe correctly some new experimental measurements of tracer transport. We are able to formulate the general solution for any time- dependent release in integral form, effectively by means of a Green’s-function-like solution. We also show the importance of along-stream dispersion mechanisms in quasi-two-dimensional jets, by comparing our full effective advection–diffusion 36 3.2 Turbulent model hypothesis model with a simple advection model. In § 3.2, we present our model hypothesis starting from the advection–diffusion equation, where the diffusive term models the dispersion by the turbulent flow field of quasi-two-dimensional jets. In § 3.3, we derive analytical solutions for both a constant volume-flux release and an in- stantaneous finite-volume release. We also show how to generalize the analytical solution for an instantaneous finite-volume release into a solution for an arbitrary time-dependent release. In the next chapter, we compare the theoretical results obtained in this chapter with experimental data. In § 4.1, we describe our experi- mental procedure. In § 4.2, we first provide a qualitative assessment of our model hypothesis, then we compare our theoretical predictions with experimental data obtained using dye tracking experiments and virtual particle tracking experiments in both the constant-flux and the finite-volume cases. In § 4.3, we analyse the statistical significance of the experimental measurements presented in § 4.2 for the cases of constant-flux releases of dye and instantaneous finite-volume releases of virtual particles. Finally, in § 4.4 we draw our conclusions for both Chapters 3 and 4. 3.2 Turbulent model hypothesis To characterize the evolution of the concentration of tracers released in quasi- two-dimensional jets, we consider the ideal model of a turbulent momentum jet in a two-dimensional semi-infinite environment. Adopting the same conventions to those used in § 2.4, the flow is considered incompressible and statistically steady. The x-direction is the lateral, cross-jet direction and the z-direction is the streamwise, axial direction. Assuming a plane flow in the domain, the velocity is labelled u = (u, w) in a Cartesian coordinate system (x, z) with the origin at the nozzle exit. The temporal and spatial evolution of the concentration of tracers C(x, z, t) (where t is time) in a two-dimensional steady turbulent jet satisfies (see e.g. Itoˆ, 1992) ∂tC +∇ · (uC) = κ∆C, (3.1) where ∇ is the gradient operator, κ is the molecular diffusivity and ∆ is the Laplacian in two dimensions. We take a point-wise ensemble average (i.e. an 37 3 Model for the streamwise transport, dispersion and mixing ensemble average at each point (x, z, t) in space and time) of equation (3.1) ∂tCE +∇ · (uECE) +∇ · ([uFCF]E) = κ∆CE, (3.2) where the subscript in XE denotes the ensemble average of a quantity X and XF denotes the fluctuations such that X = XE + XF, [XF]E = 0. Thus, the ensemble-averaged concentration is defined as CE(x, z, t) = 1 N N∑ n=1 Cn(x, z, t), (3.3) where N is the total number of realisations of an experiment and n designates the nth realisation. We then make the modelling assumptions that uECE behaves as an advective contribution and is equal to λ1uCE (where the overbar represents an appropriate average in time and λ1 is a constant), while [uFCF]E effectively acts diffusively so that [uFCF]E = −D · ∇CE (with D a turbulent eddy diffusive tensor). We expect that advection is governed by the mean flow and dispersion by eddy processes. The term [uFCF]E can be seen as a turbulent flux, which is usually defined as u′C ′ with u′ = C ′ = 0 (Mathieu & Scott, 2000), where u′ and C ′ designate the temporal fluctuations of the velocity field and the temporal fluc- tuations of the concentration field, respectively. In other words, we assume that the statistical diffusive effect of the turbulent fluctuations is equivalent whether averaged in time or over many realisations. The diffusive effect of [uFCF]E de- scribes and parameterizes physically the interaction between the high-speed core and the growing eddies described in Chapter 2. Therefore, neglecting molecu- lar diffusion under the assumption that it is less significant than eddy diffusion processes (Mathieu & Scott, 2000), equation (3.2) becomes ∂tCE + λ1∇ · (uCE) = ∇ · (D · ∇CE) , (3.4) We believe that the interaction between the high-speed core and the growing eddies has a strong streamwise dispersive effect. On the other hand, the cross- jet distribution of the concentration remains confined laterally by two linearly- expanding straight-sided boundaries (as observed in Chapter 2). As we already mentioned, the transport and dispersion of tracers in quasi-two-dimensional jets is more critical along the streamwise direction (Jirka, 2001). Therefore, we choose 38 3.2 Turbulent model hypothesis to integrate equation (3.4) across the jet ∂tφ+ λ1 ∫ ∞ −∞ ( ∂x (uCE) + ∂z (wCE) ) dx = ∫ ∞ −∞ ∇ · (D · ∇CE) dx, (3.5) where φ(z, t) = ∫ ∞ −∞ CE(x, z, t) dx. (3.6) Since CE vanishes as x→ ±∞ and u remains finite we have ∂tφ+ λ1∂z (∫ ∞ −∞ wCE dx ) = ∫ ∞ −∞ ∇ · (D · ∇CE) dx. (3.7) We assume that the eddy diffusive coefficient is independent of x and that, in the streamwise direction, it scales like the local characteristic velocity wm(z) (the maximum time-averaged streamwise velocity in the jet at height z) and the local characteristic size b(z) (the velocity spread rate or e-folding distance of the time- averaged streamwise velocity at height z) of this core and eddy structure, such that Dzz(z) ∝ b(z)wm(z). (3.8) This is essentially a ‘mixing-length’ model (Prandtl, 1925), where the mixing length is the local characteristic width of the jet, and where streamwise transport and dispersion are dominant. Therefore, since ∂xCE and ∂zCE vanish as x → ∞ and D remains finite equation (3.7) becomes ∂tφ+ λ1∂z (∫ ∞ −∞ wCE dx ) ∝ ∂z (wmb ∂zφ) . (3.9) We found in (2.5a,b) b(z) = Q0 2 √ 2πM0 ( 4 √ 2αM0zQ02 + 1 ) and wm(z) = √ 2M0 Q0 ( 4 √ 2αM0zQ02 + 1 )−1/2 , (3.10a,b) where α is the entrainment coefficient (Morton et al., 1956), Q0 is the initial volume flux of the jet, and M0 is the initial momentum flux, which is conserved with distance in the z-direction (see figure 2.6). The time-averaged streamwise velocity can be further decomposed into a spatial-averaged part and a fluctuating 39 3 Model for the streamwise transport, dispersion and mixing part: w =< w > +wˆ, (3.11) where < w >= 1 2b (∫ ∞ −∞ w dx ) . (3.12) Therefore, we obtain ∂tφ+ λ1∂z ( < w > φ+ ∫ ∞ −∞ wˆCE dx ) ∝ ∂z (wmb ∂zφ) . (3.13) We again face a closure problem with the third term on the left-hand side of (3.13), which we address by assuming that this term has an advective effect of the form < w > φ. Therefore, considering that < w >∝ wm, we can introduce two constants ka and kd to obtain ∂tφ+ ka ∂z (wmφ) = kd ∂z (wmb ∂zφ) . (3.14) We can rewrite the quantities b and wm using the power laws (2.5a) and (2.5b) (neglecting the virtual origins) respectively, to obtain the effective advection– diffusion equation for the laterally-integrated ensemble-averaged concentration φ ∂tφ+KaM01/2 ∂z ( φ z1/2 ) = KdM01/2 ∂z ( z1/2∂zφ ) , (3.15) where the constants Ka and Kd are a dimensionless advection parameter and a dimensionless dispersion parameter, respectively, which we will determine exper- imentally. The parameters Ka and Kd can be related to ka and kd using (2.5a) and (2.5b) (and, again, neglecting the virtual origins) in the following manner Ka = ka ( 2α √ 2 )1/2 and Kd = 2kd ( α √ 2 π )1/2 , (3.16a,b) with α ≈ 0.068 (as calculated in Chapter 2). It is interesting to note that in (3.15) the dispersion term increases with distance like z1/2, whereas the advection term decreases with distance like z−1/2. 40 3.3 Mathematical model 3.3 Mathematical model In order to test our turbulent model hypothesis, we impose different, appropriate initial, boundary and integral conditions on solutions to the general effective advection–diffusion equation (3.15), for example, φ(z, 0) = 0, z > 0, φ(z, t) → 0 as z →∞, and ∫ ∞ 0 φ(z, t) dz ∝ tϑ, t > 0. (3.17a–c) Equation (3.17a) imposes that the concentration is 0 everywhere initially; equa- tion (3.17b) imposes that, at all time, the concentration vanishes at infinity; and equation (3.17c) imposes that, for t > 0, the total integrated concentration evolves as a power law of time. The integral condition (3.17c) effectively controls the release of the passive tracers in the jet. In this theoretical section, we solve analytically equation (3.15) for three dif- ferent sets of initial boundary and integral conditions. We consider the simple case of a constant-flux release of passive tracers (i.e. we impose ϑ = 1 in (3.17c)), which we solve by analysing either the concentration (see § 3.3.2) or the concen- tration flux (see § 3.3.3). In the second case, presented in § 3.3.4, we consider an instantaneous release of a finite volume of passive tracers at the origin of the jet (i.e. we impose ϑ = 0 in (3.17c)). Then, based on the solution for the instan- taneous finite-volume release, we show in § 3.3.5 how to formulate, in integral form, the solution for a general and more realistic time-dependent release of trac- ers governed by an arbitrary source function (i.e. not limited to a power law of time). We give an analytical solution in the case where the source function models a constant-flux release over a finite period of time T0. We further show that the solutions for the first two simpler cases of a constant-flux release and an instantaneous finite-volume release are the two asymptotic limits of the more general solution when T0 →∞ and t≫ T0, respectively. We choose to solve the problems of a constant-flux release and a finite-volume release because we can reproduce them experimentally, and thus test our turbulent model hypothesis and the various associated assumptions, stated in § 3.2, against experimental measurements (presented in § 4.2). Before deriving the solutions of the three cases, we use below a similarity transformation to simplify the partial differential equation (3.15) into an ordinary differential equation (ODE), which we can then solve. 41 3 Model for the streamwise transport, dispersion and mixing 3.3.1 Similarity transformation We introduce the dilation transformation zˇ = εaz, tˇ = εbt, φˇ = εcφ(ε−azˇ, ε−btˇ), (3.18) and so equation (3.15) becomes εb−c∂tˇφˇ+ ε 3 2a−cKaM01/2 ∂zˇ ( φˇ zˇ1/2 ) = ε 32a−cKdM01/2 ∂zˇ ( zˇ1/2∂zˇφˇ ) . (3.19) If b = 3a/2, then equation (3.15) is invariant under this transformation. This suggests that we look for a solution for (3.15) of the form φ(z, t) = t2c/3a y(η) with η = z t2/3M01/3 . (3.20) Thus (3.15) becomes ( 2c 3a − Ka 2η3/2 ) y + ( (2Ka −Kd) 2η1/2 − 2η 3 ) y′ −Kdη1/2y′′ = 0. (3.21) The general effective advection–diffusion problem has thus been simplified to the ODE (3.21). This second-order ODE, written in similarity form, captures both the temporal and spatial streamwise evolution of the concentration of tracers in quasi-two-dimensional steady turbulent jet. Most importantly, (3.21) allows not only for streamwise advection transport, but also for streamwise turbulent dispersion (based on a mixing-length assumption). Furthermore, we can note that (3.21) depends on the ratio of two dilation constants, c/a. This ratio can be determined using the integral condition (3.17c), which becomes, using (3.20), for t > 0, ∫ ∞ 0 φ(z, t) dz = t2c/3a ∫ ∞ 0 y ( z t2/3 ) t2/3M01/3dη = M01/3t( 2c 3a+ 2 3) ∫ ∞ 0 y(η) dη ∝ tϑ. (3.22) Therefore, this condition can hold for all t > 0 if and only if c a = 3ϑ− 2 2 . (3.23) 42 3.3 Mathematical model 3.3.2 Constant-flux release: concentration In the case of a release of tracers at a constant source flux F , if the general effective advection–diffusion equation (3.15) is satisfied for z > 0, t > 0 and if, in addition, φ(z, t) satisfies (following (3.17a–c) with ϑ = 1) φ(z, 0) = 0, z > 0, φ(z, t) → 0 as z →∞, and ∫ ∞ 0 φ(z, t) dz = Ft, t > 0, (3.24a–c) then the condition (3.24c) can hold for all t > 0 if and only if a = 2c according to (3.23) with ϑ = 1. Thus, (3.20) becomes φ(z, t) = t1/3y(η) with η = z t2/3M01/3 . (3.25) In this case, the initial boundary value problem for φ(z, t), defined by (3.21) with a = 2c, (3.24a–c) and (3.25), reduces to ( 1 3 − Ka 2η3/2 ) y + ( (2Ka −Kd) 2η1/2 − 2η 3 ) y′ −Kd η1/2y′′ = 0, (3.26) subject to the conditions y(η) → 0 as η →∞, ∫ ∞ 0 y(η) dη = F M01/3 , t > 0. (3.27a,b) Equation (3.26) can then be rewritten using y(η) = s 1 3 ( Ka Kd −1 ) p(s), with s = 4η 3/2 9Kd , (3.28) to obtain p′′ + p′ +   1 3 ( Ka Kd − 2 ) s + 1 4 − ( 1 3 ( Ka Kd − 12 ))2 s2   p = 0. (3.29) Making the change of variable p = e−s/2W , we obtain the Whittaker differen- tial equation (Gradshteyn & Ryzhik, 2007). The Whittaker functions Wk,m[s] and Mk,m[s] are two linearly independent solutions of the Whittaker differential 43 3 Model for the streamwise transport, dispersion and mixing equation where k = 1 3 (Ka Kd − 2 ) , m = 1 3 (Ka Kd − 1 2 ) . (3.30a,b) Therefore, the solution of (3.29) is p(s) = e−s/2 (JWWk,m + JMMk,m) [s], (3.31) where JW and JM are constants of integration which will be determined using the boundary conditions (3.27a,b). We can rewrite equation (3.31) in the similarity form y(η) = ( 4η3/2 9Kd ) 1 3 ( Ka Kd −1 ) e− 2η3/2 9Kd (JWWk,m + JMMk,m) [ 4η3/2 9Kd ] = JWW + JMM, (3.32) defining two linearly independent solutions: W (involving Wk,m), and M (in- volving Mk,m) of the underlying equation (3.31). Since m− k − 1/2 = 0, we can actually simplify the Whittaker functionsWk,m andMk,m (see equations (13.18.5) and (13.18.4) for Wm−1/2,m and Mm−1/2,m, respectively, in National Institute of Standards and Technology, 2011-08-29) to find W(η) = ( 4η3/2 9Kd )1/3 Γ [ 2 3 (Ka Kd − 1 2 ) , 4η 3/2 9Kd ] , (3.33) M(η) = 2 3 (Ka Kd − 1 2 )( 4η3/2 9Kd )1/3 γ [ 2 3 (Ka Kd − 1 2 ) , 4η 3/2 9Kd ] , (3.34) where Γ[g, ι] = ∫∞ ι hg−1e−h dh is the upper incomplete Gamma function and γ[g, ι] = ∫ ι 0 hg−1e−h dh is the lower incomplete Gamma function. We can prove that, as η →∞, W ∼ e−η3/2η ( Ka Kd − 32 ) , M∼ η1/2, (3.35a,b) (see equation (8.11.2) in National Institute of Standards and Technology, 2011- 08-29, for the asymptotic expansion of the upper incomplete Gamma function) for Ka > Kd/2 (we will find later that for our experimental data, Ka appears to be substantially greater than Kd). So, in order to satisfy the far-field boundary condition (3.27a) requiring decay of y, we must have JM = 0 with the solution depending on W alone. JW can then be determined using the boundary condition 44 3.3 Mathematical model (3.27b): JW = F M01/3 ∫ ∞ 0 W(η) dη . (3.36) Therefore, the general solution of the effective advection–diffusion problem for the case of a constant flux release at the source is, in similarity form, for Ka > Kd/2 yF (η) = 2F 3KdM01/3Γ [ 2 3 ( Ka Kd + 1 )]η1/2Γ [ 2 3 (Ka Kd − 1 2 ) , 4η 3/2 9Kd ] , (3.37) where Γ[g] = ∫∞ 0 hg−1e−h dh is the Gamma function. We can note that the laterally-integrated concentration φF (z, t) = t1/3yF (η) (according to equation (3.25) with yF described in (3.37)) tends towards a simple asymptotic distribution φF ∝ z1/2 as t2/3M01/3 ≫ z (or η ≪ 1). In the limit t2/3M01/3 ≫ z, it appears that the laterally-integrated concentration φF depends only on z and increases with distance like z1/2. On the other hand, we will see in the next chapter that the ensemble-averaged concentration CE,F (see (3.6)) should actually decrease like z−1/2, because the experimental cross-jet distribution of φF spreads linearly with distance (see figure 4.5a). Since the asymptotic distribution of the concentration CE,F is independent of time in the limit t2/3M01/3 ≫ z, this asymptotic distri- bution represents the steady state solution. This finding is in agreement with Chen & Jirka (1999), who also showed that the time-averaged concentration of passive tracers in quasi-two-dimensional jets decays like C ∝ z−1/2 along the jet axis. Note that in the steady-state case, the ensemble average is equivalent to the time average. Mathematically, the concentration CE,F is singular at the origin z = 0 and tends to infinity. However, this is not the case in practice because the concentration of tracers must be finite at the source and the jet has a virtual origin z0. Interestingly, in the purely advective limit where Kd → 0 (corresponding to a so-called ‘top-hat’ velocity profile, see e.g. Turner, 1986) equation (3.26) becomes ( 1 3 − Ka 2η3/2 ) y + ( Ka η1/2 − 2η 3 ) y′ = 0, (3.38) 45 3 Model for the streamwise transport, dispersion and mixing which integrates to yF,a(η) = { J1η1/2, 0 ≤ η < ηa J2η1/2, ηa < η , (3.39) where J1 and J2 are integration constants, and ηa = ( 3Ka 2 )2/3 (3.40) is the location of the advective front considering ‘top-hat’ velocity profiles in the jet. Using the boundary condition at infinity (3.27a), we obtain J2 = 0. J1 can be determined using the integral condition (3.27b). Therefore, the similarity solution of the purely advective problem for the case of a constant-flux release at the source is yF,a(η) =    F KaM01/3 η1/2, 0 ≤ η < ηa 0, ηa < η . (3.41) We have plotted in figure 3.1 the non-dimensional quantities yF/ ( F/M01/3 ) and yF,a/ ( F/M01/3 ) . The five different curves show the concentration profile in similarity form for different values of Ka and Kd. As we increase Ka (deter- mining the advection strength), the maximum of the curve is displaced upwards, further away from the origin, while if we increase Kd (determining the dispersion strength), the front drops less rapidly, and there is still asymmetry about the maximum. As expected, without dispersion (i.e. in the ‘top-hat’ limit Kd → 0) the distribution of tracers yF,a/ ( F/M01/3 ) has a discontinuity at ηa, the location of the advective front (defined in (3.40)), where it vanishes. To study the distribution of yF , we can compute the location of its centroid normalized with the advective front ηa µF = ∫ ∞ 0 yF (η)η dη ηa ∫ ∞ 0 yF (η) dη (3.42) = 3 5 ( 3Kd 2Ka )2/3 Γ [ 2Ka 3Kd + 43 ] Γ [ 2Ka 3Kd + 23 ] , (3.43) 46 3.3 Mathematical model η yF / ( F/M01/3 ) Ka = 1, Kd = 0.1 Ka = 1, Kd = 0 Ka = 1, Kd = 1 Ka = 10, Kd = 1 Ka = 10, Kd = 0 0 0.2 0.4 0.6 0.8 1 0 1 2 3 4 5 6 7 8 9 10 Figure 3.1: Variation of the non-dimensional similarity solution yF / ( F/M01/3 ) , de- fined in (3.37), against the similarity variable η = z/ ( t2/3M01/3 ) for the problem of advection–dispersion in the case of a constant flux at the source and for different values of the advection and dispersion parameters, Ka and Kd respectively. In the ‘top-hat’ limit Kd → 0, we use the non-dimensional similarity solution yF,a/ ( F/M01/3 ) defined in (3.41). and its standard deviation normalized with the advective front ηa σF =   ∫ ∞ 0 yF (η)η2 dη ηa2 ∫ ∞ 0 yF (η) dη − µF 2   1/2 (3.44) = ( 3Kd 2Ka )2/3   3Γ [ 2Ka 3Kd + 2 ] 7Γ [ 2Ka 3Kd + 23 ] −   3Γ [ 2Ka 3Kd + 43 ] 5Γ [ 2Ka 3Kd + 23 ]   2  1/2 . (3.45) We plot µF in figure 3.2(a). We can see that µF decreases when Ka/Kd increases. We can prove that µF → 3/5 as Ka/Kd → ∞ (using equation (5.11.7) in Na- tional Institute of Standards and Technology, 2011-08-29), thus meaning that the centroid of yF recedes behind the advective front at a fixed relative distance. The 47 3 Model for the streamwise transport, dispersion and mixing Ka/Kd µ F Theory Data µF = 3/5 0 5 10 15 20 25 0 0.5 1 1.5 2 (a) Ka/Kd σ F √ Theory Data σF = √ 12/175 0 5 10 15 20 25 0 0.5 1 1.5 2 (b) Figure 3.2: Constant-flux case for the tracer concentration: (a) plot of the theoreti- cally predicted variation of µF (defined in (3.43)), the centroid of yF (defined in (3.37)) normalized with the advective front ηa (defined in (3.40)), as a function of Ka/Kd (plot- ted with a solid line), with the experimentally determined value (obtained from the best fit of the constant-flux case shown in figure 4.7) marked with a cross, the asymptotic value of µF is plotted with a dashed line; (b) plot of the theoretically predicted variation of σF (defined in (3.45)), the standard deviation of yF normalized with the advective front ηa, as a function of Ka/Kd (plotted with a solid line), with the experimentally de- termined value (obtained from the best fit of the constant-flux case shown in figure 4.7) marked with a cross, the asymptotic value of σF is plotted with a dashed line. normalized standard deviation σF is plotted in figure 3.2(b). σF also decreases whenKa/Kd increases. We can prove that σF → √ 12/175 asKa/Kd →∞ (using equation (5.11.7) in National Institute of Standards and Technology, 2011-08-29). Moreover, we can observe in figure 3.1 that for the solution yF of the general effective advection–diffusion problem a non-negligible portion of the volume of tracers is transported faster than the advective speed due to the combined effects of advection and dispersion processes. We can compute the portion of the total volume of tracers βF which travels ahead of the advective front βF = ∫ ∞ ηa yF dη ∫ ∞ 0 yF dη , (3.46) using equation (3.37), we obtain βF = Γ [ 2 3 ( Ka Kd + 1 ) , 2Ka3Kd ] − ( 2Ka 3Kd ) Γ [ 2 3 ( Ka Kd − 12 ) , 2Ka3Kd ] Γ [ 2 3 ( Ka Kd + 1 )] . (3.47) 48 3.3 Mathematical model The ratio βF remains constant in time and space because (3.47) does not depend on η. Moreover, βF depends only on the ratioKa/Kd. We have plotted βF against Ka/Kd in figure 3.3(a). We can prove that βF tends asymptotically towards 0 at large Ka/Kd (see equation (8.11.10) in National Institute of Standards and Technology, 2011-08-29), thus meaning that the portion of tracers in the dispersive front becomes smaller as Ka/Kd increases (see figure 3.1 for the change in the distribution of yF with various Ka and Kd). We can also compute the normalized distance between the average location of the volume of tracers travelling ahead of the advective front and the location of the advective front ηa ξF = 1 ηa   ∫ ∞ ηa yF η dη ∫ ∞ ηa yF dη − ηa   , (3.48) which yields ξF = 3 5 ( 3Kd 2Ka )2/3 Γ [ 2 3 ( Ka Kd + 2 ) , 2Ka3Kd ] − ( 2Ka 3Kd )5/3 Γ [ 2 3 ( Ka Kd − 12 ) , 2Ka3Kd ] Γ [ 2 3 ( Ka Kd + 1 ) , 2Ka3Kd ] − ( 2Ka 3Kd ) Γ [ 2 3 ( Ka Kd − 12 ) , 2Ka3Kd ] − 1. (3.49) We plot ξF against Ka/Kd in figure 3.3(b). The distance ξF can be considered as the normalized distance between the dispersive front (average location of the par- ticles travelling ahead of the advective front) and the advective front ηa (defined in (3.40)). In time and space coordinates, the distance between the dispersive front zF and the advective front za is zF − za = ξFηat2/3. So the distance between the dispersive front and the advective front increases with time like t2/3. We can also see in figure 3.3(b) that ξF → 0 as Ka/Kd →∞, thus meaning that the front becomes sharper as Ka/Kd increases (see also figure 3.1). 3.3.3 Constant-flux release: concentration flux A somewhat more physically relevant quantity, which we can now study in space and time for the case of a constant-flux release at the source, is the streamwise 49 3 Model for the streamwise transport, dispersion and mixing Ka/Kd β F Theory Data 0 5 10 15 20 25 0 0.2 0.4 0.6 0.8 1 (a) Ka/Kd ξ F Theory Data 0 5 10 15 20 25 0 1 2 3 (b) Figure 3.3: Constant-flux case for the tracer concentration: (a) plot of the theoret- ically predicted variation of βF (defined in (3.47)), the portion of the total volume of tracers released which travels ahead of the advective front ηa (defined in (3.40)), as a function of Ka/Kd (plotted with a solid line), with the experimentally determined value (obtained from the best fit of the constant-flux case shown in figure 4.7) marked with a cross; (b) plot of the theoretically predicted variation of ξF (defined in (3.49)), the normalized distance between the average location of the volume of tracers travelling ahead of the advective front and the location of the advective front ηa, as a function of Ka/Kd (plotted with a solid line), with the experimentally determined value (obtained from the best fit of the constant-flux case shown in figure 4.7) marked with a cross. concentration flux of tracers in a steady quasi-two-dimensional jet, defined as: Mφ = ∫ ∞ −∞ wC dx. (3.50) We can take the point-wise ensemble average (as defined in (3.3) for the concen- tration) of (3.50) and, neglecting the second-order turbulent contribution to the flux (Wang & Law, 2002, found that the turbulent mass flux for round turbulent jets was approximately 7.6% of the mean mass flux, so can be ignored to leading order), we find MφE = ∫ ∞ −∞ wECE dx. (3.51) Using the same modelling assumptions we made in § 3.2, (3.51) becomes MφE = λ1 ∫ ∞ −∞ wCE dx, (3.52) where we assume that the ensemble-averaged streamwise velocity is proportional to the time-averaged streamwise velocity. Then, the time-averaged streamwise 50 3.3 Mathematical model velocity can be further decomposed into a spatially-averaged part < w >, defined in (3.12), and a fluctuating part wˆ, so that MφE = λ1 < w > φ+ λ1 ∫ ∞ −∞ wˆCE dx, (3.53) Again, if we assume that the term ∫∞ −∞ wˆCE dx in (3.53) has an advective effect similar to < w > φ, Mφ can be related to Ka and φ as (hereafter omitting the subscript E for simplicity) Mφ = KaM01/2 φ z1/2 . (3.54) Therefore, the solution of the concentration flux of tracers for the case of a con- stant source flux is, for Ka > Kd/2, Mφ(z, t) = yM(η) = F 1− Kd2Ka Γ [ 2 3 ( Ka Kd − 12 ) , 4η3/29Kd ] Γ [ 2 3 ( Ka Kd − 12 )] , with η = z t2/3M01/3 , (3.55) where we use the solution for the laterally-integrated concentration φ = φF = t1/3yF , with yF defined in (3.37). In the limit t2/3M01/3 ≫ z (or η ≪ 1), the concentration flux is independent of time or space and tends towards a constant Mφ → F (1−Kd/(2Ka)). For comparison with a purely advective flow, in the limit Kd → 0 (relevant, as already noted, to ‘top-hat’ velocity profiles) the concentration flux is yM,a(η) = { F, 0 ≤ η < ηa 0, ηa < η , (3.56) according to yF,a, defined in (3.41), and (3.54) with φ = φF,a = t1/3yF,a. We have plotted the normalized tracer flux yM/F as well as yM,a/F in figure 3.4. The five different curves show the concentration profile in similarity form for different values of Ka and Kd. As we increase the advection parameter the flux of tracers extends from the origin into a plateau before dropping smoothly at the front and eventually vanishing at large η. In the purely advective case (i.e. in the ‘top-hat’ limit Kd → 0), the solution yM,a/F has a discontinuity at the location of the advective front ηa (defined in (3.40)). The steepness of the front tends 51 3 Model for the streamwise transport, dispersion and mixing η yM/F Ka = 1, Kd = 0.1 Ka = 1, Kd = 0 Ka = 1, Kd = 1 Ka = 10, Kd = 1 Ka = 10, Kd = 0 0 0.5 1 1.5 0 1 2 3 4 5 6 7 8 9 10 Figure 3.4: Plot of the variation of the normalized similarity solution yM/F , defined in (3.55), against the similarity variable η = z/ ( t2/3M01/3 ) for the concentration flux of tracers in the case of a constant flux at the source F and for different values of the advection and dispersion parameters, Ka and Kd respectively. In the ‘top-hat’ limit Kd → 0, we use the normalized piecewise-constant similarity solution yM,a/F , defined in (3.56). to decrease with increasing dispersion parameter. Moreover, we can see that the value at the origin yM(η = 0)/F decreases with Ka/Kd, from yM(0)/F → ∞ as Ka/Kd → 0 to yM(0)/F → 1 as Ka/Kd →∞. Similarly to the previous section, we can compute the centroid of the distribu- tion of yM normalized with the advective front ηa µM = ∫ ∞ 0 yM(η)η dη ηa ∫ ∞ 0 yM(η) dη (3.57) = 1 2 ( 3Kd 2Ka )2/3 Γ [ 2Ka 3Kd + 1 ] Γ [ 2Ka 3Kd + 13 ] , (3.58) 52 3.3 Mathematical model and its standard deviation normalized with the advective front ηa σM =   ∫ ∞ 0 yM(η)η2 dη ηa2 ∫ ∞ 0 yM(η) dη − µM 2   1/2 (3.59) = ( 3Kd 2Ka )2/3   Γ [ 2Ka 3Kd + 53 ] 3Γ [ 2Ka 3Kd + 13 ] −   Γ [ 2Ka 3Kd + 1 ] 2Γ [ 2Ka 3Kd + 13 ]   2  1/2 . (3.60) We plot µM against Ka/Kd in figure 3.5(a). We can see that, similarly to µF (defined in (3.43)), µM decreases when Ka/Kd increases. We can prove that µM → 1/2 as Ka/Kd → ∞ (using equation (5.11.7) in National Institute of Standards and Technology, 2011-08-29), thus meaning that the centroid of yM recedes behind the advective front at a fixed relative distance. The normalized standard deviation σM is plotted in figure 3.5(b) against Ka/Kd. Similarly to σF (defined in (3.45)), σM decreases when Ka/Kd increases. We can prove that σM → √ 3/6 as Ka/Kd → ∞ (using equation (5.11.7) in National Institute of Standards and Technology, 2011-08-29). In a similar fashion to the previous subsection (cf. (3.46) and (3.48)), we can compute the portion of the total concentration flux of tracers βM which is ahead of the advective front ηa βM = ∫ ∞ ηa yM dη ∫ ∞ 0 yM dη , (3.61) using equation (3.55), we obtain βM = Γ [ 2 3 ( Ka Kd + 12 ) , 2Ka3Kd ] − ( 2Ka 3Kd )2/3 Γ [ 2 3 ( Ka Kd − 12 ) , 2Ka3Kd ] Γ [ 2 3 ( Ka Kd + 12 )] . (3.62) As before, the ratio βM remains constant in time and space because (3.62) does not depend on η; and βM depends only on the ratio Ka/Kd. We have plotted βM against Ka/Kd in figure 3.6(a). Similarly to βF , βM appears to vanish at large Ka/Kd, thus meaning that the portion of the tracer flux in the dispersive front becomes smaller as Ka/Kd increases (see figure 3.4 for the change in the 53 3 Model for the streamwise transport, dispersion and mixing Ka/Kd µ M Theory Data µM = 1/2 0 5 10 15 20 25 0 0.5 1 1.5 2 (a) Ka/Kd σ M Theory Data σM = √ 3/6 0 5 10 15 20 25 0 0.5 1 1.5 2 (b) Figure 3.5: Constant-flux case for the tracer concentration flux: (a) plot of the the- oretically predicted variation of µM (defined in (3.58)), the centroid of yM (defined in (3.55)) normalized with the advective front ηa (defined in (3.40)), as a function of Ka/Kd (plotted with a solid line), with the experimentally determined value (obtained from the best fit of the constant-flux case shown in figure 4.7) marked with a cross, the asymptotic value of µM is plotted with a dashed line; (b) plot of the theoretically predicted variation of σM (defined in (3.60)), the standard deviation of yM normalized with the advective front ηa, as a function of Ka/Kd (plotted with a solid line), with the experimentally determined value (obtained from the best fit of the constant-flux case shown in figure 4.7) marked with a cross, the asymptotic value of σM is plotted with a dashed line. distribution of yM with various Ka and Kd). We can also compute the normalized distance between the average location of the tracer flux ahead of the advective front and the location of the advective front ηa ξM = 1 ηa   ∫ ∞ ηa yM η dη ∫ ∞ ηa yM dη − ηa   , (3.63) which yields ξM = 1 2 ( 3Kd 2Ka )2/3 Γ [ 2 3 ( Ka Kd + 32 ) , 2Ka3Kd ] − ( 2Ka 3Kd )4/3 Γ [ 2 3 ( Ka Kd − 12 ) , 2Ka3Kd ] Γ [ 2 3 ( Ka Kd + 12 ) , 2Ka3Kd ] − ( 2Ka 3Kd )2/3 Γ [ 2 3 ( Ka Kd − 12 ) , 2Ka3Kd ] − 1. (3.64) We plot ξM against Ka/Kd in figure 3.6(b). Similarly to ξF (see figure 3.3b), we can also see in figure 3.6(b) that ξM → 0 as Ka/Kd →∞, thus meaning that the front becomes sharper as Ka/Kd increases. 54 3.3 Mathematical model Ka/Kd β M Theory Data 0 5 10 15 20 25 0 0.2 0.4 0.6 0.8 1 (a) Ka/Kd ξ M Theory Data 0 5 10 15 20 25 0 1 2 3 (b) Figure 3.6: Constant-flux case for the tracer flux: (a) plot of the theoretically pre- dicted variation of βM (defined in (3.62)), the portion of the total concentration flux of tracers ahead of the advective front ηa (defined in (3.40)), as a function of Ka/Kd (plotted with a solid line), with the experimentally determined value (obtained from the best fit of the constant-flux case shown in figure 4.7) marked with a cross; (b) plot of the theoretically predicted variation of ξM (defined in (3.64)), the normalized distance between the average location of the concentration flux of tracers ahead of the advective front and the location of the advective front ηa, as a function of Ka/Kd (plotted with a solid line), with the experimentally determined value (obtained from the best fit of the constant-flux case shown in figure 4.7) marked with a cross. 3.3.4 Finite-volume release: instantaneous release funda- mental solution We can also consider an instantaneous finite-volume release localized at the source of a quasi-two-dimensional steady turbulent jet. If the general equation (3.15) is satisfied for z > 0, t > 0 and if, in addition, φ(z, t) satisfies (following (3.17a–c) with ϑ = 0) φ(z, 0) = Bδ(z), φ(z, t) → 0 as z →∞, ∫ ∞ 0 φ(z, t) dz = B, t > 0, (3.65a–c) where B is a constant representing the total volume of tracers released and δ(z) is a Dirac delta function, then the condition (3.65c) can hold for all t > 0 if and only if c = −a according to (3.23) with ϑ = 0. Thus, (3.20) becomes φ(z, t) = t−2/3y(η) with η = z t2/3M01/3 . (3.66) 55 3 Model for the streamwise transport, dispersion and mixing In this case, the initial boundary value problem for φ(z, t), defined by (3.21) with c = −a, (3.65a–c) and (3.66), reduces to ( −2 3 − Ka 2η3/2 ) y + ( (2Ka −Kd) 2η1/2 − 2η 3 ) y′ −Kd η1/2y′′ = 0, (3.67) subject to the conditions y(η) → 0 as η →∞, ∫ ∞ 0 y(η) dη = B M01/3 , t > 0. (3.68a,b) Equation (3.67) can be rearranged −2 3 (ηy)′ +Ka ( y η1/2 )′ −Kd ( η1/2y′ )′ = 0, (3.69) and thus integrated twice to obtain y(η) = ηKa/Kd exp [ − 4 9Kd η3/2 ]( J4 + 2J3 3 ( − 4 9Kd ) 2 3 ( Ka Kd −1 ) γ [ 2 3 ( 1− KaKd ) ,− 4 9Kd η3/2 ]) , (3.70) where J3 and J4 are two integration constants and γ[g, ι] = ∫ ι 0 hg−1e−h dh is the lower incomplete gamma function. Since η > 0 and the function γ[g, ι] is complex for ι < 0, J3 must equal 0. J4 can be determined by integrating equation (3.70) ∫ ∞ 0 y(η) dη = J4 ( 3 2 ) 4 3 ( Ka Kd + 14 ) K 2 3 ( Ka Kd +1 ) d Γ [ 2 3 (Ka Kd + 1 )] , (3.71) and applying the integral condition (3.68b) to obtain J4 = B ( 3 2 ) 4 3 ( Ka Kd + 14 ) K 2 3 ( Ka Kd +1 ) d Γ [ 2 3 ( Ka Kd + 1 )] M01/3 . (3.72) Therefore, the ‘fundamental’ solution of the effective advection–diffusion problem for the case of an instantaneous finite-volume release initially localized as a delta 56 3.3 Mathematical model function at z = 0 is, in similarity form, yδ(η) = B ( 3 2 ) 4 3 ( Ka Kd + 14 ) K 2 3 ( Ka Kd +1 ) d Γ [ 2 3 ( Ka Kd + 1 )] M01/3 ηKa/Kd exp [ − 4 9Kd η3/2 ] . (3.73) We can note that the concentration φδ = t−2/3yδ(η) (from (3.66) with yδ described in (3.73)) vanishes in time for all values of η, because of the streamwise dispersion (i.e. for Kd > 0). Furthermore, we expect the actual concentration Cδ to vanish even more rapidly due to the cross-jet dispersion as the flow transports the finite volume of tracers (see the experimental results in figure 4.5a for finite-volume releases in quasi-two-dimensional jets). We have plotted the non-dimensional quantity yδ/ ( B/M01/3 ) in figure 3.7. The three different curves show the concentration profile in similarity form for different values of Ka and Kd. Unsurprisingly, we find that the location of the peak, ηmax = (3Ka/2)2/3, only depends on Ka. Thus, increasing Ka shifts the peak upwards (away from the origin), while increasing Kd spreads the width of the distribution. There is always to a greater or lesser extent asymmetry, with the leading edge being more diffuse than the rear. Interestingly, in the ‘top-hat’, purely advective limit Kd → 0 equation (3.69) integrates to ( Ka η1/2 − 2η 3 ) y = J5, (3.74) where J5 is a constant of integration. In order to satisfy the boundary condition at infinity (3.68a) as well as the integral condition (3.68b) we must have J5 = 0 for all 0 ≤ η < ηa and ηa < η, where ηa = (3Ka/2)2/3 is the location of the advective front as defined in (3.40) (note that ηa is the same in both the constant-flux case and the finite-volume case). Therefore, the similarity solution of the purely advective problem for the case of an instantaneous finite-volume release initially localized as a delta function at z = 0 is yδ,a = Bδ (η − ηa) . (3.75) As expected, without dispersion (i.e. in the ‘top-hat’ limit Kd → 0) the dis- tribution of tracers remains the same in time (i.e. distributed as the initial Dirac delta function). The delta function is located in the similarity domain at 57 3 Model for the streamwise transport, dispersion and mixing η yδ/ ( B/M01/3 ) Ka = 1, Kd = 0.1 Ka = 1, Kd = 1 Ka = 10, Kd = 1 0 0.2 0.4 0.6 0.8 1 0 2 4 6 8 10 12 Figure 3.7: Plot of the variation of the non-dimensional fundamental similar- ity solution yδ/ ( B/M01/3 ) , defined in (3.73), against the similarity variable η = z/ ( t2/3M01/3 ) for the problem of advection–dispersion in the case of an instantaneous finite-volume release at the source and for different values of the advection and disper- sion parameters, Ka and Kd respectively. ηa = (3Ka/2)2/3, the location of the (purely) advective front. In time and space coordinates, it means that the volume of tracers is located at za = (3Ka/2)2/3 t2/3 and travels at the speed wa = KaM01/2z−1/2 in the streamwise direction. We can notice that the location of the advective front ηa is the same as the location of the peak of the tracer concentration in the general effective advection–diffusion problem: ηa = ηmax = (3Ka/2)2/3. Similarly to the previous section, we can compute the centroid of the distribu- tion of yδ normalized with the advective front ηa µB = ∫ ∞ 0 yδ(η)η dη ηa ∫ ∞ 0 yδ(η) dη (3.76) 58 3.3 Mathematical model µB = ( 3Kd 2Ka )2/3 Γ [ 2Ka 3Kd + 43 ] Γ [ 2Ka 3Kd + 23 ] , (3.77) and its standard deviation normalized with the advective front ηa σB =   ∫ ∞ 0 yδ(η)η2 dη ηa2 ∫ ∞ 0 yδ(η) dη − µB2   1/2 (3.78) = ( 3Kd 2Ka )2/3   Γ [ 2Ka 3Kd + 2 ] Γ [ 2Ka 3Kd + 23 ] −   Γ [ 2Ka 3Kd + 43 ] Γ [ 2Ka 3Kd + 23 ]   2  1/2 . (3.79) We plot µB against Ka/Kd in figure 3.8(a). We can see that, similarly to µF (defined in (3.43)) and µM (defined in (3.58)), µB decreases when Ka/Kd in- creases. We can prove that µB → 1 as Ka/Kd → ∞ (using equation (5.11.7) in National Institute of Standards and Technology, 2011-08-29), thus meaning that the centroid of yB recedes precisely to the location of the advective front, which is also the location of the peak. It is also important to note that µB does not only depend on Ka but actually on the ratio Ka/Kd. Since the distribution is not sym- metric with respect to its centroid, then both advection and dispersion processes can affect the centroid. We believe that the underlying physical interpretation of this asymmetry can be related to the asymmetry between the advective and the dispersive terms in the general effective advection–diffusion equation (3.15). The advection term decreases with distance like z−1/2, whereas the diffusion term increases with distance like z1/2. The normalized standard deviation σB is plotted in figure 3.8(b) against Ka/Kd. Similarly to σF (defined in (3.45)) and σM (de- fined in (3.60)), σB decreases when Ka/Kd increases. We can prove that σB → 0 as Ka/Kd → ∞ (using equation (5.11.7) in National Institute of Standards and Technology, 2011-08-29). σB vanishes at large Ka/Kd because, as we mentioned previously, the concentration becomes distributed spatially according to a Dirac delta function δ(z). Similarly to the constant-flux case, in the general effective advection–diffusion problem a non-negligible portion of the volume of tracers is transported faster than the advective speed due to the combined effects of advection and dispersion 59 3 Model for the streamwise transport, dispersion and mixing Ka/Kd µ B Theory Data µB = 1 0 5 10 15 20 25 0 1 2 3 (a) Ka/Kd σ B Theory Data 0 5 10 15 20 25 0 1 2 3 (b) Figure 3.8: Instantaneous finite-volume case for the tracer concentration: (a) plot of the theoretically predicted variation of µB (defined in (3.77)), the centroid of yδ (defined in (3.73)) normalized with the advective front ηa (defined in (3.40)), as a function of Ka/Kd (plotted with a solid line), with the experimentally determined value (obtained from the best fit of the constant-flux case shown in figure 4.7) marked with a cross, the asymptotic value of µB is plotted with a dashed line; (b) plot of the theoretically predicted variation of σB (defined in (3.79)), the standard deviation of yδ normalized with the advective front ηa, as a function of Ka/Kd (plotted with a solid line), with the experimentally determined value (obtained from the best fit of the constant-flux case shown in figure 4.7) marked with a cross. processes. We can compute the portion of the total volume of tracers βB which travels ahead of the advective front βB = ∫ ∞ ηa yδ dη ∫ ∞ 0 yδ dη , (3.80) using equation (3.73), we obtain βB = Γ [ 2 3 ( Ka Kd + 1 ) , 2Ka3Kd ] Γ [ 2 3 ( Ka Kd + 1 )] , (3.81) where, once again, Γ[g, ι] = ∫∞ ι hg−1e−h dh is the upper incomplete Gamma func- tion. As in the constant-flux release case βF defined in (3.47), the ratio βB remains constant in time and space because (3.81) does not depend on η. Moreover, βB depends only on the ratio Ka/Kd. We have plotted βB against Ka/Kd in figure 3.9(a). However, in contrast to βF , we can prove that βB → 1/2 (plotted with a 60 3.3 Mathematical model dashed line) as Ka/Kd →∞ (see equation (8.11.10) in National Institute of Stan- dards and Technology, 2011-08-29), thus meaning that the distribution of tracers yδ becomes more symmetrical with respect to the peak value as Ka/Kd increases (see figure 3.7 for the change in the distribution of yδ with various Ka and Kd). We can also compute the normalized distance between the average location of the volume of tracers travelling ahead of the advective front and the location of the advective front ηa ξB = 1 ηa   ∫ ∞ ηa yδ η dη ∫ ∞ ηa yδ dη − ηa   , (3.82) which yields ξB = ( 3Kd 2Ka )2/3 Γ [ 2 3 ( Ka Kd + 2 ) , 2Ka3Kd ] Γ [ 2 3 ( Ka Kd + 1 ) , 2Ka3Kd ] − 1. (3.83) We plot ξB against Ka/Kd in figure 3.9(b). Similarly to the constant-flux case ξF defined in (3.49), the normalized distance ξB can also be considered as the distance between the dispersive front (average location of the particles travelling ahead of the advective front) and the advective front ηa. In time and space coordinates, the distance between the dispersive front zB and the advective front za is zB − za = ξBηat2/3. This distance increases with time as t2/3, as we observed in the constant- flux case. We can also see in figure 3.9(b) that ξB → 0 as Ka/Kd → ∞, thus meaning that the spreading of the tracer distribution becomes small compared with the distance between the peak and the origin as Ka/Kd increases (see also figure 3.7). 3.3.5 Finite-volume release: time-dependent release gen- eral solution The solution φδ(z, t) = t−2/3yδ(η) is the response of the system described by the effective advection–diffusion equation (3.15) to a finite volume released instanta- neously at t = 0 and distributed spatially according to a Dirac delta function δ(z). Due to the linearity of equation (3.15), we can construct from this ‘fundamental’ solution φδ an integral expression for the general solution φg for a finite volume B being released at the origin z = 0 over a period of time such that φg(0, t) = f(t). 61 3 Model for the streamwise transport, dispersion and mixing Ka/Kd β B β Theory Data βB = 0.5 0 5 10 15 20 25 0 0.2 0.4 0.6 0.8 1 (a) Ka/Kd ξ B Theory Data 0 5 10 15 20 25 0 1 2 3 (b) Figure 3.9: Finite-volume case for an instantaneous release: (a) plot of the theoret- ically predicted variation of βB (defined in (3.81)), the portion of the total volume of tracers released which travels ahead of the advective front ηa (defined in (3.40)), as a function of Ka/Kd (plotted with a solid line), with the experimentally determined value (obtained from the best fit of the constant-flux case shown in figure 4.7) marked with a cross, and the asymptotic value βB = 0.5 is plotted with a dashed line; (b) plot of the theoretically predicted variation of ξB (defined in (3.83)), the normalized distance between the average location of the volume of tracers travelling ahead of the advective front and the location of the advective front ηa, as a function of Ka/Kd (plotted with a solid line), with the experimentally determined value (obtained from the best fit of the constant-flux case shown in figure 4.7) marked with a cross. Without loss of generality, we choose to normalize the source function f(t): ∫ ∞ −∞ f(t) dt = 1. (3.84) Therefore, the general solution φg can be expressed as the following integral φg(z, t) = ∫ t 0 (t− τ)−2/3 yδ(ητ )f(τ) dτ, with ητ = z (t− τ)2/3M01/3 . (3.85) The case of a truly instantaneous release of a finite volume at (z, t) = (0, 0) is physically impossible to realize in an experiment. It is also not ideal in the modelling of real flows. A more realistic set of initial boundary conditions is to have a finite volume released at a constant flux over a finite period of time 0 ≤ t ≤ T0. This problem can be defined in terms of the following conditions φT0(z, t) → 0 as z →∞, ∫ ∞ 0 φT0(z, t) dz =    Bt T0 , 0 ≤ t ≤ T0 B, T0 < t , (3.86a,b) 62 3.3 Mathematical model with φT0 satisfying the general equation (3.15) for z > 0, t > 0. The solution to this initial boundary value problem can be computed using equation (3.85) with the source function fT0(t) = H(t)−H(t− T0) T0 , (3.87) where H is the Heaviside function (i.e. H(t) = 0 for all t < 0 and H(t) = 1 for all t > 0). We find that the solution to the integral (3.85) with the source function fT0 described by (3.87) is φT0(z, t) = 2Bz1/2 3KdM01/2T0Γ [ 2 3 ( Ka Kd + 1 )]  Γ [ 2 3 (Ka Kd − 1 2 ) , 4z 3/2 9KdM01/2t ] −    0, 0 < t ≤ T0 Γ [ 2 3 (Ka Kd − 1 2 ) , 4z 3/2 9KdM01/2(t− T0) ] , T0 < t   . (3.88) The upper incomplete Gamma function, Γ[g, ι] = ∫∞ ι hg−1e−h dh, requires g > 0, hence this solution is well-defined only for Ka > Kd/2. As we mentioned previ- ously, we will find later that for our experimental data Ka appears to be substan- tially greater than Kd. Note that this solution cannot be written in similarity form because of the dependence on the time constant T0. We can prove (see Appendix A.1) that the solution φT0(z, t), described in (3.88), satisfies φT0(z, t) = φδ(z, t), for t T0 ≫ 1. (3.89) So, the general solution for a rectangular source function converges asymptotically to the fundamental solution φδ(z, t) (defined by (3.73) and (3.66)) in the limit t ≫ T0. It is interesting to study how fast φT0 converges towards φδ. We can non-dimensionalize the distance z and the time t using the scalings for length and time scales T01/3M01/3 and T0, respectively, such that z = T01/3M01/3z˘, t = T0t˘, (3.90a,b) where breves denote non-dimensional variables. The evolution in time of the normalized absolute deviation of the general solution φT0 from the fundamental 63 3 Model for the streamwise transport, dispersion and mixing t/T0 de v Ka = 1, Kd = 0.1 Ka = 1, Kd = 1 Ka = 10, Kd = 1 0 5 10 15 20 25 30 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 Figure 3.10: Variation with scaled time t˘ = t/T0 of the normalized absolute deviation dev(t˘), defined in (3.91), of the general solution φT0(z, t), defined in (3.88), and the fundamental solution φδ(z, t) (defined by (3.73) and (3.66)). The data are computed numerically for different values of the advection and dispersion parameters. solution φδ is dev(t˘) = ∫∞ 0 |φT0(z˘, t˘)− φδ(z˘, t˘)| dz˘∫∞ 0 φδ(z˘, t˘) dz˘ , for tT0 ≥ 1, (3.91) a non-dimensional quantity which only depends on the advection and dispersion parameter Ka and Kd, and in particular does not depend on the total injected volume of tracers B, on the initial momentum M0 or on the period of injection T0. We plot dev(t˘) in figure 3.10 for 1 ≤ t/T0 ≤ 30. We compute the deviation numerically for three different sets of values of Ka and Kd. We can see that all the curves decrease asymptotically towards 0 as t/T0 increases. The deviation is smaller than 0.1 (which can be considered as a threshold value of near conver- gence) for t/T0 > 11, t/T0 > 4.7 and t/T0 > 11 for the sets of advection and dis- persion parameters (Ka = 1, Kd = 0.1), (Ka = 1, Kd = 1) and (Ka = 10, Kd = 1), respectively. It appears that the deviation depends mainly on the ratio Ka/Kd and only very weakly on Kd. Furthermore, we can note that in equation (3.88), if we take the limit T0 →∞ 64 3.3 Mathematical model and define F = B/T0, then we find φT0→∞(z, t) = t1/3yF (η), (3.92) with η = z/ ( t2/3M01/3 ) consistently with (3.25). So, equation (3.37) is equivalent to the asymptotic solution of the general solution φT0 if the period of release T0 extends to infinity. We have developed in this chapter a theoretical model, for various source con- ditions, describing the streamwise transport and dispersion in quasi-two-dimen- sional jets. In the following chapter we test the predictions of this model through comparison with a range of experimental measurements. 65 Chapter 4 Streamwise transport, dispersion and mixing in quasi-two-dimensional jets: experimental results 4.1 Experimental procedure We conduct our experiments in a slight modification of the experimental ap- paratus we presented in Chapter 2, as shown schematically in figure 4.1. We conduct three distinct sets of experiments using two qualitatively different tech- niques. Each set of experiments is designed to provide experimental data that can be compared with the three theoretical predictions derived in Chapter 3 for: a constant-flux release; an instantaneous finite-volume release; and a non- instantaneous finite-volume release. In the first set of experiments (whose results are presented in § 4.2.2), we measure the distribution of the concentration of dye 67 4 Experimental results for the streamwise dispersion and mixing x z d = 5mm Electronic valves Peristaltic pump Overflow H = 1 m L = 1m W = 0.01m u w 2b(z) CCD camera Blue back-light tape PIV study area Figure 4.1: Schematic diagram of the experimental apparatus. as it is released at a constant flux at the source of quasi-two-dimensional steady turbulent jets. The second set of experiments (whose results are presented in § 4.2.3) involves what we believe to be a new technique, which consists of track- ing large quantities of virtual particles evolving as passive tracers in the velocity field of quasi-two-dimensional steady turbulent jets. The velocity field is mea- sured in experiments with real jets (as opposed to numerically computed jets) by using particle image velocimetry. We designed this technique, which we desig- nate as virtual particle tracking, to obtain data for an instantaneous release to compare with our mathematical model (derived in Chapter 3). In the third set of experiments (whose results are presented in § 4.2.4), we measure the distribu- tion of the concentration of dye as it is released as finite volumes at the origin of quasi-two-dimensional steady turbulent jets. For physical reasons, which will be detailed below, we cannot release finite volumes of dye instantaneously in the jets, and so such physical dye releases inevitably extended over a finite time interval. 4.1.1 Constant-flux releases of dye We fill the 1m (L)×0.01m (W )×1m (H) tank displayed in figure 4.1 with fresh tap water. A vertical jet of constant source volume flow rate is discharged into the 68 4.1 Experimental procedure tank using a peristaltic pump (520DU/R2 Watson-Marlow variable speed pump) fed by a constant-head tank. The injection mechanism for the constant-flux releases of dye in steady turbu- lent quasi-two-dimensional jets consists of a syringe-pump connected to a small needle inserted into a single main tube. The needle is located 0.2 m upstream of the nozzle. After the jet has reached a steady state in the tank, a mixture of red food dye ‘Fiesta Red’ (Allura Red AC, E129) and tap water (with a dye concen- tration of 1.8% per weight) is injected at a constant flow rate, 0.11 cm3 s−1 . We study 19 constant-flux releases of dye in steady turbulent jets with jet Reynolds number 2240 ≤ Rej = dws/ν ≤ 3870, where ws is the source velocity and ν is the kinematic viscosity of water. To measure the dye concentration, we perform the experiments in a dark room. Following Dalziel et al. (2008) we attach a 0.54 m × 0.54 m electroluminescent Light Tape (Electro-LuminX Lighting Corporation) to the external surface of the rear side of the tank, centred on the jet axis and with the bottom of the tape at the height of the nozzle. It provides a constant and uniform source of near-monochromatic cyan light of approximately 400 cd m−2. This wave length is close to the peak of the ‘Fiesta Red’ dye absorption spectrum. We measure the transmitted light intensity with a high-speed 8 bit grey-scale camera (Fastcam SA1.1 - Photron) mounted with an 85 mm focal-length lens (f-stop 5.6). The camera is located 3 m away from the tank, which is sufficient to have negligible parallax error. We also take care to reduce any light pollution from reflection or other sources, in particular by installing a black frame around the study area. The camera records 640× 848 pixel images covering the entire study area, which spans −40 ≤ x/d ≤ 40 and 0 ≤ z/d ≤ 100 (where x is the coordinate in the lateral, cross-jet direction, and z is the coordinate in the streamwise direction; the origin is at the centre of the nozzle slot and d = 5 mm is the nozzle width), and part of the black frame (in order to have a black intensity reference). For each video we set the origin in time, t = 0, at the image preceding the first image in which dye is seen by the camera. The frequency of image acquisition is set at 60 frames per second. Following the calibration method and the algorithm described by Coomaraswamy (2011) and based on Cenedese & Dalziel (1998), we perform the calibration in situ. We record the intensity measured by the camera for 23 known concentrations of dye, ranging from 0 to 2 % per weight. A fitting 69 4 Experimental results for the streamwise dispersion and mixing curve using a third-order polynomial in the logarithm of the normalized intensity gives us a continuous and monotonic relationship between the intensity and the spanwise- (or y-) integrated concentration. All the images recorded by the camera, either for the calibration process or for the experiments, are analysed using the software code DigiFlow (Sveen & Dalziel, 2005). This procedure enables us to obtain accurate measurements of the (spanwise-integrated) dye concentration in time and space for each experiment. 4.1.2 Instantaneous finite-volume releases of clusters of vir- tual particles We track virtual particles in experimentally measured velocity fields of quasi- two-dimensional steady turbulent jets. We use the velocity fields measured by us previously as presented in Chapter 2 and obtained using a particle-image- velocimetry technique (as described in Sveen & Dalziel, 2005). We measure the jet velocity in a 0.4m×0.4m study area centred on the jet axis (as shown in figure 4.1) and covering a height from z = 0.2–0.6 m. We use the camera described above (mounted with a 62mm focal-length lens) at a frequency of image acquisition 250 frames per second and for a duration of 21.8 s. The 1024 × 1024 pixel images provide us with spatially and temporally resolved velocity fields for six steady turbulent jets at source volume flow rates 33.2, 37.0 and 40.3 cm3 s−1 . The jet Reynolds number ranges from 3320 ≤ Rej ≤ 4030. We find that the divergence of each velocity field is insignificant (typically mean(|∇·u|)/mean(|∇×u|) ≈ 5%, wheremean(·) represents an average in time and space), so they can be considered as incompressible. Using these computed velocity fields, we seed in each of them 201×51pixel clusters of (massless) virtual particles located in a rectangular evenly- distributed cluster at −8.8 ≤ x/d ≤ 7 and 44.4 ≤ z/d ≤ 48.3 (i.e. within the characteristic local width of the jet). The release can be considered instantaneous as a cluster of virtual particles is injected in the flow field within a single time step. The possibility of releasing instantaneously a large number of particles constitutes the main reason for the use of this technique in this study. This important advantage, compared with the non-instantaneous dye finite-volume releases (discussed below), allows us to reproduce more easily the instantaneous release constraint imposed in the mathematical model in (3.65a). We release individual clusters every 0.4s in each experiment and study a total of 70 4.1 Experimental procedure 256 clusters representing 2,624,256 virtual particles. For each cluster the virtual particles evolve in time and space as passive tracers transported by the flow. For each simulation we set the origin in time, t = 0, at the first image in which the particle cluster is seeded. The simulation of a cluster stops as soon as a virtual particle reaches the top boundary of the velocity field. Finally, we record the location in time and space of the tracers and analyse the results using DigiFlow. By averaging 256 virtual-particle experiments we obtain a smooth distribution of the particle concentration in time and space, which we compare with the dye experiments and the theoretical prediction in § 4.2. Different techniques involving particle tracking have been used to study disper- sion, mixing and transport in jets or other types of flows. In previous studies, the particles were either real and tracked by imaging analysis technique (see e.g. Yang et al., 2000; Sveen & Dalziel, 2005), or purely numerical and evolving in nu- merically resolved flows (see e.g. Dutkiewicz et al., 1993; Luo et al., 2006; Picano et al., 2010). However, we have not been able to find any mention in the literature of using virtual particles in the velocity field of real flows. This technique requires a spatially and temporally resolved computation of the velocity field, which can be done, for example, using a particle-image-velocimetry technique. We can then seed some (massless) virtual particles in the velocity field and track their trajec- tory as they are transported as passive tracers by the flow. The advantages of this technique are numerous: the resolution is only limited by the resolution of the acquisition of the velocity field; it is not restricted to the computation limitations encountered in full numerical simulations, but can be used for any laboratory experiments; a large quantity of virtual particles can be seeded instantaneously in the jet (thus satisfying, in our case, the constraint imposed in the theoretical model for an instantaneous finite-volume release); and their initial distribution can be completely arbitrary. 4.1.3 Finite-volume releases of dye The experimental procedure for the finite-volume releases of dye in steady turbu- lent quasi-two-dimensional jets is very similar to the experimental procedure for the constant-flux releases of dye (described in § 4.1.1). We fill the tank displayed in figure 4.1 with fresh tap water. A vertical jet of constant source volume flow rate is discharged into the tank using the same peristaltic pump described above 71 4 Experimental results for the streamwise dispersion and mixing and fed by a constant-head tank. For the injection mechanism of the finite-volume releases of dye, the main tube divides into two approximately 80 cm before the nozzle (see figure 4.1). The two tubes are recombined approximately 15 cm before the nozzle. Two valves located just before the recombining junction control the flow for each pipe separately. We monitor the valves to allow the flow to go through one section or the other exclusively. We open and close the valves electronically so that a steady jet flow is maintained in the tank before and after switching the valves. Although we observe a small perturbation (a pressure wave) in the tank we believe it does not perturb the experiment significantly. The purpose of this two-tube system is to release a finite volume of dye in a steady turbulent jet. The procedure for each experiment is as follows. We inject a 5 cm3 mixture of the same red food dye described above and tap water (with a dye concentration of 2 % per weight) into the closed tube approximately 5 mm upstream of the valve. Meanwhile, water flows at a constant source volume flow rate through the other tube to produce a turbulent jet in the tank. After the jet reaches a steady state, we switch the valves to redirect the whole flow into the section containing the red dye, thus releasing a finite volume of dye into the established jet. We conduct 26 finite-volume releases of dye in steady turbulent jets with jet Reynolds number 2170 ≤ Rej ≤ 4870. It is important to note that, although great care is taken during the experiments and different protocols have been tested, instantaneous finite-volume releases of dye cannot be achieved for practical reasons. We find that the time of injection, although relatively short (of the order of 0.5 s), cannot be considered as instantaneous, as we will discuss in § 4.2.4. We believe that the main reason for this injection delay is due to some Taylor dispersion (Taylor, 1953) of the dye as it is transported in the short section of tube (approximately 0.2 m long) leading to the tank. We perform the measurements of the dye concentration for the finite-volume releases using exactly the same technique as described for the constant-flux re- leases. From the transmitted light intensity recorded by the high-speed camera described above, we can compute the dye concentration in the study area, span- ning −40 ≤ x/d ≤ 40 and 0 ≤ z/d ≤ 100, at a frequency of 60 frames per second. We obtain accurate measurements of the dye concentration in time and space for each experiment. 72 4.2 Experimental results 4.2 Experimental results Similarly to (2.7a,d), we find that the natural scalings for length and time in our problem are d, the source width, and (d2/Q0), respectively. Therefore when con- sidering our experimental data we will always scale quantities with these scalings, i.e. z = dz˜, t = ( d2 Q0 ) t˜, (4.1a,b) where tildes denote non-dimensional variables. Although the initial momentum fluxM0 is also a natural scaling parameter in the theoretical model (see equations (2.5b), (3.15) and (3.20)), we do not use it as a scaling parameter in this section because we could not measure it directly in the experiments. Instead ofM0, we use the equivalent ratio Q02/d (in § 2.4, we found M0 ≈< M >= 0.55 ( Q02/d ) , where < M > is the space- and time-averaged momentum flux in quasi-two-dimensional jets). In particular, the non-dimensional similarity variable η = z/ ( t2/3M01/3 ) , defined in the model (see § 3.3.1), is replaced by ηexp = z/ ( t2/3 ( Q02/d )1/3) , so that ηexp η = (dM0 Q02 )1/3 ≈ 0.82. (4.2) This non-dimensionalization also affects slightly the advection and dispersion pa- rameters Ka and Kd, defined in the model (see § 3.2). As a consequence, the advection and dispersion parameters Ka,exp and Kd,exp, that we use in this sec- tion, are related to Ka and Kd such that Ka,exp Ka = Kd,exp Kd = (dM0 Q02 )1/2 ≈ 0.74. (4.3) We omit the subscript exp in ηexp, Ka,exp and Kd,exp hereafter in this section. To test our turbulent model hypothesis developed in § 3.2 and which led to the general effective advection–diffusion (3.15), we choose to compare the theoretical predictions, developed in § 3.3, first with experiments realized in the constant- flux case. The initial boundary and integral conditions (3.24a–c) imposed in the constant-flux case are simpler to satisfy experimentally than the initial boundary and integral conditions imposed in the finite-volume case (3.65a–c), which require an instantaneous release of finite volumes of tracers. Instantaneous finite-volume releases of virtual particles are then tested against the theoretical prediction, 73 4 Experimental results for the streamwise dispersion and mixing before studying the more challenging case of a non-instantaneous finite-volume release of dye. In each case, we are particularly interested in whether the natural scaling of the model z ∝ t2/3 agrees with the experimental results and, if so, we then estimate from the experimental data the two key parameters: the advection parameter Ka and the dispersion parameter Kd. Since the experiments in the constant-flux case are simpler to realize, we believe that the estimates of Ka and Kd measured in this case are more accurate than in the other two cases. There- fore, we consider the values of Ka and Kd measured in the constant-flux case as reference values, while the values measured in the other two cases are used to determine the confidence interval of Ka and Kd. Before presenting the quanti- tative experimental results, we give below a qualitative assessment of our tur- bulent model hypothesis and motivate the utility of the virtual-particle-tracking technique (described in § 4.1.2) in understanding the transport, dispersion and mixing properties of quasi-two-dimensional jets. 4.2.1 Qualitative assessment The purpose of this qualitative assessment is two-fold. Firstly, we want to study how the dynamical structure of steady turbulent quasi-two-dimensional jets affects their transport and dispersion properties. We have developed our turbulent model hypothesis, stated in § 3.2, from the qualitative understanding of these properties. Secondly, we use in this study a new technique to analyse the transport and dispersion properties of the jets, which we introduced in the previous section as virtual particle tracking. We give a qualitative overview of this technique, as well as some justifications and motivations for its use in a more systematic and rigorous approach to obtain quantitative results (which will be presented in § 4.2.3). As we discussed in Chapter 2, in the far-field of quasi-two-dimensional jets (i.e. z ≥ 20 d for W = 2 d Dracos et al., 1992), the core forms a high-speed undulating region, which grows on average in an expanding straight-sided triangular section. Outside the core we observe large counter-rotating eddies, which develop on al- ternate sides of the core and grow linearly with distance. Moreover, we showed in Chapter 2 that the core–eddy structure is self-similar with distance z. The characteristic sinuous core and the large growing eddies can be observed in fig- ure 4.2(a), which is an instantaneous grey-scale picture of a constant-flux release of dye in a steady-state quasi-two-dimensional jet with Rej = 3850 (shown five 74 4.2 Experimental results 0 20 40 60 80 100 120 140 160 z (d ) (a) (b) 0.0 200.0 400.0 600.0 800.0 1000.0 (c) (d) 0.0 200.0 400.0 600.0 800.0 1000.0(e) Figure 4.2: (a) Grey-scale picture of a dyed jet (Rej = 3850) rising in the tank. The average dye edges are plotted with black lines (half-spreading angle, < θdye >= 12.4◦, as measured in figure 2.5). (b) Passive tracers (Pliolite particles) shown as streaks in a typical jet (Rej = 4080). (c) Trajectories of the passive tracers shown in (b) and identified by imaging analysis (for a duration of 0.2 s). (d) Instantaneous velocity field (arrows) of the jet shown in (b). (e) Trajectories of virtual particles (for approximately 0.3 s) seeded at the same initial locations as the particles identified in (c) and evolving as passive tracers in the time-dependent velocity field shown in (d). seconds after injection; the average dye edges are plotted with black lines, half- spreading angle < θdye >= 12.4◦, as measured in figure 2.4). The instantaneous core–eddy structure can also be seen in figure 4.2(b). In figure 4.2(b), a superpo- sition of 50 images (i.e. for a duration of 0.2 s) of the filming of an experiment (see § 2.2), where passive tracers (0.23 mm Pliolite VTAC particles) were mixed with a quasi-two-dimensional jet (Rej = 4080), depicts the tracers as streaks, thus revealing the Eulerian structures in the flow (see discussion in § 2.5). We compute two different types of results from the experiment with passive tracers shown in figure 4.2(b). We can consider the tracers as Lagrangian parti- cles and track their trajectory in time using a particle tracking algorithm imple- mented in DigiFlow (Dalziel, 1992; Sveen & Dalziel, 2005). Figure 4.2(c) shows 75 4 Experimental results for the streamwise dispersion and mixing the trajectories identified by the algorithm, at the same time instant as the jet displayed in figure 4.2(b). Particles have been tracked for 50 images (i.e. for a duration of 0.2 s) and reveal very similar flow patterns to the streaks in figure 4.2(b). However, this technique has some limitations as the number of particles tracked for a certain time period decreases quickly with increasing time period. We have also very little control over the initial distribution of the particles (usu- ally spatially homogeneous), and cannot, for example, reproduce an instantaneous finite-volume release of these particles. To remedy these limitations, we have de- veloped a virtual-particle-tracking technique, which we presented in § 4.1.2. We seed in the velocity field (displayed in figure 4.2d) of the experimental jet shown in figure 4.2(b) some virtual particles in order to track their trajectory as they are advected as passive tracers by the flow. As a qualitative validation of this technique, we have seeded the virtual particles so that their initial distribution is identical to the initial distribution of the (real) particles identified in figure 4.2(c). The resulting trajectories of the virtual particles are plotted in figure 4.2(e) for a period of approximately 0.3 s. The trajectories of the virtual particles are very similar to the trajectories of the particles in figures 4.2(b) and 4.2(c), and thus reveal the same core–eddy structure. We believe that the virtual-particle-tracking technique can provide meaningful information about the transport and dispersion properties of quasi-two-dimensional jets. The schematic diagram displayed in figure 4.3(a) summarizes the structure of quasi-two-dimensional jets. The time-averaged mean picture of quasi-two-di- mensional jets is associated with a triangular shape encapsulating all the flow structures, while the time-dependent picture shows a sinuous core flanked by large growing eddies. We believe that the interaction between the core and the eddies results in large streamwise dispersion as the fluid experiences intense stretching at the interface between the core and the eddies. The eddies also play a crucial role in the entrainment and mixing of ambient fluid. From the observations of dyed jets such as the jet illustrated in figure 4.2(a), we find that fluid can be entrained from the ambient by the eddies and then either drawn within the eddies or incorporated into the core. We also believe that fluid can be exchanged between the eddies and the core. On the other hand, we have not observed any dyed fluid being detrained completely from the jet to the ambient. These processes can be revealed by applying the virtual-particle-tracking tech- 76 4.2 Experimental results nique to the core and the eddies of a quasi-two-dimensional jet. In the velocity field of the jet presented in figure 4.2(d) and reproduced in figure 4.3(b), we seed three clusters of virtual particles. The first cluster, composed of 3721 virtual particles, distributed in a square and initially seeded at the centre of an eddy is shown in light grey in figure 4.3(b). The second cluster, composed of 7381 virtual particles, distributed in a rectangle and initially seeded between the eddy and the core is shown in grey in figure 4.3(b). The last cluster, composed of 3721 virtual particles, distributed in a square and initially seeded in the core of the jet is shown in dark grey in figure 4.3(b). Figure 4.3(c) shows the typical trajectories of one single particle from each cluster. The particle locations are plotted every 0.02 s and each colour corresponds to a time period of 0.2 s (see colour scale). The particle starting in the eddy (plotted with pluses) moves slower than the other two particles and its trajectory forms two loops characteristic of the fact that it is transported within the eddy. The particle starting in the core (plotted with crosses) is transported quickly and has a slightly sinuous trajectory, which is characteristic of the transport within the core. On the other hand, the trajectory of the particle chosen approximately at the interface between the eddy and the core (see § 2.5 for a thorough discussion on the identification of the core and eddy structures) is often more complex (plotted with squares) and can be transported from the core to the eddy, or indeed from the eddy to the core. In the present case the particle starts in the core and then is drawn into the neighbouring eddy as the trajectory forms one loop. This is a simple illustration of the possible exchange of fluid parcels between the different structures. Figure 4.4 shows the simultaneous evolution in time of all the particles in the three clusters as they are passively transported by the jet velocity field shown in figure 4.3(b). Each colour corresponds to a particular time instant, starting from black and finishing with white and with a time step of 0.2 s between each colour (we use the same colour scale to that used in figure 4.3c). Again, we can clearly see that the virtual particles are transported much faster in the core of the jet (see figure 4.4c) than in the eddy (see figure 4.4a). On the other hand, mixing is more intense in the eddy than in the core. The cluster initially seeded in the eddy disintegrates very rapidly compared to the cluster initially seeded in the core. The cluster initially seeded between the eddy and the core (see figure 4.4b) experiences considerable stretching in the streamwise direction (its streamwise 77 4 Experimental results for the streamwise dispersion and mixing (a) (b) 0 t (s)1 2 3 (c) Figure 4.3: (a) Schematic diagram describing the structure of quasi-two-dimensional jets. (b) Instantaneous velocity field displayed in figure 4.2(d) with three rectangular clusters of virtual particles initially seeded: at the centre of an eddy (plotted in light grey); between the eddy and the core (plotted in grey); and in the core of the jet (plotted in dark grey). (c) Typical trajectories of three virtual particles evolving in the time-dependent velocity field shown in (b) and initially seeded: in an eddy (cluster outlined in light grey) (plotted with pluses); between the eddy and the core (cluster outlined in grey) (plotted with squares); and in the core (cluster outlined in dark grey) (plotted with crosses). The particle locations are plotted every 0.02 s and each colour corresponds to a time period of 0.2 s (see colour scale). maximum extent is ten times larger than its cross-stream maximum extent after a few time steps), owing to the shear layer at the interface between the core and the eddy. We can notice that some virtual particles are drawn into the eddy while others remain in the core. This emphasizes the time-dependent exchange of fluids between the core and the eddies pointed out above. We can also observe the delaying effect (with the colour scheme) of the eddies, in which tracers have a longer residency time than in the core. In Chapter 5, we investigate further the turbulent relative dispersion of the particle clusters presented in figure 4.4. When ensemble-averaged, we believe that the streamwise dispersive mecha- nisms revealed by the virtual particles in figure 4.4 can be modelled as an en- hanced dispersion coefficient, as stated in the turbulent hypothesis presented in § 3.2. The main assumption we make in equation (3.8), pertaining to the tur- bulent eddy diffusive coefficient (Dzz ∝ b wm, where Dzz is the streamwise com- ponent of the turbulent eddy diffusive tensor), can be physically justified from the study of both the structures and the velocity profile of quasi-two-dimensional jets (see figures 2.5, 2.7 and 2.13 for velocity measurements in quasi-two-dimen- 78 4.2 Experimental results (a) 0 t (s)1 2 3 (b) (c) Figure 4.4: Evolution in time of the virtual particles seeded in the velocity field shown in figure 4.3(b) as they are transported by the flow (each colour corresponds to a particular time instant): (a) cluster initially distributed at the centre of an eddy and shown in light grey in figure 4.3(b); (b) cluster initially distributed between the eddy and the core and shown in grey in figure 4.3(b); (c) cluster initially distributed in the core of the jet and shown in dark grey in figure 4.3(b). Each colour corresponds to a time period of 0.2 s, the colour scale shown at the bottom of (b) is the same to that used in figure 4.3(c). sional jets). The core–eddy structure is self-similar with height, thus the local characteristic size of the jet, b(z), appears as a relevant length-scale. Moreover, the local maximum time-averaged streamwise velocity is the second physically meaningful variable in the problem of dispersion, because all mixing and disper- sive mechanisms should scale like wm(z). In the rest of this section, we compare ensemble-averaged experimental results with the theoretical predictions found in § 3.3 and based on our turbulent model hypothesis. 79 4 Experimental results for the streamwise dispersion and mixing 4.2.2 Constant-flux releases of dye We present in figures 4.5(a–c) experimental results and theoretical predictions of constant-flux releases of dye in quasi-two-dimensional steady turbulent jets. The spatial distribution of the concentration C(x, z, t) is plotted using a colour scale (see colour scale at the top of figures 4.5a–c) at different non-dimensional times, 74 ≤ t˜ ≤ 374, to show the evolution of the dye concentration in the jet. In figure 4.5(a), we plot the ensemble-averaged concentration of the 19 experi- ments, which were conducted at different jet Reynolds number, 2240 ≤ Rej ≤ 3870 (see § 4.1.1). We also plot the average dye edges (half-spreading angle, < θdye >= 12.4◦) with thick white lines and the average boundaries of the core (half-spreading angle, 7◦ starting from z = 20 d) with thin white lines. We can observe some dispersion of the dye at the leading edge, which indicates the stream- wise dispersion discussed above. It is also apparent that the dye is transported first through the core (i.e. within the thin white lines) before mixing across the full width of the jet (i.e. filling the triangle delimited by the average dye edges shown with thick white lines). The characteristic sinuous instability of the core (clearly visible in figure 4.2a) does not appear in figure 4.5(a) because of the averaging process. Our model is inherently one-dimensional, and so obviously cannot predict the distribution of the concentration across the jet (i.e. in the x-direction). In order to be able to solve the partial differential equation (3.4), we integrate the concen- tration along the x-axis and study the evolution of φ(z, t) rather than C(x, z, t). We present the laterally-integrated experimental concentration φF,exp(z, t) in fig- ure 4.5(b) in normalized and re-distributed form using C(x, z, t) =    φF,exp(z, t) 2l(z) , −l(z) ≤ x ≤ l(z) 0, otherwise , (4.4) where l(z) = tan (< θdye >)(z − z0), for z ≥ 0 (4.5) is the local lateral distance between the average dye edges (plotted with thick white lines in figure 4.5a) and z0 is the space virtual origin defined below in (4.7a). Alongside in figure 4.5(c), we show the equivalent theoretical prediction 80 4.2 Experimental results t ~ = 74 (a) (b) (c) 0 1.8C (% wt) t ~ = 56 (d) (e) (f ) 0 2C (% wt) t ~ = 149 (a) (b) (c) t ~ = 108 (d) (e) (f ) t ~ = 224 (a) (b) (c) t ~ = 222 (d) (e) (f ) t ~ = 299 (a) (b) (c) t ~ = 332 (d) (e) (f ) t ~ = 374 (a) (b) (c) t ~ = 533 (d) (e) (f ) Figure 4.5: Distribution in space and non-dimensional time t˜ = t/(d2/Q0) of the concentration of dye (plotted using the two colour scales shown at the top for figures a–c and d–f, respectively) in the case of constant-flux releases (a–c) and finite-volume releases (d–f ) in quasi-two-dimensional jets for: (a) ensemble average of 19 experiments, the average dye edges are plotted with thick white lines (half-spreading angle, < θdye >= 12.4◦, as measured in figure 2.4) and the average boundaries of the core are plotted with thin white lines (half-spreading angle, 7◦ starting from z = 20 d, as measured in figure 2.12); (b) spatial lateral average of the distribution shown in (a) (defined in (4.4)); (c) theoretical prediction based on (3.37) and using Ka = 1.65 and Kd = 0.09; (d) ensemble average of 26 experiments, similarly to (a) the average dye edges are plotted with thick white lines and the average boundaries of the core are plotted with thin white lines; (e) spatial lateral average of the distribution shown in (d) (defined in (4.9)); (f ) theoretical prediction based on (3.88) using Ka = 1.65, Kd = 0.09 and T0 = 183 ( d2/Q0 ) . 81 4 Experimental results for the streamwise dispersion and mixing computed from equation (3.37) for yF (η), based on the assumption of a constant- flux release at the origin of the jet. To compute the theoretical prediction yF , we useKa = 1.65 andKd = 0.09 for the advection and dispersion parameters, respec- tively. These parameters are optimized by obtaining the best least-squares fit be- tween the experimental concentration yF,exp (i.e. the similarity form of φF,exp(z, t), transformed using (3.25)), and the theoretical prediction yF . Before plotting the theoretical prediction yF in figure 4.5(c), we transform yF into its physical form φF (z, t) using (3.25), then normalize it (similarly to φF,exp(z, t)) with the local dis- tance 2l(z) between the average dye edges, and finally re-distribute it uniformly, assuming a top-hat spatially-averaged profile, within these boundaries, i.e. C(x, z, t) =    φF (z, t) 2l(z) , −l(z) ≤ x ≤ l(z) 0, otherwise , (4.6) where l(z) is defined in (4.5). As we noted in § 3.3.2, we can see that the cross- stream distribution of the concentration spreads linearly with distance. Law (2006) modelled mathematically the cross-stream distribution of the concentra- tion of passive tracers in round and plane turbulent jets. He also found that the cross-stream distribution spreads linearly with distance. The model predicts that, in steady state, the laterally-integrated concentration φF increases like z1/2. How- ever, due to the cross-stream dispersion, the concentration C should decrease like z−1/2. We can actually see in figure 4.5(b,c) that the experimental and theoretical concentrations, respectively, decrease with distance. Comparing the data (figure 4.5b) with the theoretical prediction (figure 4.5c), we can see that the propagation of the front as well as its dispersion appear to have been correctly modelled (i.e. the scaling is correct), with only a small difference near the source. This mismatch is probably due to the zone of flow establishment of the jet (see e.g. Yannopoulos & Noutsopoulos, 1990). There is a necessary time and distance of adjustment before the experimental data can match the theoretical prediction, because the theoretical prediction is based around the assumption that the jet characteristic properties are given by the similarity power laws (2.5a,b). Giger et al. (1991) and Dracos et al. (1992) reported that the structure of qua- si-two-dimensional jets was different near the source, where three-dimensional effects were important. They found that the self-similar core and eddy structure 82 4.2 Experimental results (which is key in the dispersion mechanisms of the jet) only developed beyond approximately z ≥ 20 d (for the aspect ratio W/d = 2). Therefore, we might expect our model to be appropriate for z ≥ 20 d. We display in figure 4.6(a) the evolution in time of the non-dimensional inte- grated concentration of dye released in the experiments shown in figure 4.5(a). We can see that the experimental data (plotted with pluses) increase approxi- mately linearly in time (a linear fit is plotted with a black line). Therefore, the constant-flux integral condition (3.24c) assumed in the model is satisfied experi- mentally. We show in figure 4.6(b) the evolution in time of the distribution in similarity space of the normalized experimental data yF,exp, plotted for nine successive time periods in the range 2 ≤ t˜ ≤ 353. As we explained earlier, yF,exp is computed from the ensemble-averaged laterally-integrated experimental concentration for the constant-flux releases φF,exp using equation (3.25) at every instant in time t˜. We also use the following virtual origins in space (see equation (2.6)) and time: z0 = − Q02 4 √ 2αM0 , t0 = z0d Q0 . (4.7a,b) The space virtual origin z0 is simply the virtual origin of quasi-two-dimensional jets. The time virtual origin t0 represents the time needed to travel the distance |z0|, from the jet virtual origin to the nozzle, at the average source jet velocity Q0/d. We shift the origins in space and time from (z = 0, t = 0) (where z = 0 corresponds to the height of the nozzle and t = 0 corresponds to the time instant when the dye first appears from the nozzle) to (z0, t0) by applying the following transformation between the new and old coordinates znew = zold − z0, tnew = told − t0. (4.8a,b) For simplicity, we omit the subscripts new and old hereafter. In Chapter 2, we found α ≈ 0.068 and M0 ≈< M >= 0.55 ( Q02/d ) . So, the non-dimensional virtual origins in space and time are z˜0 = t˜0 ≈ −4.7. Except for the data in the time interval, 2 ≤ t˜ ≤ 118 (plotted with dashed curves), the data corresponding to the time interval, 118 ≤ t˜ ≤ 353 (plotted with thin solid curves), seem to have a similar distribution. The experimental concentration distribution converges rapidly, in time, towards an asymptotic profile in similarity space (y, η). We 83 4 Experimental results for the streamwise dispersion and mixing t/(d2/Q0) ∫ ∞ 0 φ F ,e x p (z ,t ) dz /d 2 Data Linear fit 0 100 200 300 0 0.01 0.02 0.03 0.04 (a) η yF,exp/ ( F ( d/Q02 )1/3) Time-averaged 2 ≤ t˜ ≤ 118 118 ≤ t˜ ≤ 353 0 0.5 1 0 1 2 3 (b) Figure 4.6: (a) Evolution in time of the non-dimensional integrated concentration of dye: the experimental data are plotted with pluses, a linear fit is plotted with a black line. (b) Evolution in time of the distribution of the normalized ensemble-averaged laterally-integrated experimental concentration shown in similarity form yF,exp in the case of constant-flux releases (plotted with dashed curves against the similarity variable η = z/ ( t2/3 ( Q02/d )1/3) for the time interval 2 ≤ t˜ ≤ 118 and with thin solid curves for the time interval 118 ≤ t˜ ≤ 353). The time-averaged data yF,exp, for 118 ≤ t˜ ≤ 353, are plotted with a thick solid curve. approximate this asymptotic distribution by the time-averaged distribution yF,exp for 118 ≤ t˜ ≤ 353 (plotted with a thick solid curve in figure 4.6b). The rapid convergence of the data in similarity space is very important because it means that the similarity scalings derived from the model, φF (z, t) = t1/3yF (η) (with η ∝ z/t2/3), are the appropriate scalings for this phenomenon. We can notice in figure 4.6(b) that near η = 0 the data are incomplete. Small values of η ∝ z/t2/3 are equivalent to small values of z compared with t2/3, or large values of t2/3 compared with z. The incomplete data near η = 0 are simply due to a lack of spatial resolution near the source and a finite time of observation in the experiments. We present the experimental data yF,exp in figure 4.7 (the ensemble average is plotted with pluses and the standard deviation, std, with dotted curves). We compute the best least-squares fit using the theoretical formula (3.37), where Ka and Kd are optimized under the constant-flux constraint (3.27b). The best fit (plotted with a solid curve) is found for Ka = 1.65 and Kd = 0.09. We can see that the model captures the main characteristics of the data. The concentration increases from zero at the origin (where the first derivative is infinite) to a peak value and then decreases smoothly at the front. The front of the curve agrees with 84 4.2 Experimental results η yF / ( F ( d/Q02 )1/3) Average data yF,exp std data Best fit yF : Ka = 1.65, Kd = 0.09 0 0.2 0.4 0.6 0.8 1 0 0.5 1 1.5 2 2.5 3 3.5 Figure 4.7: Constant-flux case, in similarity form: plots of the ensemble average (pluses) and standard deviation (std) (thin dotted curves) of the normalized exper- imental dye concentration yF,exp (pluses) and best least-squares fit using yF from (3.37) and with Ka = 1.65 and Kd = 0.09 (solid curve) against the similarity vari- able η = z/ ( t2/3 ( Q02/d )1/3). the theoretical fit, and so, the dispersion processes appears to have been correctly modelled. The rear of the experimental data appears slightly more linear than the theoretical prediction. This mismatch is probably due to the zone of flow establishment discussed above. The ratio between the advection parameter and the dispersion parameter is approximately Ka/Kd = 18.3. Using the advection parameter, we can compute theoretically the location of the advective front (considering ‘top-hat’ velocity profiles in the jet), ηa = 1.83, based on (3.40). We find that the position of the centroid relative to ηa is, for the experimental data, µF,exp = 0.65 (computed using (3.42)), which is close to the theoretical prediction µF = 0.62 (shown with a cross in figure 3.2a and computed using (3.43) with Ka/Kd = 18.3). The standard deviation of yF,exp is σF,exp = 0.29 (computed using (3.44)), which is almost identical to the theoretical prediction σF = 0.30 (shown with a cross in figure 3.2b and computed using (3.45) with Ka/Kd = 18.3). We can also 85 4 Experimental results for the streamwise dispersion and mixing measure from the experimental data the portion of the dye which travels ahead of the advective front βF,exp = 0.12 (computed using (3.46)), which is close to the theoretical prediction βF = 0.10 (shown with a cross in figure 3.3a) based on the ratio Ka/Kd = 18.3 and using (3.47). Thus, at each instant in time a non-negligible proportion of the total volume of tracers having been released travels ahead of the advective front ηa. Finally, we can also determine from the experimental data the normalized distance between the average location of the volume of tracers travelling ahead of the advective front and the location of the advective front ηa, ξF,exp = 0.16 (computed using (3.48)). This value is slightly larger than the theoretical prediction based on the ratio Ka/Kd = 18.3 and using (3.49), ξF = 0.13 (shown with a cross in figure 3.3b). ξF is a measure of the spread of the front of the distribution compared with the distance of the peak from the origin. All these agreements between the data yF,exp and the best least-squares fit yF suggest that our model can predict the shape of the concentration distribution of a finite-volume release of tracers in quasi-two-dimensional jets. We believe that the constant-flux experiments are the most straightforward experiments performed in this chapter. Therefore, the values of the advection and the dispersion parameters Ka = 1.65 and Kd = 0.09, respectively, found in this case will be used in the next cases as reference values. Furthermore, these results clearly reveal the importance of dispersion processes in the transport of passive tracers by quasi-two-dimensio- nal jets. As is clear in figure 4.7, the front of the distribution of the concentration in the similarity space (y, η) is not sharp but smooth due to dispersion. Were the transport of passive tracers by quasi-two-dimensional jets purely governed by advective processes alone, the distribution of the concentration in similarity space would drop much more rapidly at the front, as shown by the distributions of yF,a in figure 3.1 (plotted with a thin solid curve and a thin dashed curve). It is also important to note that more than 10% of the total volume of tracers released, at any time, propagates ahead of the advective front. We plot the normalized ensemble-averaged experimental results for the concen- tration flux of dye yM,exp/F in figure 4.8 with pluses, while the standard deviation of the data (std) is plotted with thin dotted curves. The experimental concentra- tion flux of dye Mφ,exp is computed using the expression (3.54) with Ka = 1.65 (as found above for the best fit of yF,exp in the constant-flux case, see figure 4.7) 86 4.2 Experimental results and the virtual origins described in (4.7a) and (4.7b). Then, according to (3.55), the similarity form is yM,exp = Mφ,exp. We compute the theoretical prediction yM (plotted with a solid curve) using the theoretical formula (3.55) with Ka = 1.65 and Kd = 0.09 (the reference values obtained in the constant-flux case for yF , see figure 4.7). We also compute the best least-squares fit yM,fit using the theoret- ical formula (3.55), where Ka,fit and Kd,fit are optimized. The best fit (plotted with a dashed curve) is found for Ka,fit = 1.55 and Kd,fit = 0.07. (The values of the advection and dispersion parameters for the best fit and the theoretical prediction are actually very similar.) The theoretical prediction matches with the data at the front, with the dispersion processes appearing to have been correctly modelled, but near the origin the data drop towards zero instead of remaining constant. The absence of a plateau near the origin in the experimental results is presumably due to the time and distance of adjustment before the experimen- tal data can match the theoretical prediction, which we mentioned previously as being associated with the zone of flow establishment. We find that the position of the centroid relative to ηa = 1.83 (computed using (3.40) with Ka = 1.65) is, for the experimental data, µM,exp = 0.58 (computed using (3.57)), which is somewhat larger than the theoretical prediction µM = 0.50 (shown with a cross in figure 3.5a and computed using (3.58) with Ka/Kd = 18.3). The standard deviation of yM,exp is σM,exp = 0.31 (computed using (3.59)), which is identical to the theoretical prediction σM = 0.31 (shown with a cross in figure 3.5b and computed using (3.60) with Ka/Kd = 18.3). We can also measure from the experimental data the proportion of the dye flux being ahead of the advective front βM,exp = 0.09 (computed using (3.61)), which is close to the theoretical prediction βM = 0.06 (shown with a cross in figure 3.6a) based on the ratio Ka/Kd = 18.3 and using (3.62). We can also determine from the experimental data the normalized distance between the average location of the volume of tracers travelling ahead of the advective front and the location of the advective front (considering ‘top-hat’ velocity profiles in the jet) ηa, ξM,exp = 0.19 (computed using (3.63)). This value is somewhat larger than the theoretical prediction based on the ratio Ka/Kd = 18.3 and using (3.64), ξM = 0.12 (shown with a cross in figure 3.6b). The study of the flux of dye in the constant-flux case also demonstrates the ability of the model to predict both advective and diffusive processes. It is clear 87 4 Experimental results for the streamwise dispersion and mixing η yM/F Average data yM,exp std data Theoretical prediction yM : Ka = 1.65, Kd = 0.09 Best fit yM,fit: Ka,fit = 1.55, Kd,fit = 0.07 0 0.5 1 1.5 0 0.5 1 1.5 2 2.5 3 3.5 Figure 4.8: Constant-flux case, in similarity form: plots of the ensemble average (pluses) and standard deviation (std) (thin dotted curves) of normalized experimen- tal dye flux yM,exp, theoretical prediction yM using (3.55) and with Ka = 1.65 and Kd = 0.09 (solid curve), and best least-squares fit yM,fit using (3.55) and with Ka,fit = 1.55 and Kd,fit = 0.07 (dashed curve) against the similarity variable η = z/ ( t2/3 ( Q02/d )1/3). from the observation of the front of the profile in figure 4.8 that quasi-two-dimen- sional jets diffuse tracers in a qualitatively different manner from the ‘top-hat’ purely advective case yM,a presented in figure 3.4 (plotted with a thin solid curve and a thin dashed curve). Moreover, we measure that approximately 10 % of the total concentration flux of tracers is located ahead of the advective front. 4.2.3 Instantaneous finite-volume releases of clusters of vir- tual particles We now compare our effective advection–diffusion model with experiments con- ducted using finite-volume releases of tracers in quasi-two-dimensional jets. The initial boundary and integral conditions imposed in the finite-volume case (3.65a– c) are more difficult to reproduce experimentally because they require an instan- taneous release. An instantaneous release is not physically possible in laboratory 88 4.2 Experimental results experiments (as we discussed in § 4.1.3), but it can be achieved using virtual particles. So, we first investigate the case of finite volumes of virtual particles released in the velocity field of quasi-two-dimensional jets, before analysing the more difficult problem of finite-volume releases of dye (presented in § 4.2.4). Figure 4.9(a) shows the dimensionless streamwise profile, at different times, of the laterally-integrated normalized concentration φv,exp ( z˜, t˜ ) /φmax(t˜) (where φmax(t˜) is the maximum value of φv,exp(z˜, t˜) in time, and z˜ = z/d and t˜ = t/ (d2/Q0) as defined in (4.1a) and (4.1b), respectively) of the ensemble average of 256 virtual-particle clusters released instantaneously, as finite volumes, in the experimental velocity field of quasi-two-dimensional jets with source volume flow rates 33.2, 37.0 and 40.3 cm3 s−1 (see § 4.1.2). At each time, we bin the data into 100 evenly-spaced intervals extending from the origin to the maximum stream- wise extent of the ensemble-averaged cluster. The thick dashed curve shows the location of the front zf of the ensemble-averaged cluster in time, which reaches the top boundary of the velocity field at approximately t˜ = 290, after the release time. The location of the front follows the expected power law zf ∝ t2/3, derived from (2.5b). As we can see, the ensemble-averaged cluster rapidly changes from an initial rectangular shape to a smoother rounded profile as it is advected by the jet. At early times t˜ ≤ 150, the dispersion of the particles appears to differ slightly between the front and the rear of the ensemble-averaged cluster. The front is sharper and drops more rapidly, while the rear has a longer tail. This is probably due to the fact that at the beginning most particles are advected quickly by the core of the jet, while the rest are trapped in the lateral eddies where they move more slowly (the time-averaged streamwise speed of an eddy is approximately 25% of the maximum speed of the core, as measured in § 2.5). However, at later times the cluster seems to spread more symmetrically between the front and the rear. We believe this is due to the continuous exchange of material between the core and the eddies. We apply the similarity transformation (3.66) to the ensemble-averaged exper- imental concentration φv,exp to obtain the similarity form yv,exp, normalized by the total volume of virtual particles Bv,exp = 2,624,256. We use the space vir- tual origin z0 defined in (4.7a). The time virtual origin cannot be the same as defined in the simple equation (4.7b) because the jet velocity is not constant be- tween the jet virtual origin z0 and the location of release of the virtual particles 89 4 Experimental results for the streamwise dispersion and mixing 0 50 100 150 200 250 300 350 t/(d2/Q0) 50 60 70 80 90 100 110 120 z/ d φv,exp(z˜, t˜)/φmax(t˜) Front (a) η yv,exp/ ( B ( d/Q02 )1/3) Time-averaged 48 ≤ t˜ ≤ 205 205 ≤ t˜ ≤ 401 0 0.5 1 1.5 2 2.5 0 1 2 3 4(b) Figure 4.9: (a) Streamwise distribution of the normalized laterally-integrated con- centration of virtual particles φv,exp(z, t)/φmax(t) (solid curves) at different non- dimensional times. The results have been averaged for 256 releases of identical clusters in the velocity fields of quasi-two-dimensional turbulent jets of source volume flow rates 33.2, 37.0 and 40.3 cm3 s−1 . The location of the front of the ensemble-averaged cluster is plotted versus time with a thick dashed curve. (b) Evolution in time of the dis- tribution in similarity form of the normalized ensemble-averaged laterally-integrated experimental concentration of virtual particles yv,exp (dashed curves for the time in- terval 48 ≤ t˜ ≤ 205 and with thin solid curves for the time intervals 205 ≤ t˜ ≤ 401) against the similarity variable η = z/ ( t2/3 ( Q02/d )1/3). The time-averaged data yv,exp, for 205 ≤ t˜ ≤ 401, are plotted with a thick solid curve. (i.e. 44.4 ≤ z/d ≤ 48.3). We determine the time virtual origin so that the loca- tion of the front in time z˜f (t˜) (plotted with a thick dashed curve in figure 4.9a) best fits (using a least-squares fit) a straight line in a log–log plot. We show in figure 4.9(b) the evolution in time of the distribution in similarity space of the nor- malized experimental data yv,exp, plotted for nine successive time periods in the range 48 ≤ t˜ ≤ 401. We can see that yv,exp seems to converge towards an asymp- totic distribution after 205 ≤ t˜ (the data for 48 ≤ t˜ ≤ 205 are plotted with dashed curves, while the data for 205 ≤ t˜ ≤ 401 are plotted with thin solid curves). We approximate the asymptotic distribution by the time-averaged distribution yv,exp for 205 ≤ t˜ ≤ 401 (plotted with a thick solid curve in figure 4.9b). Similarly to the constant-flux case, the convergence of these finite-volume data in similarity space implies that the similarity scalings derived from the model, φδ(z, t) = t−2/3yδ(η) (with η ∝ z/t2/3), are the appropriate scalings for this phenomenon. In figure 4.10, we compare the time-averaged ensemble-averaged virtual par- ticle data yv,exp (the ensemble average is plotted with crosses and the standard 90 4.2 Experimental results deviation, std, is plotted with dotted curves) with the theoretical prediction of the fundamental solution yδ (plotted with a solid curve), which assumes an instan- taneous release. We compute the theoretical prediction yδ using equation (3.73) with Ka = 1.65 and Kd = 0.09 (the reference values obtained in the constant- flux case for yF , see figure 4.7). We also compute the best least-squares fit yδ,fit (plotted with a dashed curve in figure 4.10) using (3.73), where Ka,fit and Kd,fit are optimized under the finite-volume constraint (3.68b). The best least-squares fit between yv,exp and yδ,fit is obtained for Ka,fit = 1.62 and Kd,fit = 0.09. Once again, these best-fit values are quite similar to the reference values. We can see that the model captures the main characteristics of the data. The concentration increases from zero at the origin (where the first and second deriva- tives also vanish) to a peak value and then decreases at the front, following the theoretical prediction yδ. The location of the peak of yv,exp (which is also the location of the advective front) is at ηa,exp = 1.83. Using the advection parameter Ka = 1.65, we can compute theoretically a very similar value ηa = 1.83, based on (3.40). We find that the position of the centroid relative to ηa is, for the ex- perimental data, µB,exp = 0.99 (computed using (3.76)), which is slightly smaller than the theoretical prediction µB = 1.03 (shown with a cross in figure 3.8a and computed using (3.77) with Ka/Kd = 18.3). Thus, the centroid is very close to the location of the concentration peak. The standard deviation of yv,exp is σB,exp = 0.17 (computed using (3.78)), which is close to the theoretical prediction σB = 0.19 (shown with a cross in figure 3.8b and computed using (3.79) with Ka/Kd = 18.3). We can also measure from the experimental data the portion of the virtual particles which travels ahead of the advective front βB,exp = 0.49 (com- puted using (3.80)), which is very close to the theoretical prediction βB = 0.54 (shown with a cross in figure 3.9a) based on the ratio Ka/Kd = 18.3 and using (3.81). (A value βB of 0.5 means that the virtual particles are symmetrically distributed with respect to the concentration peak.) Finally, we can also deter- mine from the experimental data the normalized distance between the average location of the volume of tracers travelling ahead of the advective front and the location of the advective front (considering ‘top-hat’ velocity profiles in the jet) ηa, ξB,exp = 0.13 (computed using (3.82)). This value is somewhat smaller than the theoretical prediction based on the ratio Ka/Kd = 18.3 and using (3.83), ξB = 0.17 (shown with a cross in figure 3.9b). ξB is a measure of the spread of 91 4 Experimental results for the streamwise dispersion and mixing the distribution compared with the distance of the peak from the origin. η y/ ( B ( d/Q02 )1/3) Average data yv,exp std data Theoretical prediction yδ: Ka = 1.65, Kd = 0.09 Best fit yδ,fit: Ka,fit = 1.62, Kd,fit = 0.09 0 0.5 1 1.5 0 0.5 1 1.5 2 2.5 3 3.5 Figure 4.10: Finite-volume case, instantaneous release, in similarity form: plots of the variation with similarity variable η = z/ ( t2/3 ( Q02/d )1/3) of the ensemble av- erage (pluses) and standard deviation (std) (thin dotted curves) of the normalized time-averaged experimental concentration of virtual particles yv,exp (pluses), theoreti- cal prediction yδ defined by (3.73) with Ka = 1.65 and Kd = 0.09 (solid curve), and best least-squares fit using yδ,fit defined by (3.73) with Ka,fit = 1.62 and Kd,fit = 0.09 (dashed curve). All these agreements between the data yv,exp and the theoretical prediction yδ, and between the advection and dispersion parameters of the constant-flux case and the finite-volume case, suggest that our model can predict the shape of the concentration distribution of an instantaneous finite-volume release of tracers in quasi-two-dimensional jets. Furthermore, it clearly reveals the importance of dis- persion processes in the transport of passive tracers by quasi-two-dimensional jets. As is clear in figure 4.9(b), the distribution of the concentration in the similarity space (y, η) converges in time towards a distribution with a finite width. Were the transport of passive tracers by quasi-two-dimensional jets purely governed by advective processes alone, the distribution of the concentration in similarity space would rather shrink towards a distribution of negligible width (similar to a Dirac delta function), even with a non-instantaneous release of tracers. It is 92 4.2 Experimental results also important to note that approximately half of the total volume of tracers in figure 4.10 travels ahead of the advective front, at a normalized averaged distance ξB ≈ 0.17 (defined in (3.83) with Ka/Kd = 18.3). 4.2.4 Finite-volume releases of dye We also present in figures 4.5(d–f ) experimental results and theoretical predictions of finite-volume releases of dye in quasi-two-dimensional steady turbulent jets. The spatial distribution of the concentration C(x, z, t) is plotted using a colour scale (see colour scale at the top of figures 4.5d–f ) at different non-dimensional times, 56 ≤ t˜ ≤ 533, to show the evolution of the patch of dye as it is ad- vected, mixed and dispersed by the jet. In figure 4.5(d) we plot the ensemble average of the 26 experiments, which were conducted at different jet Reynolds number, 2170 ≤ Rej ≤ 4870 (see § 4.1.3). We also plot the average dye edges (half-spreading angle, < θdye >= 12.4◦) with thick white lines and the average boundaries of the ‘core’ (half-spreading angle, 7◦ starting from z = 20d) with thin white lines. Similarly to the constant-flux results presented in figures 4.5(a–c), we can observe that the interaction between the core and the eddies, as described in § 2.5, results in large streamwise dispersion. As we explained earlier, we model this streamwise dispersion using an enhanced turbulent eddy diffusive coefficient Dzz ∝ bwm. We present the laterally-integrated experimental concentration φB,exp(z, t) in figure 4.5(e) in normalized and re-distributed form using C(x, z, t) =    φB,exp(z, t) 2l(z) , −l(z) ≤ x ≤ l(z) 0, otherwise , (4.9) where l(z) = tan (< θdye >) (z−z0), as defined in (4.5), is the local lateral distance between the average dye edges (plotted with thick white lines in figure 4.5d), and z0 is the space virtual origin defined in (4.7a). Alongside in figure 4.5(f ), we show the equivalent theoretical prediction φT0(z, t) computed from equation (3.88) and based on the assumption of a finite volume being released at a constant-flux during a finite period of time T0 = 183 (d2/Q0) (we discuss this value in more detail below). To compute φT0 , we use Ka = 1.65 and Kd = 0.09 for the advection and dispersion parameters, respectively (the reference values obtained in the constant- 93 4 Experimental results for the streamwise dispersion and mixing flux case for yF , see figure 4.7). Before plotting the theoretical prediction φT0 in figure 4.5(f ), we normalize it (similarly to φB,exp(z, t)) with the local distance 2l(z) between the average dye edges, and finally re-distribute it uniformly, assuming a top-hat spatially-averaged profile, within these boundaries, i.e. C(x, z, t) =    φT0(z, t) 2l(z) , −l(z) ≤ x ≤ l(z) 0, otherwise , (4.10) where l(z) is defined in (4.5). Although the comparison between the experimental data in figure 4.5(e) and the theoretical prediction in figure 4.5(f ) is not perfect at early times and near the origin (the theoretical concentration seems to travel slightly slower than the experimental concentration for t˜ ≤ 222), it improves at later times as the jet advects and diffuses the dye. As we mentioned above, this mismatch is probably due to the zone of flow establishment of the jet. There is a necessary time and distance of adjustment before the experimental data can match the theoretical prediction, because the theoretical prediction is based around the assumption that the jet characteristic properties are given by the similarity power laws (2.5a,b). In these experiments, we naturally are not able to release finite volumes of dye instantaneously. Aspects of the experimental dye release are revealed in figure 4.11(a), where we show the evolution in time of the integral of the dye con- centration over the whole domain ∫∞ 0 φB,exp(z, t) dz (plotted with pluses). These data represent the total volume of dye ‘seen’ by the imaging analysis in the win- dow frame −40 ≤ x/d ≤ 40 and 0 ≤ z/d ≤ 100. The dashed line indicates the time instant t˜90 = 183, when approximately 90 % of the total volume of dye has entered the tank. We can see that the total volume of dye increases almost steadily for t˜ ≤ t˜90. Then, the total volume of dye reaches a maximum at t˜ ≈ 290 before decreasing smoothly as the dye is transported outside the window frame. These data clearly show that the release of dye occurs over a finite period of time and not instantaneously. The effect of the spreading in time of the release of dye can also be seen in the evolution in time of the concentration distribution in the jets. In figure 4.11(b), we show the non-dimensional experimental concentration in similarity form yB,exp (computed from φB,exp using (3.66) at each instant in time). We normalize yB,exp 94 4.2 Experimental results t/(d2/Q0) ∫ ∞ 0 φ (z ,t ) dz Data Theory based on φδ Theory based on φT0 t˜90 0 200 400 600 800 0 0.02 0.04 0.06 0.08 (a) 800 η yB,exp/ ( B ( d/Q02 )1/3) 0 0.5 1 0 1 2 3 (b) Figure 4.11: (a) Evolution in time of the integral of the dye concentration over the whole domain for: the experimental data φB,exp (pluses); the theoretical prediction φδ (solid line), defined by (3.73); and the theoretical prediction φT0 (dotted line), defined by (3.88). The dashed line indicates the time instant t˜90 = 183 when approximately 90 % of the total integrated concentration of dye has entered the tank. (b) Plots of the variation with similarity variable η = z/ ( t2/3 ( Q02/d )) of the evolution in time of the normalized ensemble-averaged laterally-integrated experimental concentration plotted in similarity form, yB,exp (computed from φB,exp using (3.66)), in the case of finite-volume releases of dye. The data are plotted at 12 different time instants for 0 ≤ t˜ ≤ 979, with time increasing as the amplitude of the data increases. with the total injected volume B and plot it at 12 different instants in time for 0 ≤ t˜ ≤ 979, with time increasing as the amplitude of the data increases. The space and time virtual origins described in (4.7a) and (4.7b) are used to compute yB,exp. Ideally, if the dye were released instantaneously at the origin (as described in the integral and initial boundary conditions (3.65a–c)) all the curves should be identical and collapse on a single profile. Instead, we observe a gradual increase of the area under the curves. The data do not appear to have yet reached an asymptotic distribution. It can also be noticed that the curves at late times (for 290 ≤ t˜) are not plotted over the whole range 0 ≤ η ≤ 3.5, but stop at some values η < 3.5. These curves are incomplete because for 290 ≤ t˜, the front of the dye (located at the height zf ) has already moved outside the image frame, i.e. zf/d > 100, and thus we cannot visualize the full distribution of the dye in space. It is clear from both figures 4.11(a) and 4.11(b) that the release of the dye is not instantaneous and that the data have not yet reached an asymptotic distribution in similarity space. Thus, we cannot use the theoretical prediction yδ defined in (3.73) to model these experiments (as we did in the case of finite-volume releases 95 4 Experimental results for the streamwise dispersion and mixing of virtual particles presented above) because the fundamental solution yδ assumes an instantaneous release of the finite volume of tracers (see the integral and initial boundary conditions (3.65a–c)). Therefore, we compare the experimental data φB,exp(z, t) with the general solution φg(z, t), described in (3.85) and based on the convolution of the fundamental solution φδ with a source function f(t) = φg(0, t). The source function can model the more general and realistic case of a time- dependent release. To compute the general solution φg(z, t), we need to define the source function f(t), which represents the rate at which the overall integrated volume of tracers changes with time. In figure 4.11(a), we observe that the total integrated con- centration of dye ∫∞ 0 φB,exp(z, t)dz increases almost linearly with time for t˜ ≤ t˜90. Hence, we choose to model the source function as simply a non-zero constant for 0 ≤ t˜ ≤ t˜90 and zero for t˜90 ≤ t˜, ft˜90(t˜) = H(t˜)−H(t˜− t˜90) t˜90 , (4.11) where H is the Heaviside function. Using such a rectangular source function, the general solution φg(z, t) corresponds to the particular solution φT0 (with T0 = t90), described in (3.88). We plot the resulting theoretical integrated concentration∫∞ 0 φT0(z, t) dz with a dotted curve in figure 4.11(a). We can see that the match with the data (plotted with pluses) is, at least until the dye is advected beyond the spatial range of the camera (for t˜ ≤ 290), better than for the model assuming an instantaneous release φδ (plotted with a solid line). We compute the theoretical prediction φT0 , based on the source function ft˜90 with T0 = t90 = 183(d2/Q0), using the virtual origins described in (4.7a,b). We compare the distribution of the experimental data φB,exp (plotted with pluses) and the theoretical prediction φT0 (plotted with solid curves) in figure 4.12 at nine different times for 0 ≤ t˜ ≤ 418. We compute φT0 using the advection and dispersion parametersKa = 1.65 andKd = 0.09, respectively (the reference values obtained in the constant-flux case for yF , see figure 4.7). We also show the best least-squares fit φT0,fit (plotted with dashed curves in figure 4.12), computed using the theoretical formula (3.88) and the source function ft˜90 (see equation (4.11)) with T0 = t90 = 183(d2/Q0). The advection and dispersion parameters Ka,fit and Kd,fit, respectively, are optimized under the finite-volume constraint (3.65c). 96 4.2 Experimental results The best least-squares fit between φB,exp and φT0,fit is obtained for Ka,fit = 1.75 and Kd,fit = 0.09, still quite close to the reference values. Overall, we observe a reasonable agreement between φT0 and φB,exp. At early time, for t˜ ≤ 100, the match between the data and the model is not perfect because the experimental concentration profile adjusts partially due to the lack of self-similarity in the jet (this issue is related to the zone of flow establishment discussed previously). Then, both the advection (location of the peak in time) and the dispersion (width of the curve) seem to agree. There is a consistent mismatch at the rear where the data seem to be more spread out. This is probably due to some residue of dye in the tube still being injected in the jet at late time, and apparently stretching and diffusing the experimental dye concentration. According to equation (3.89), the solution φT0(z, t) converges in time towards φδ(z, t). Hence, we also expect the data φB,exp(z, t) to converge in time towards φδ(z, t). We demonstrate this convergence by plotting in figure 4.13 the similarity form of the theoretical prediction φT0 at t˜ = 150 (plotted with a thin solid curve), t˜ = 300 (plotted with a dotted curve) and t˜ = 450 (plotted with a dashed curve). We also show the asymptotic solution yδ, defined by (3.73) and computed using the (reference) advection and dispersion parameters Ka = 1.65 and Kd = 0.09. We can measure the absolute deviation, based on equation (3.91), between yT0 at t˜ = 300 (when the integrated concentration of dye is approximately maxi- mum, see figure 4.11a) and the asymptotic solution yδ. We find dev = 0.85, computed for t˜/T˜0 = 300/183 ≈ 1.64 and using (3.91). If we consider that con- vergence is ‘achieved’ if dev ≤ 0.1, then we find that our experimental data would be expected to achieve convergence for t˜/T˜0 ≥ 13.6, or at t˜ ≥ 2488. We can estimate that the distance at which we should observe the concentration distri- bution of the finite volumes of dye converge towards an asymptotic distribution is z ≥ ηa(13.6T˜0)2/3 d ≈ 2 m (based on the location of the concentration peak at convergence). Finally, we can predict the key characteristics of yδ, the asymptotic distribution of yB,exp (the similarity form of φB,exp computed using (3.66)), which are actually identical to the characteristics of the theoretical prediction found for the virtual particles because the advection and dispersion parameters are the same. So, we can expect that the maximum concentration of the asymptotic distribution of yB,exp is located at ηa = 1.83, based on (3.40) with Ka = 1.65. The position of the centroid relative to ηa is µB = 1.03 (shown with a cross in 97 4 Experimental results for the streamwise dispersion and mixing z/ d 1000φ/d t/(d2/Q0) = 0 Data Theoretical prediction Best fit 0 0.5 1 0 20 40 60 80 100 z/ d 1000φ/d t/(d2/Q0) = 51 0 0.5 1 0 20 40 60 80 100 z/ d 1000φ/d t/(d2/Q0) = 103 0 0.5 1 0 20 40 60 80 100 z/ d 1000φ/d t/(d2/Q0) = 155 0 0.5 1 0 20 40 60 80 100 z/ d 1000φ/d t/(d2/Q0) = 207 0 0.5 1 0 20 40 60 80 100 z/ d 1000φ/d t/(d2/Q0) = 259 0 0.5 1 0 20 40 60 80 100 z/ d 1000φ/d t/(d2/Q0) = 311 0 0.5 1 0 20 40 60 80 100 z/ d 1000φ/d t/(d2/Q0) = 363 0 0.5 1 0 20 40 60 80 100 z/ d 1000φ/d t/(d2/Q0) = 418 0 0.5 1 0 20 40 60 80 100 Figure 4.12: Plots at various times of the streamwise distribution of the laterally- integrated concentration of dye against the non-dimensional distance z/d in the case of finite-volume releases for: ensemble-averaged experimental data φB,exp (pluses); theo- retical prediction φT0 (solid curves), based on equation (3.88) using the reference advec- tion and dispersion parameters Ka = 1.65 and Kd = 0.09, respectively, and the source function ft˜90(t) as defined in (4.11); and best least-squares fit φT0,fit (dashed curves), based on equation (3.88) using the advection and dispersion parameters Ka,fit = 1.75 and Kd,fit = 0.09, respectively, and the source function ft˜90(t) as defined in (4.11). figure 3.8a and computed using (3.77) with Ka/Kd = 18.3). The theoretically predicted standard deviation is σB = 0.19 (shown with a cross in figure 3.8b and computed using (3.79) with Ka/Kd = 18.3). The portion of the virtual particles which travels ahead of the advective front is βB = 0.54 (shown with a cross in figure 3.9a), based on the ratio Ka/Kd = 18.3 and using (3.81). The average lo- cation of the volume of tracers travelling ahead of the advective front is ξB = 0.17 (shown with a cross in figure 3.9b), based on the ratio Ka/Kd = 18.3 and using (3.83). 98 4.3 Statistical significance of the experimental results η y/ ( B ( d/Q02 )1/3) Theoretical prediction yT0 : t˜ = 150 Theory prediction yT0 : t˜ = 300 Theory prediction yT0 : t˜ = 450 Theory prediction yT0 : t˜ → +∞ 0 0.5 1 1.5 0 0.5 1 1.5 2 2.5 3 3.5 Figure 4.13: Finite-volume case, in similarity form: plots of the variation with simi- larity variable η = z/ ( t2/3 ( Q02/d )) of the non-dimensionalized theoretical prediction yT0 , based on equation (3.88), computed at t˜ = 150 (thin solid curve), t˜ = 300 (dot- ted curve), and t˜ = 450 (dashed curve), using the reference advection and dispersion parameters Ka = 1.65 and Kd = 0.09, respectively, and the source function ft˜90(t) as defined in (4.11). The asymptotic distribution of yT0 (thick solid curve) is equal to yδ and can be computed using (3.73) with Ka = 1.65 and Kd = 0.09. 4.3 Statistical significance of the experimental re- sults We investigate the statistical significance of the experimental results, presented in § 4.2 above, for the constant-flux releases of dye and the instantaneous finite- volume releases of virtual particles in quasi-two-dimensional jets. We compute the probability density function (p.d.f.) of all the measurements of the laterally- integrated concentration y (in similarity form) at different values of the similarity variable η ∝ z/t2/3. We do not compute the p.d.f. of the experimental results found in the case of finite-time finite-volume releases of dye. (We describe the experiments in § 4.1.3 and plot the ensemble-averaged concentration in figure 4.12 with pluses). As we discuss in § 4.2.4, the distribution of the concentration has not yet reached an 99 4 Experimental results for the streamwise dispersion and mixing asymptotic distribution in similarity form. Therefore, the study of the statistical significance is not meaningful while the concentration profile is in a transient time-dependent regime. In the problem of river pollution, predicting and assessing the risk of encoun- tering harmful concentration levels is crucial. We show in this section how we can predict and assess this risk for the constant-flux and instantaneous finite-volume cases. Moreover, we discuss how this risk varies in time and space, depending on the p.d.f. of the measurements of the concentration. 4.3.1 Constant-flux release of dye We compute numerically the p.d.f. fF of the laterally-integrated experimental concentration yF,exp (in similarity form) for the Ne = 19 constant-flux release experiments (presented in § 4.1.1) such that, for all 1 ≤ l ≤ Nη and 1 ≤ m ≤ Ny (with l and m two integers) fF (ym, ηl) = Ne,Nz ,Nt∑ i,j,k ∆ei,zj ,tk(ym, ηl) Ny∑ m Ne,Nz ,Nt∑ i,j,k ∆ei,zj ,tk(ym, ηl) δy , (4.12) where δy = y2 − y1 is the concentration step and, for all 1 ≤ i ≤ Ne, 1 ≤ j ≤ Nz and 1 ≤ k ≤ Nt (with i, j and k three integers), ∆ei,zj ,tk(ym, ηl) =    1 if ym−1 ≤ yF,exp(ei, zj , tk) F ( d/Q02(ei) )1/3 < ym and ηl−1 ≤ η = zj tk2/3 ( Q02(ei)/d )1/3 < ηl, 0 otherwise , (4.13) where zj (the discretized streamwise coordinate) are linearly distributed from approximately 0 to 100 d (depending on the experiment and with Nz = 756), tk (the kth frame of the experiment) are linearly distributed from approximately 0 to tNt (with Nt of the order of 300, depending on the experiment), ηl (the discretized similarity variable) are linearly distributed from 0 to 3.5 (with Nη = 200, the number of bins) and ym (the discretized laterally-integrated concentration 100 4.3 Statistical significance of the experimental results 0 0.2 0.4 0.6 0.8 1.0 yF / ( F ( d/Q02 )1/3) 0 5 10 15 20 25 30 35 f F η = 0.45 η = 0.89 η = 1.33 η = 1.76 η = 2.20 η = 2.64 η = 3.08 (a) 0 0.5 1.0 1.5 2.0 2.5 η 0 0.2 0.4 0.6 0.8 1.0 P ∗ F t˜ = 1 t˜ = 2 t˜ = 3 t˜ = 4 t˜ = 5 t˜ = 6 t˜ = 7 t˜ = 8 (b) Figure 4.14: (a) Probability density function of the experimental dye concentration yF,exp/ ( F ( d/Q02 )1/3), non-dimensionalized and in similarity form, at different values of the similarity variable η in the case of constant-flux releases. We describe the exper- iments in § 4.1.1 and plot the ensemble-averaged concentration yF,exp in figure 4.7 with pluses. (b) Distribution against the similarity variable η at different non-dimensional times (plotted with different colours) of the probability that the concentration of trac- ers φF,exp is greater than a critical value φ∗ in the case of a constant-flux release in a quasi-two-dimensional jet. in similarity form) are linearly distributed from 0 to 1.07 (with Ny = 200, the number of bins). In figure 4.14(a), we show the p.d.f. fF , computed in (4.12), at seven different values of the similarity variable η for 0.45 ≤ η ≤ 3.08 (plotted with different colours).The distribution of the ensemble-averaged concentration yF,exp is plotted in figure 4.7 with pluses. As we can see in figure 4.14(a), the p.d.f. fF decreases with increasing concentration yF,exp. In figure 4.7, the maximum of yF,exp is found at η = 1.17, which corresponds, in figure 4.14(a), to the rightmost and lowest profile of fF (see light green curve at η = 1.33). The amplitude of fF is large for either large values of η or small values: η > 2.20 and η < 0.45. In fact, we find that the standard deviation of the p.d.f. of yF,exp grows approximately linearly with its average value, although with a hysteresis between the values before the concentration peak (i.e. η < 1.17) and the values after the concentration peak (i.e. η > 1.17). In the case of a constant-flux release of pollutants in a quasi-two-dimensional turbulent jet, the probability P ∗F to find concentrations of pollutants larger than a critical concentration level φ∗ (laterally-integrated concentration) at a given value 101 4 Experimental results for the streamwise dispersion and mixing η is for t > 0 P ∗F (η, t˜) = ∫ ∞ t˜−1/3φ˜∗ fF (y˜F,exp, η) dy˜F,exp, (4.14) where tildes denote non-dimensional values (see (4.1a,b)), and where we use equa- tion (3.25), φ(z, t) = t1/3y(η). It is interesting to note that the distribution of P ∗F increases in time, starting from 0 at t = 0. The increase in time of P ∗F is due to the constant flux of tracer concentration at the source of the jet and to the decrease (like z−1/2) of the velocity of the jet with distance. Thus, the laterally- integrated tracer concentration tends to increase at a fixed value of η ∝ z/t2/3 as time increases (i.e. for z increasing). As an example, we have plotted P ∗F in figure 4.14(b) against η at eight different non-dimensional times 1 ≤ t˜ ≤ 8 (plotted with different colours), for the critical non-dimensional concentration φ˜∗ = 1. We can clearly see that the probability P ∗F increases rapidly in time. As time increases, the maximum value of P ∗F appears to move to the left, towards η = 0, and is found in the range 0.9 ≤ η ≤ 1.3 (the peak of yF,exp, in figure 4.14a, is found at η = 1.17). We find that, in this example, the peak of the probability P ∗F first becomes greater than 0.05 (i.e. statistically significant) from t˜ ≥ 1.0 at the location η = 1.24, corresponding to z˜ = 1.2. The peak of the probability P ∗F becomes greater than 0.95 from t˜ ≥ 12.0 at the location η = 0.95, corresponding to z˜ = 5.0. 4.3.2 Instantaneous finite-volume release of virtual parti- cles Similarly to the constant-flux case presented above, we compute numerically the p.d.f. fv of the laterally-integrated numerical concentration yv,exp (in similarity form), for the Ne = 256 clusters, representing 2,624,256 virtual particles, released instantaneously in the velocity field of quasi-two-dimensional jets (see § 4.1.2). The p.d.f. fv is, for all 1 ≤ l ≤ Nη and 1 ≤ m ≤ Ny (with l and m two integers), fv(ym, ηl) = Ne,Nz ,Nt∑ i,j,k ∆ei,zj ,tk(ym, ηl) Ny∑ m Ne,Nz ,Nt∑ i,j,k ∆ei,zj ,tk(ym, ηl) δy , (4.15) 102 4.3 Statistical significance of the experimental results where δy = y2 − y1 is the concentration step and, for all 1 ≤ i ≤ Ne, 1 ≤ j ≤ Nz and 1 ≤ k ≤ Nt (with i, j and k three integers), ∆ei,zj ,tk(ym, ηl) =    1 if ym−1 ≤ yv,exp(ei, zj , tk) B ( d/Q02(ei) )1/3 < ym and ηl−1 ≤ η = zj tk2/3 ( Q02(ei)/d )1/3 < ηl, 0 otherwise , (4.16) where zj (the discretized streamwise coordinate) are linearly distributed from approximately 47 d to 128 d (depending on the experiment and with Nz = 1022), tk (the kth frame of the experiment) are linearly distributed from approximately 0 to tNt (with Nt of the order of 300, depending on the experiment), ηl (the discretized similarity variable) are linearly distributed from 0 to 5 (with Nη = 200, the number of bins) and ym (the discretized laterally-integrated concentration in similarity form) are linearly distributed from 0 to 5 (with Ny = 400, the number of bins). We present in figure 4.15(a) the p.d.f. fv, computed in (4.15), at eight different values of the similarity variable η for 1.52 ≤ η ≤ 3.50 (plotted with different colours). The distribution of the ensemble-averaged concentration yv,exp is plotted in figure 4.10 with pluses. As we can see in figure 4.15(a), the p.d.f. fv decreases even more rapidly than fF (shown in figure 4.14a) with increasing concentration yv,exp. In figure 4.10, the maximum of yv,exp is found at η = 1.83, which is close to the rightmost and lowest profile of fv in figure 4.15(a) (see curve at η = 2.07). Similarly to fF , the amplitude of the fv is the largest for either large values of η or small values: η > 2.4 and η < 1.8. Moreover, we also find that the standard deviation of the p.d.f. of yv,exp grows approximately linearly with its average value, although with more scatter than for the p.d.f. of yF,exp and with a stronger hysteresis between the values before the concentration peak (i.e. η < 1.83) and the values after the concentration peak (i.e. η > 1.83). In the case of an instantaneous finite-volume release of pollutants in a quasi-two- dimensional turbulent jet, the probability P ∗δ to find concentrations of pollutants larger than a critical concentration level φ∗ at a given value η is for t > 0 P ∗δ (η, t˜) = ∫ ∞ t˜2/3φ˜∗ fv(y˜v,exp, η) dy˜v,exp, (4.17) 103 4 Experimental results for the streamwise dispersion and mixing 0 0.1 0.2 0.3 0.4 0.5 yv,exp/ ( B/ ( d/Q02 )1/3) 0 10 20 30 40 50 f v η = 1.52 η = 1.80 η = 2.07 η = 2.37 η = 2.65 η = 2.92 η = 3.22 η = 3.50 (a) 0 0.5 1.0 1.5 2.0 2.5 η 0 0.2 0.4 0.6 0.8 1.0 P ∗ δ t˜ = 1 t˜ = 2 t˜ = 3 t˜ = 4 t˜ = 5 t˜ = 6 t˜ = 7 t˜ = 8 (b) Figure 4.15: (a) Probability density function of the concentration of virtual parti- cles yv,exp/ ( B ( d/Q02 )1/3), non-dimensionalized and in similarity form, at different values of the similarity variable η in the case of instantaneous finite-volume releases. We describe the experiments in § 4.1.2 and plot the ensemble-averaged concentration yv,exp in figure 4.10 with pluses. (b) Distribution against the similarity variable η at different non-dimensional times (plotted with different colours) of the probability that the concentration of tracers φv,exp is greater than a critical value φ∗ in the case of an instantaneous finite-volume release in a quasi-two-dimensional jet. where we use equation (3.66), φ(z, t) = t−2/3y(η). It is interesting to note that, contrary to P ∗F , the probability P ∗δ decreases in time. The probability P ∗F de- creases in time because finite volumes of tracers become more dilute, due to the streamwise dispersion, as they are transported by the jet. As an example, we have plotted P ∗δ in figure 4.15(b) against η at eight different non-dimensional times 1 ≤ t˜ ≤ 8 (plotted with different colours), for the critical non-dimensional concentration φ˜∗ = 1. We can clearly see that the probability P ∗δ decreases rapidly in time. The maximum value of P ∗δ is located at approximately η = 0.7, a secondary, much smaller, local maximum is found in the range 1.1 ≤ η ≤ 1.5. We find that, in this example, the peak of the probability P ∗δ first becomes less than 0.95 from t˜ ≥ 0.3 at the location η = 0.82, corresponding to z˜ = 0.4. The peak of the probability P ∗δ becomes less than 0.05 (i.e. statistically insignificant) from t˜ ≥ 9.1 at the location η = 0.77, corresponding to z˜ = 3.4. 4.3.3 Discussion Owing to the large number of experiments conducted in the cases of constant-flux releases of dye and instantaneous releases of finite volumes of virtual particles, 104 4.3 Statistical significance of the experimental results we have been able to compute the statistical significance of the measurements of the laterally-integrated concentration, in similarity form. We find that, in both cases, the p.d.f. of the concentration tends to decrease and spread rapidly with increasing concentration. This means that near the location of the concentration peak, the difference between the concentration predicted by the model and the experimental (or real) concentration is likely to be much larger than at the tail or front of the distribution where the concentrations are smaller. Therefore, the model is less accurate in the prediction of the values of the largest concentrations than in the prediction of the values of the lowest concentrations. From the experimental results of the p.d.f. in the constant-flux and instanta- neous finite-volume cases, we can determine the probability of having a certain range of tracer concentrations at a certain location in time and space. We discuss the problem, relevant to pollution control, of how to calculate the probability to find concentration levels greater than a critical value in a quasi-two-dimensional jet. We find that in the constant-flux case, the probability increases rapidly in time. On the other hand, the probability decreases for a finite-volume release. In both cases, the location (in terms of the similarity variable η) of the maximum of the probability seems to remain constant in time. It is important to note that, in this section, we present and discuss the results of the laterally-integrated concentration φ instead of the actual concentration C. We observed previously in figure 4.5 that the concentration C tends to disperse linearly with distance across the jet, for both the constant-flux and the finite- volume cases. Therefore, the concentration of tracers is diluted not only because of the streamwise dispersion but also because of the lateral spreading of the jet. As a result, the distribution of the probability P ∗δ to encounter some concentrations C greater than a critical value should decrease more rapidly in time in the case of finite-volume releases of tracers. In the case of constant-flux releases, we believe that instead of increasing in time (for φ) the distribution of the probability P ∗F should actually decrease in time for the concentration C. This is related to the fact that, in steady state, the concentration CF decreases like z−1/2 whereas the laterally-integrated concentration φF increases like z1/2 (as discussed in § 3.3.2). 105 4 Experimental results for the streamwise dispersion and mixing 4.4 Conclusion In Chapters 3 and 4, we have analysed the time-dependent transport and dis- persion properties along the streamwise direction of quasi-two-dimensional jets. We model the evolution in time and space of the concentration of passive tracers using a one-dimensional time-dependent effective advection–diffusion equation. We integrate the concentration across the jet in order to be able to solve the effective advection–diffusion equation (3.4). From the analysis of experimental results we find that this simplification appears to be appropriate, because the tracer distribution remains confined within the quasi-two-dimensional jet between two linearly-expanding straight-sided boundaries (see figure 2.4). Neglecting any molecular diffusion, we assume a streamwise turbulent eddy diffusive coefficient Dzz proportional to the product of the local half-width of the jet b(z) ∝ z and the local time-averaged maximum streamwise velocity wm(z) ∝ z−1/2 (essentially based on mixing length theory). The streamwise turbulent eddy diffusive diffusion coefficient models physically the interaction between the core and eddy structures of quasi-two-dimensional jets. (In § 2.5, we showed that the core–eddy structure was self-similar with height, with characteristic local length-scale b(z), and with characteristic local velocity scale wm(z).) Using Dzz ∝ z1/2 we are able to transform the effective advection–diffusion equation into a similarity form. We solve analytically the resulting ordinary dif- ferential equation in the cases of a constant-flux release and an instantaneous finite-volume release yielding a ‘fundamental solution’. The solutions depend on two parameters, an advection parameterKa and a dispersion parameterKd, which we determine using experimental measurements. We also provide an integral for- mulation for the general problem of an arbitrary time-dependent release of tracers governed by a source function. The integral formulation for this more realistic case is the convolution between the fundamental solution found for the instan- taneous finite-volume release and the source function. We present an analytical solution for the general problem in the case of a rectangular source function (i.e. the flux of tracers at the jet source is constant for a finite period of time, T0, and zero otherwise, thus releasing a finite volume). At large time (t ≫ T0), this solution converges towards the fundamental solution found for the instantaneous finite-volume release. On the other hand, for T0 → ∞, this solution converges towards the solution found for the constant-flux release. 106 4.4 Conclusion Furthermore, we show theoretically that, owing to dispersion mechanisms, a non-negligible portion of the total volume of tracers released travels ahead of the advective front, in both the finite-volume and the constant-flux cases. The advective front corresponds to the location of the volume of tracers (in the finite- volume case) or the front of the tracer distribution (in the constant-flux case) if all dispersion mechanisms are ignored and Kd = 0. We also find that the streamwise dispersion increases in time as t2/3. In this chapter, we compare the theoretical model developed in Chapter 3 with experimental measurements obtained by tracking the concentration of dye or vir- tual particles in time and space. We conduct both constant-flux and finite-volume releases of dye in quasi-two-dimensional steady turbulent jets. We also release fi- nite volumes of virtual particles (transported as passive tracers) instantaneously in the fully resolved time-dependent velocity fields of quasi-two-dimensional steady turbulent jets. We consider the experimental data for constant-flux releases of dye more accurate because the initial, boundary and integral conditions imposed in the theoretical model are more straightforward to satisfy experimentally. We find that the experimental results agree well with the theoretical prediction, using either the laterally-integrated concentration of dye φ or the streamwise concentra- tion flux of dyeMφ as defined in (3.6) and (3.50), respectively. The similarity scal- ing derived from the model η ∝ z/t2/3 is appropriate to study this phenomenon. We find that what we refer to as our ‘reference’ values for the advection and dis- persion parameters are Ka = 1.65 and Kd = 0.09, respectively, determined from the study of the concentration in the constant-flux dye experiments. We largely confirm these results by the experimental data obtained with finite- volume releases of virtual particles. The data converge in similarity form towards the fundamental theoretical solution assuming an instantaneous finite-volume re- lease. The similarity scaling η ∝ z/t2/3 is also appropriate in this case. We find that the best fits to the advection and dispersion parameters are Ka = 1.62 and Kd = 0.09, respectively. In the case of finite-volume releases of dye, we find that the experimental concentration distribution has not converged towards the asymptotic fundamental solution assuming an instantaneous release. We believe that this is principally due to the fact that the dye could not be released instan- taneously in the experiment. The duration of the dye release introduces a new time scale T0, which affects the concentration distribution. Until t ≫ T0, the 107 4 Experimental results for the streamwise dispersion and mixing Case Theory Ka Kd Ka/Kd ηa µ σ β ξ CF yF 1.65 0.09 18.3 1.83 0.65–0.62 0.29–0.30 0.12–0.10 0.16–0.13 CF yM 1.55 0.07 23.6 1.83 0.58–0.50 0.31–0.31 0.09–0.06 0.19–0.12 IFV yδ 1.62 0.09 18 1.83–1.83 0.99–1.03 0.17–0.19 0.49–0.54 0.13–0.17 FV φT0 1.75 0.09 19.4 1.83 a 1.03a 0.19a 0.54a 0.17a aTheoretical value after φT0 converges to φδ. Table 4.1: Summary of the key experimental results found in the constant-flux case (CF) for dye releases, in the instantaneous finite-volume case (IFV) for virtual-particle releases and the finite-volume case (FV) for dye releases. The values for the advection and dispersion parameters Ka and Kd are obtained from the best least-squares fit of the experimental data. On the other hand, ηa, µ, σ, β and ξ are computed theoretically using the ‘reference’ parameters Ka = 1.65 and Kd = 0.09, found in the constant-flux case; if two values are indicated: the first value is measured experimentally while the second value is computed theoretically using Ka = 1.65 and Kd = 0.09. concentration distribution is in a transition regime, which we model using the general model φT0 defined in (3.88), assuming a rectangular source function. We find that the best fits to the advection and dispersion parameters are Ka = 1.80 and Kd = 0.08, respectively. We also calculate that, in this case, the distribution should ‘converge’ (i.e. the normalized absolute deviation between φT0 and φδ, de- fined in (3.91), is smaller than 0.1) towards the fundamental solution φδ defined in (3.73) after a duration equal to approximately 14 times the time of release of the dye (i.e. t ≥ 14T0). In other words, the dye distribution should converge towards an asymptotic distribution at z ≈ 2m (i.e. at a distance larger than four times the maximum distance of our study area). Our model appears to be robust to variations in the initial boundary conditions of the experiments. In the experiments with finite-volume releases of virtual par- ticles, even though the particles are released instantaneously but far away from the source, the particle concentration distribution seems to converge rapidly in time towards a stable asymptotic distribution predicted by the model. In the experiments with finite-volume releases of dye, even though the dye is released near the source but not instantaneously, we can prove that the dye concentra- tion distribution will eventually converge in time towards a stable asymptotic distribution predicted by the model. Moreover, we can estimate the time before convergence and provide a model for the transition regime. Overall, the model largely appears to agree with the data, especially at the dispersive front of the distribution. In table 4.1 we collect all the various key 108 4.4 Conclusion experimentally determined quantities. By comparing the various models with all the experiments, we are able to give an estimated range for the advection and dispersion parameters. We find that the advection and dispersion parameters are Ka = 1.65 ± 0.10 and Kd = 0.09 ± 0.02 respectively, and the ratio between the two is within the range 18 ≤ Ka/Kd ≤ 23.6. For both the constant-flux case and the instantaneous finite-volume case, the location in similarity space of the advective front (as defined in (3.40)) is found at ηa = 1.83. Then, in the case of constant-flux releases of tracers, we find that the ratio between the centroid and the advective front is approximately µF = 0.635±0.015 with a standard deviation normalized with ηa σF = 0.295 ± 0.005. At each instant in time, approximately βF = 11 % ±1 % (as defined in (3.47)) of the total volume of tracers having already been released is transported ahead of the advective front, at an averaged normalized distance in similarity space ξF = 0.145± 0.015 (as defined in (3.49)). In the case of an instantaneous finite-volume release of tracers, the ratio between the centroid and the advective front is approximately µB = 1.01 ± 0.02 with a standard deviation normalized with ηa σB = 0.18± 0.01. At each instant in time, approximately βB = 51.5 % ±2.5 % (as defined in (3.81)) of the total volume of tracers released is transported ahead of the advective front ηa, at an averaged normalized distance in similarity space ξB = 0.15± 0.02 (as defined in (3.83)). The analysis of the statistical significance of the experimental measurements of the laterally-integrated concentration reveals that experimental or real concen- trations are more likely to differ from the concentrations predicted by the model at large concentration levels than at low concentration levels. We find that the distribution, against the similarity variable η, of the probability to encounter laterally-integrated concentrations greater than a critical value increases in time for the case of constant-flux releases of tracers. On the other hand, the distribu- tion of the analogous probability decreases in time for the case of finite-volume releases of tracers. However, if we study the actual (non-laterally-integrated) con- centration, we believe that the probability distribution in the constant-flux case should also decrease in time due to lateral dispersion across the jet with distance. In § 3.1, we discussed the importance of modelling correctly the transport and dispersion of tracers in quasi-two-dimensional jet flows. We believe that the model developed in Chapter 3 provides not only a strong insight into these mechanisms but also a quantitative basis to predict them. In this chapter, comparisons with 109 4 Experimental results for the streamwise dispersion and mixing experimental data obtained using different techniques support the predictions of the model. From this comparison, we can also measure accurately the strength of the advection and the strength of the dispersion in quasi-two-dimensional jets, using only an advection parameter Ka and a dispersion parameter Kd, respec- tively. Finally, we have discovered that the streamwise dispersion increases in time like t2/3. In other words, a significant amount of tracers released in quasi- two-dimensional jets is transported faster than the speed predicted by a simple advection model. Such predictions are crucial to many applications, particularly in the event of environmental pollutions in rivers and lakes. 110 Chapter 5 Two-point statistics for turbulent relative dispersion in quasi-two-dimensional jets 5.1 Introduction The dispersion and mixing mechanisms in the turbulent flow of quasi-two-di- mensional jets are closely related to the dynamics of the large-scale structures identified as core and eddies. The core and eddy structures display very different flow properties. The velocity field of the core is very high in the streamwise direction, and it appears to be subject to a sinuous instability. The velocity field of the eddies is inherently vortical, with a time-averaged mean component in the streamwise direction. At the interface between the eddies and the core, the streamwise velocity has a large lateral (or cross-jet) gradient. Moreover, the 111 5 Two-point statistics for turbulent relative dispersion flow is turbulent everywhere in a quasi-two-dimensional jet. We believe that such distinctive Eulerian characteristics (of the flow in the core, in the eddies and at the interface between the two) also have distinctive dispersive and mixing properties. From the interaction between these structures, in time and space, results the global, mean dispersion mechanism of quasi-two-dimensional jets, which we model in Chapter 3 along the streamwise direction. Conversely, in this chapter, we adopt a Lagrangian approach to investigate the dispersion and mixing properties of the core and eddy structures of quasi- two-dimensional jets. In figure 4.4 presented in the previous chapter, we showed the evolution in time of clusters of virtual particles (or passive tracers) released in different parts of a quasi-two-dimensional jet: in an eddy (see figure 4.4a), between the eddy and the core (see figure 4.4b), and in the core (see figure 4.4c). We qualitatively described how the clusters of particles disperse and mix, while being transported by the jet. The virtual particles seeded in the eddy travel significantly slower than the virtual particles seeded in the core. The virtual particles seeded in the eddy appear to experience more vigorous stirring than the virtual particles seeded in the core. We also noticed that the virtual particles seeded in the core disperse laterally as they are advected by the flow. On the other hand, the virtual-particle cluster seeded between the eddy and the core display intense streamwise stretching. The aim of this study is to quantify these observations about the dispersion and mixing of the virtual particles in figure 4.4. We use statistical analysis to understand the underlying physical mechanisms. We study the probability dis- tribution of two-point properties, such as the lateral (or x-) distance between two points, the streamwise (or z-) distance between two points, the distance be- tween two points, and the ratio of the lateral distance to the streamwise distance between two points. We apply these probabilistic tools to clusters of virtual parti- cles released in quasi-two-dimensional jets (such as those in figure 4.4), where the two-point measurements are made between pairs of virtual particles. We compute the probability distribution of the two-point properties at each instant in time to obtain meaningful quantitative insight about the temporal and spatial dispersive dynamics of the jet structures. The work of Richardson (1926) pioneers the use of two-point statistics to study diffusion in turbulent flows (see e.g. Sawford, 2001; Salazar & Collins, 2009, for 112 5.1 Introduction recent reviews). Observing considerable discrepancies (by more than ten orders of magnitude) in the measurements of the atmospheric diffusivity, he argued that two-point statistics are more appropriate to explain diffusion in the atmosphere than single-point statistics (used previously to measure the diffusivity in the sense described by Fick’s law). Two-point statistics (such as the time average of the distance between two points) enable the study of the dispersion in the flow at each spatial scale (defined, for example, by the eddy size), without being influenced by the larger scales. From the probability density function of the distance between particles, Richardson derived his famous 4/3 law of diffusion. Batchelor (1952) developed a rigorous mathematical framework for the idea of Richardson (1926) to use two-point statistics in order to study turbulent relative dispersion. He applied two-point statistics to the diffusion of passive scalars in homogeneous isotropic turbulence. The concept of two-point statistics has then been used to study turbulent dis- persion in the ocean and in the atmosphere (see e.g. Monin & Yaglom, 1975, pp. 556–567, for a review). Salazar & Collins (2009) and Yeung (2002) give a summary of experimental and numerical works investigating turbulent relative dispersion. In experimental turbulent flows, two-point statistics can be calculated by tracking Lagrangian particles. According to Toschi & Bodenschatz (2009), the most successful current technique to perform Lagrangian particle tracking is called particle tracking velocimetry. For example, Bourgoin et al. (2006) used particle tracking velocimetry to measure the mean square distance between particles in a turbulent flow (generated “between coaxial counter-rotating baffled disks in a closed chamber”). They confirmed the theoretical prediction of Batchelor (1950) that the temporal evolution of the distance between pairs of particles during the superdiffusion stage (i.e. the regime when the mean square distance between par- ticles increases in time like tα with α > 1, Bourgoin et al., 2006) is influenced by the initial distance separation of the particles. Bourgoin et al. (2006) also commented on the scarcity of direct experimental evidence for turbulent relative dispersion. Toschi & Bodenschatz (2009) attributed the lack of experimental ev- idence to the technical difficulties of the implementation of Lagrangian particle tracking in fully turbulent flow. We believe that applying two-point statistics to the turbulent flow of quasi- two-dimensional jets can give new insight about turbulent relative dispersion in 113 5 Two-point statistics for turbulent relative dispersion the case of a non-homogeneous and anisotropic turbulent flow. We use what we believe to be a new method to calculate these two-point statistics, which we call virtual particle tracking (see § 4.1). The virtual-particle-tracking technique (which we use to produce the results shown in figure 4.4, mentioned above) con- sists of seeding and tracking virtual passive tracers in velocity fields measured using particle image velocimetry. The results presented in this study focus pri- marily on the dispersion properties of quasi-two-dimensional jets, but not directly on the transport or turbulent mixing properties. By definition, two-point statis- tics do not depend on mean transport motion, and thus cannot investigate it. (The transport properties of the jet have actually been studied extensively in Chapters 2, 3 and 4.) On the other hand, we believe that mixing properties can- not be directly examined from the results we present in this thesis for technical reasons. The measurements of the velocity fields (performed using particle image velocimetry), though well-resolved in time (the time resolution is one order of magnitude smaller than the Kolmogorov time scale, τηK ≈ 40 ms), do not have the spatial accuracy necessary to investigate the finest scales of turbulence in our flow (the Kolmogorov length scale is of the order of ηK ≈ 0.2 mm, as discussed in § 2.2.2). In Chapter 4, we quantify the mixing through the dilution of the dye concentration. Likewise, in this chapter, we infer indirectly the turbulent mixing processes from the dispersion, stretching and folding of our particle distributions. In order to comprehend fully the temporal evolution of the probability distri- butions of two-point properties applied to virtual-particle clusters seeded in the different parts of the flow of quasi-two-dimensional jets, we compare our results with other flow fields. As a preliminary study, we apply our statistical tools to simple distributions of points (such as a circle, an ellipse and a square) evolving in diverging velocity fields. The purpose of this preliminary study is to understand how the probability distributions of two-point properties are related to a given initial distribution of particles, and how they evolve in time. Then, we compare the results for the time-dependent flow field of a quasi-two-dimensional jet with results obtained using the time-averaged flow field of the same jet. This compari- son allows us to identify some key dispersive mechanisms due to the core and eddy structures and emphasizes the importance of their time-dependent interactions. The rest of this chapter is organized as follows. In § 5.2, we describe mathe- matically how to compute the probability distribution of the two-point properties 114 5.2 Mathematical definitions of two-point probability distributions stated above, in both the continuous case and the discrete case. In § 5.3, we present a preliminary study of three analytical and numerical test cases: an ax- isymmetric expansion of a circular domain, a non-axisymmetric expansion of an elliptical domain, and a diffusive expansion of a square domain. In § 5.4, we present the results of the probability distributions for the three clusters of virtual particles seeded in the quasi-two-dimensional jet shown in figure 4.4. We compare these results with similar results obtained in the equivalent time-averaged velocity field of the jet. Finally, we draw our conclusions in § 5.5. 5.2 Mathematical definitions of two-point proba- bility distributions 5.2.1 Continuous formulation The probability density function (p.d.f.) fY of a real-valued random variable Y is the derivative of the cumulative distribution function (c.d.f.) F Y (see e.g. Pope, 1985) such that, for any real number y , fY (y ) = dF Y (y ) dy . (5.1) The c.d.f. can be defined as the probability that the random variable Y takes on a value less than or equal to y , F Y (y ) = P (Y ≤ y ). (5.2) In the present study, we wish to compute the p.d.f. of four characteristic properties between pairs of points (x1,x2) distributed in a domain A . The first characteristic property is the lateral distance between two points: H (x1,x2) = |x1 − x2|, with x1 = (x1, z1), x2 = (x2, z2) ∈ A . (5.3) The second characteristic property is the streamwise distance between two points: V (x1,x2) = |z1 − z2|, with x1 = (x1, z1), x2 = (x2, z2) ∈ A . (5.4) 115 5 Two-point statistics for turbulent relative dispersion The third characteristic property is the (Euclidean) distance between two points: D (x1,x2) = ‖x1 − x2‖, with x1 = (x1, z1), x2 = (x2, z2) ∈ A , (5.5) where ‖x1 − x2‖ = ( (x1 − x2)2 + (z1 − z2)2 )1/2 . The fourth characteristic prop- erty is the ratio of the lateral distance to the streamwise distance between two points: M (x1,x2) = |x1 − x2| |z1 − z2| , with x1 = (x1, z1), x2 = (x2, z2) ∈ A . (5.6) The probability that the random variable Y (x1,x2) (= H (x1,x2), V (x1,x2), D (x1,x2) or M (x1,x2)), with x1, x2 ∈ A , takes on a value less than or equal to y (= h , v , d or m , respectively) is PA (Y (x1,x2) ≤ y ) = 1∫ A ς(x1)dτ1 ∫ A PA (Y (x1,x2) ≤ y |x1)ς(x1) dτ1, (5.7) where dτ1 is an appropriate differential for the domain A with respect to the first point x1, and ς(x) is the density of the probability distribution (i.e. it is a measure of the local probability, which may not be uniform, at the point x ∈ A ). PA (Y (x1,x2) ≤ y |x1) is the conditional probability that the random variable Y (x1,x2) takes on a value less than or equal to y knowing x1 and is defined as PA (Y (x1,x2) ≤ y |x1) = 1∫ A ς(x2)dτ2 ∫ A ∩B(x1,y ) ς(x2) dτ2, (5.8) where the domain B(x1, y ) is defined such that x2 ∈ B(x1, y ) if Y (x1,x2) ≤ y with x1 known, and dτ2 is an appropriate differential for the domain A with respect to the second point x2. Finally, according to (5.1) and (5.2), the p.d.f. of the random variable Y in the domain A is fYA (y ) = dPA (Y (x1,x2) ≤ y ) dy , (5.9) where the probability PA (Y (x1,x2) ≤ y ) is defined by (5.7) and (5.8). 116 5.3 Test studies in diverging velocity fields 5.2.2 Discrete formulation The probability density functions of the four characteristic properties of the dis- tribution of particles defined in (5.3), (5.4), (5.5), and (5.6), for H , V , D and M , respectively, can also be formulated in the discrete case. For n particles xi = (xi, zi) of weight ωi (similarly to the concept of density ς used previously for the continuous case) distributed in a domain A , the discrete p.d.f. of a random variable Y (x1,x2) (= H (x1,x2), V (x1,x2), D (x1,x2) or M (x1,x2)) in this domain is fYA (yk) = 1 n∑ i>j ωiωj δy n∑ i>j Yi,j(yk), k ∈ N, 1 ≤ k ≤ N, (5.10) with δy = y1 − y0 and, for all 1 ≤ i ≤ n and 1 ≤ j ≤ n (where i and j are two integers), Yi,j(yk) = { ωiωj if yk−1 ≤ Y (xi,xj) < yk 0 otherwise . (5.11) Here, yk are distributed from y0 = 0 to yN , the maximum value taken by Y (x1,x2) in the domain A , while N is the number of bins. The distribution of yk is linear for Y = H , V and D , and logarithmic for Y = M . 5.3 Test studies in diverging velocity fields We calculate analytically or numerically the time evolution of the p.d.f.s of the two-point properties H , V , D , and M (as defined in (5.3), (5.4), (5.5) and (5.6), respectively) in the case of continuous or discrete distributions of points in simple two-dimensional domains. Firstly, we study a circular domain expanding axisymmetrically in a diverging velocity field. Secondly, we investigate the non- axisymmetric expansion of an elliptical domain, in order to understand the effect on the p.d.f.s of a variation in the aspect ratio of the domain. Thirdly, we study the effect of molecular diffusion on the p.d.f.s for an initially square distribution of discrete points. Following the example of Richardson (1926), who derived the p.d.f. of the dis- tance between particles distributed on a straight line, we believe this preliminary 117 5 Two-point statistics for turbulent relative dispersion analysis can help us understand the turbulent relative dispersion in quasi-two-di- mensional jets, which is studied in § 5.4. In these test studies, the actual velocity fields in which we set our domains (the disc, the ellipse and the square) are not important because we do not intend to relate them directly to the velocity fields of quasi-two-dimensional jets. The velocity fields of the test studies are merely a means to change the shape of our three domains in time. In fact, this is rather the effect of the time evolution of the shape of our domains on the p.d.f.s that we intend to compare with the time evolution of the p.d.f.s for the three particle clusters presented in § 5.4. 5.3.1 Circular domain in an axisymmetric diverging veloc- ity field We define in R2 a continuous uniform distribution Dt where the initial density ςDt(x, 0) ≡ 1 for all x = (x, z) such that x2 + z2 ≤ R02 (where R0 is the initial radius of the disc Dt) and ςDt(x, 0) ≡ 0 otherwise. The domain Dt evolves in time due to a constant diverging radial velocity field (u, w) defined by u(x) = x, w(z) = z. (5.12) Hence, the radius of the disc increases uniformly in time at an exponential rate: R(t) = etR0; (5.13) and the density decreases in time such that ςDt(x, t) =    R02 R2(t) = e −2t ∀ x = (x, z), x2 + z2 ≤ R2(t) 0 otherwise . (5.14) Since the domain Dt is axisymmetric and expands radially at all times, the p.d.f. of the lateral distance between two points is equal to the p.d.f. of the streamwise distance between two points in the same domain, i.e. fHDt = fVDt . Moreover, the p.d.f. is self-similar in time and depends only on the radius R(t). Using equations (5.7) and (5.8) with Y = H (or V ), defined in (5.3) (and (5.4)), y = h (or v ), 118 5.3 Test studies in diverging velocity fields the c.d.f. of H (x1,x2) (or V (x1,x2)) with x1, x2 in A = Dt is, for h ≥ 0, FHDt (h ) = 4 πR2(t) ∫ R(t) 0 PDt(H (x1,x2) ≤ h |x1) √ R2(t)− x12 dx1, (5.15) where we use the fact that the c.d.f. is symmetric with respect to both the x- axis and the z-axis, the conditional probability does not depend on z1 (as long as x12 + z12 ≤ R2(t)) and the density is uniform over the whole domain Dt. For 0 ≤ h ≤ R(t), the conditional probability is PDt(H (x1,x2) ≤ h |x1) = 1pi ( arcsin ( x1+h R(t) ) +( x1+h ) R(t) √ 1− ( x1+h R(t) )2 −arcsin ( x1−h R(t) ) − (x1−h )R(t) √ 1− ( x1−h R(t) )2 ) , 0≤x1≤R(t)−h , (5.16) and PDt(H (x1,x2) ≤ h |x1) = 1pi ( pi 2−arcsin ( x1−h R(t) ) − (x1−h )R(t) √ 1− ( x1−h R(t) )2 ) , R(t)−h≤x1≤R(t). (5.17) For R(t) ≤ h ≤ 2R(t), the conditional probability is PDt(H (x1,x2) ≤ h |x1) = 1, 0 ≤ x1 ≤ −R(t) + h , (5.18) and PDt(H (x1,x2) ≤ h |x1) = 1pi ( pi 2−arcsin ( x1−h R(t) ) − (x1−h )R(t) √ 1− ( x1−h R(t) )2 ) , −R(t)+h≤x1≤R(t). (5.19) For 2R(t) ≤ h PDt(H (x1,x2) ≤ h |x1) = 1, 0 ≤ x1 ≤ R(t). (5.20) Details about the calculation of the conditional probability PDt(H (x1,x2) ≤ h |x1) can be found in Appendix B.1. We calculate the c.d.f. FHDt (as defined in (5.15)) by computing the integral numerically. We then obtain the p.d.f. fHDt (or fVDt) by differentiating this calculated c.d.f. numerically, i.e. fHDt (h ) = dFHDt (h ) dh . (5.21) 119 5 Two-point statistics for turbulent relative dispersion We plot the non-dimensional p.d.f. fHDtR(t) in figure 5.1(a) against the non- dimensional variable h /R(t). We can see that fHDtR(t) decreases smoothly from fHDt (0) = 32/ (3π2R(t)) to fHDt (2R(t)) = 0 (see details in Appendices B.2 and B.3). The p.d.f. fHSt of H in a square domain St defined by the density ςSt(x, 0) ≡ 1 for all x = (x, z) such that −R0 ≤ x ≤ R0 and −R0 ≤ z ≤ R0 and 0 otherwise (with R(t) described by (5.13) and ςSt(x, t) = e−2t if −R(t) ≤ x ≤ R(t) and −R(t) ≤ z ≤ R(t) and 0 otherwise) is fHSt(h ) = 2R(t)− h 2R2(t) . (5.22) The full derivation of (5.22) can be found in Appendix B.4. We can see in figure 5.1(a) that the ‘disc’ fHDt is somewhat similar to the ‘square’ fHSt (plotted with a dashed line), which decreases linearly from fHSt(0) = 1/ (R(t)) to fHSt (2R(t)) = 0. Similarly, we can compute the c.d.f. of the distance between two points using equations (5.7) and (5.8) with Y = D (defined in (5.5)), y = d , with x1, x2 in A = Dt. We have in polar coordinates, for d ≥ 0, FDDt(d ) = 2 R2(t) ∫ R(t) 0 PDt(D (x1,x2) ≤ d |x1)r1dr1, (5.23) where r1 = √ x12 + z12, and where we use the fact that the conditional probability is independent of the angle θ1 (as long as r1 ≤ R(t)) and the density is uniform over the whole domain Dt. For 0 ≤ d ≤ R(t), the conditional probability is PDt(D (x1,x2) ≤ d |x1) = d 2 R2(t) , 0 ≤ r1 ≤ R(t)− d , (5.24) and PDt(D (x1,x2) ≤ d |x1) = d 2piR2(t) ( pi−arccos(xI−r1d )+ (xI−r1) d √ 1−(xI−r1d ) 2 ) + 1 pi ( arccos( xIR(t))− xI R(t) √ 1−( xIR(t)) 2 ) , R(t)−d ≤r1≤R(t), (5.25) where xI = R2(t) + r12 − d 2 2r1 , (5.26) 120 5.3 Test studies in diverging velocity fields h /R(t) f H D t R (t ) Domain: disc Domain: square 0 0.5 1.0 1.5 2.0 0 0.2 0.4 0.6 0.8 1.0 (a) d /R(t) f D D t R (t ) 0 0.5 1.0 1.5 2.0 0 0.2 0.4 0.6 0.8 1.0 (b) square m f M D t (t ) 10−5 1 105 10−10 10−5 1 (c) Figure 5.1: Probability density functions in the case of a uniformly distributed disc Dt in an axisymmetric diverging velocity field described in (5.12) for: (a) the lateral (or streamwise) distance between two points fHDt (solid curve), defined in (5.21), the p.d.f. of the lateral (or streamwise) distance between two points in a square fHSt (defined in (5.22)) is plotted with a dashed line for comparison; (b) the distance between two points fDDt , defined in (5.30); (c) the ratio of the lateral distance to the streamwise distance between two points fMDt , defined in (5.34). is the x-coordinate of the intersection between the perimeter of Dt and the circle defined by x = (x, z) such that (x− r1)2 + z2 = d 2. For R(t) ≤ d ≤ 2R(t), the conditional probability is PDt(D (x1,x2) ≤ d |x1) = 1, 0 ≤ r1 ≤ −R(t) + d , (5.27) 121 5 Two-point statistics for turbulent relative dispersion and PDt(D (x1,x2) ≤ d |x1) = d 2piR2(t) ( pi−arccos(xI−r1d )+ (xI−r1) d √ 1−(xI−r1d ) 2 ) + 1 pi ( arccos( xIR(t))− xI R(t) √ 1−( xIR(t)) 2 ) , −R(t)+d ≤r1≤R(t). (5.28) For 2R(t) ≤ d , PDt(D (x1,x2) ≤ d |x1) = 1, 0 ≤ r1 ≤ R(t). (5.29) Details about the calculation of the conditional probability PDt(D (x1,x2) ≤ d |x1) can be found in Appendix B.5. As before, we calculate the c.d.f. FDDt (as defined in (5.23)) by computing the integral numerically. We then obtain the p.d.f. fDDt by differentiating this calculated c.d.f. numerically, i.e. fDDt(d ) = dFDDt(d ) dd . (5.30) We plot the non-dimensional p.d.f. fDDtR(t) in figure 5.1(b) against the non- dimensional variable d /R(t). It is interesting to note that the p.d.f. fDDt starts from 0 at d = 0 (as can be proved from (5.23), (5.24) and (5.25) using a similar technique to that used in Appendix B.2), increases to a maximum value (which appears to occur for d /R(t) < 1) and then vanishes at d = 2R(t) (as can be proved from (5.23), (5.27) and (5.28) using a similar technique to that used in Appendix B.3). We can also compute the c.d.f. of the ratio of the lateral distance to the streamwise distance between two points using equations (5.7) and (5.8) with Y = M , defined in (5.6), y = m , and with x1, x2 in A = Dt. We have in polar coordinates, for m ≥ 0, FMDt (m , t) = 4 πR2(t) ∫ R(t) 0 ∫ pi 2 0 PDt(M (x1,x2) ≤ m |x1)r1 dr1dθ1, (5.31) where x1 = r1 cos θ1 and z1 = r1 sin θ1, and where we use the fact that the density is uniform over the whole domain Dt and that the conditional probability is symmetric with respect to both the x-axis and the z-axis. The conditional 122 5.3 Test studies in diverging velocity fields probability, which, in this case, depends on θ1, is PDt(M (x1,x2) ≤ m |x1) = F (π − υ)−F (υ) + F (−υ)−F (υ − π), (5.32) with υ = arctan (1/m ), and where F (θ2) = 1piR2(t)  R2(t) 2 θ2+ r12 4 sin(2(θ2−θ1))+ R2(t) 2 ( arccos( r1R(t) sin(θ2−θ1))− r1 R(t) sin(θ2−θ1) √ 1−( r1R(t) sin(θ2−θ1)) 2 )) . (5.33) Details about the calculation of the conditional probability PDt(M (x1,x2) ≤ m |x1) can be found in Appendix B.6. As before, we calculate the c.d.f. FMDt (as defined in (5.31)) by computing the integral numerically. We then obtain the p.d.f. fMDt by differentiating this calculated c.d.f. numerically, i.e. fMDt (m ) = dFMDt (m ) dm . (5.34) We plot the dimensionless p.d.f. fMDt in figure 5.1(c) against the dimensionless variable m using a logarithmic scale. We can see that the p.d.f. vanishes at r → 0 and r → ∞. Moreover, fMDt is symmetric with respect to m = 1, owing to the axisymmetry of the domain Dt at all time. In other words, fMDt (m ) = fMDt (1/m ). 5.3.2 Elliptical domain in a non-axisymmetric diverging ve- locity field Now, we study the case of a non-axisymmetric diverging velocity field. We define in R2 a continuous uniform distribution Lt, which is identical to the disc Dt (described previously) at t = 0. The initial density of Lt is ςLt(x, 0) ≡ 1 for all x = (x, z) such that x2 + z2 ≤ R02 (where R0 is the initial radius of Lt) and ςDt(x, 0) ≡ 0 otherwise. For t > 0, the domain Lt evolves in time due to a constant diverging non-axisymmetric velocity field (u, w) defined by u(x) = x, w(z) = zc , (5.35) 123 5 Two-point statistics for turbulent relative dispersion where c > 1 is a constant. The case c = 1 corresponds to the axisymmetric expansion of the disc studied in the previous section. Also, we do not need to study the case c < 1 owing to the symmetry between the x and z spatial coordinates (or u and w components of the velocity field). The domain Lt has an elliptical contour for t > 0, whose semi-major axis and semi-minor axis are denoted a and b in the x- and z-directions, respectively. The semi-axes a and b increase in time at different exponential rates: a(t) = etR0, b(t) = et/cR0, (5.36a,b) and the density decreases in time such that ςLt(x, t) =    R02 a(t)b(t) = e −t(c+1)/c ∀ x = (x, z), ( x a(t) )2 + ( z b(t) )2 ≤ 1 0 otherwise . (5.37) The domain Lt expands radially at all time but not axisymmetrically. The domain Lt remains symmetric with respect to both the x-axis and the z-axis. The p.d.f. of the lateral distance between two points fHLt is not equal to the p.d.f. of the streamwise distance between two points fVLt , but can be computed in a similar manner by substituting a(t) and b(t). Using equations (5.7) and (5.8) with Y = H , defined in (5.3), y = h , the c.d.f. of H (x1,x2) with x1, x2 in A = Lt is, for h ≥ 0, FHLt (h ) = 4 πa(t)b(t) ∫ a(t) 0 ∫ b(t) √ 1−(x1/a(t))2 0 PLt(H (x1,x2) ≤ h |x1) dz1dx1, (5.38) where we use the fact that the c.d.f. is symmetric with respect to both the x-axis and the z-axis, and the density is uniform over the whole domain Lt. Since the conditional probability does not depend on z1 (as long as (x1/a(t))2+(z1/b(t))2 ≤ 1), we can integrate (5.38) with respect to z1, to obtain FHLt (h ) = 4 πa2(t) ∫ a(t) 0 PLt(H (x1,x2) ≤ h |x1) √ a2(t)− x12 dx1. (5.39) We can notice that (5.39) is exactly the same as the c.d.f. of the lateral distance between two points in the domain Dt FHDt (see equation (5.15)), but with R(t) = 124 5.3 Test studies in diverging velocity fields (a) (b) (c) Figure 5.2: Distribution of a cluster of virtual particles seeded in the non-axisymmetric diverging velocity field described in (5.35) with c = 5 at successive non-dimensional times: (a) t = 0; (b) t = 1; (c) t = 2. a(t). Thus, the conditional probability in (5.39) is given by equations (5.16), (5.17), (5.18), (5.19) and (5.20) with R(t) = a(t). Therefore, the p.d.f. fHLt is equivalent to fHDt , but with R(t) = a(t). Similarly, we find by symmetry that the p.d.f. of the streamwise distance between two points in the domain Lt (i.e. fVLt) is equivalent to fVDt , but with R(t) = b(t). We have plotted the non-dimensional p.d.f. fHLta(t) and fVLtb(t) in figures 5.3(a) and 5.3(b), respectively, with solid curves. We can see that fHLta(t) and fVLtb(t) are similar when plotted against h /a(t) and v /b(t), respectively. In comparison, we plot the non-dimensional p.d.f. of the lateral (streamwise) distance between two points in a rectangular domain (defined such that −a(t) ≤ x ≤ a(t) and −b(t) ≤ z ≤ b(t)) with a dashed line in figure 5.3(a) (5.3b, respectively). The p.d.f. of the lateral (streamwise) distance between two points in a rectangular domain is equivalent to the p.d.f. in a square domain, described in (5.22), with R(t) = a(t) (R(t) = b(t), respectively). The calculation of the p.d.f. of D (the distance between two points) and M (the ratio of the lateral to the streamwise distances between two points) in the elliptical domain Lt is apparently more difficult. Instead of computing fDLt and fMLt analytically using the continuous formulation, we use the discrete formulation described in § 5.2.2. We distribute 7845 virtual passive tracers (or particles) of similar weight ω = 1 uniformly in a disc of initial radius R0 = 50 centred at the origin of a two-dimensional (x, z) infinite domain. The particles are seeded in the non-axisymmetric diverging velocity field described in (5.35), with c = 5. Using a discrete time step δt = 1, we find that the position of a given particle at t is xt = 2tx0, zt = (c+ 1 c )t z0. (5.40) We display in figures 5.2(a–c) the distribution of the particles at dimensionless times t = 0, t = 1 and t = 2, respectively. 125 5 Two-point statistics for turbulent relative dispersion From the location of all the particles at each instant in time, given by (5.40), we can compute the p.d.f. of the distance between two particles fDLt using the discrete formulation described in (5.10) and (5.11), with Y (x1,x2) = D (x1,x2), yk = dk and N = 100. We plot the non-dimensional p.d.f. fDLta(t) in figure 5.3(c) against d /a(t) for t = 0 (black), t = 1 (blue) and t = 2 (red). We can clearly see that the p.d.f. is no longer self-similar in time. The peak of the curve increases and move towards d = 0 (i.e. to the left) as time increases. The spurious fluctuations that can be seen in the p.d.f. fDLt are due to discretization issues. These fluctuations, particularly prominent at t = 0, are due to the fact that the discrete particle distribution has not enough randomness. Thus, despite a large number of pairs of particles (30,768,090) there cannot be a statistically good partition of all their separation distances among the N = 100 bins of the discretized variable dk. Again, from the location of all the particles at each instant in time, given by (5.40), we can compute the p.d.f. of the ratio of the lateral distance to the stream- wise distance between two particles fMLt using the discrete formulation described in (5.10) and (5.11), with Y (x1,x2) = M (x1,x2), yk = m k and N = 100. We plot the dimensionless p.d.f. fMLt in figure 5.3(d) against m for t = 0 (black), t = 1 (blue) and t = 2 (red). We can clearly see that for t > 0 the p.d.f. is not symmetric with respect to m = 1, as it is at t = 0 when the domain is circular. The p.d.f. fMLt displaces to the right as the aspect ratio a(t)/b(t) of the domain increases in time. Therefore, the evolution in time of fMLt can reveal a change in the aspect ratio of the domain studied. The spurious fluctuations that can be seen in the p.d.f. fMLt are also due to the discretization issue mentioned previously. 5.3.3 Square cluster of virtual particles in a diffusive veloc- ity field Now, we study the case of a diffusive velocity field. We distribute 3721 virtual passive tracers (or particles) of similar weight ω = 1 uniformly in a square of unit size centred at the origin of a two-dimensional (x, z) infinite domain. At each time step, the particles move following a two-dimensional random walk of length 500(t+1). We designate by Kt the diffusing distribution of particles. We display in figures 5.4(a–c) the distribution of the particles at dimensionless times t = 0, t = 1 and t = 2, respectively. From the location of all the particles at each instant in time, we can compute 126 5.3 Test studies in diverging velocity fields h /a(t) f H L t a( t) Domain: ellipse Domain: rectangle 0 0.5 1.0 1.5 2.0 0 0.2 0.4 0.6 0.8 1.0 (a) v /b(t) f V L t b( t) Domain: ellipse Domain: rectangle 0 0.5 1.0 1.5 2.0 0 0.2 0.4 0.6 0.8 1.0 (b) 0 0.5 1.0 1.5 2.0 d /a(t) 0 0.2 0.4 0.6 0.8 1.0 1.2 f D L t a( t) t = 0 t = 1 t = 2 (c) 1.0 0.01 0.1 1 10 100 m 0.01 0.1 1 f M L t t = 0 t = 1 t = 2 (d) Figure 5.3: Evolution in time of the probability density functions in the case of a uniformly distributed elliptical domain Lt in a non-axisymmetric diverging velocity field described in (5.35) for: (a) the lateral distance between two points fHLt (solid curve) computed in the continuous case using (5.15) with R(t) = a(t), the p.d.f. of the lateral distance between two points in a rectangle (defined in (5.22) with R(t) = a(t)) is plotted with a dashed line for comparison; (b) the streamwise distance between two points fVLt (solid curve) computed in the continuous case using (5.15) with R(t) = b(t), the p.d.f. of the streamwise distance between two points in a rectangle (defined in (5.22) with R(t) = b(t)) is plotted with a dashed line for comparison; (c) the distance between pairs of particles fDLt computed in the discrete case using (5.10) and (5.11) with n = 7845 and N = 100 at times t = 0 (black), t = 1 (blue) and t = 2 (red); and (d) the ratio of the lateral distance to the streamwise distance between pairs of particles fMLt computed in the discrete case using (5.10) and (5.11) with n = 7845 and N = 100 at times t = 0 (black), t = 1 (blue) and t = 2 (red), note that, in this case, the distribution is in a log–log plot. the p.d.f. of the lateral distance between two particles fHKt using the discrete formulation described in (5.10) and (5.11), with Y (x1,x2) = H (x1,x2), yk = hk and N = 100. We plot the p.d.f. fHKt in figure 5.5(a) against h for t = 0 127 5 Two-point statistics for turbulent relative dispersion (a) (b) (c) Figure 5.4: Time evolution of an initially square distribution of particles following random walks, at successive non-dimensional times: (a) t = 0; (b) t = 1; (c) t = 2. (black), t = 1 (blue) and t = 2 (red). We can see that starting from the expected distribution for a square domain, the p.d.f. rapidly drops and becomes smoother (similarly to the p.d.f. for the circular domain fHDt displayed in figure 5.1a). Similarly, we can compute the p.d.f. of the streamwise distance between two particles fVKt using the discrete formulation described in (5.10) and (5.11), with Y (x1,x2) = V (x1,x2), yk = vk and N = 100. We plot the p.d.f. fVKt in figure 5.5(b) against v for t = 0 (black), t = 1 (blue) and t = 2 (red). The p.d.f. of the streamwise distance is very similar to the p.d.f. of the lateral distance shown in figure 5.5(a), owing to the axisymmetry of the diffusion process. Similarly, we can compute the p.d.f. of the distance between two particles fDKt using the discrete formulation described in (5.10) and (5.11), with Y (x1,x2) = D (x1,x2), yk = dk and N = 100. We plot the p.d.f. fDKt in figure 5.5(c) against d for t = 0 (black), t = 1 (blue) and t = 2 (red). The p.d.f. gradually drops and spreads in time. For t > 0, the p.d.f. is similar to the p.d.f. for the circular domain fDDt displayed in figure 5.1(c). Similarly, we can compute the p.d.f. of the ratio between the lateral distance to the streamwise distance between two particles fMKt using the discrete formulation described in (5.10) and (5.11), with Y (x1,x2) = M (x1,x2), yk = m k and N = 100. We plot the p.d.f. fMKt in figure 5.5(d) against m for t = 0 (black), t = 1 (blue) and t = 2 (red). It is clear that the distribution of particles remain symmetric as the p.d.f. is centred around m = 1 at all time. Again, the spurious fluctuations that can be observed in the p.d.f.s fHKt , fVKt , fDKt , and fMKt are due to the discretization issue mentioned previously. 128 5.3 Test studies in diverging velocity fields 0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 h 0 0.5 1.0 1.5 2.0 f H K t t = 0 t = 1 t = 2 (a) 2 0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 v 0 0.5 1.0 1.5 2.0 f V K t t = 0 t = 1 t = 2 (b) 0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 d 0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 f D K t t = 0 t = 1 t = 2 (c) 2 10−2 0.1 1 10 102 m 10−2 0.1 1 f M K t t = 0 t = 1 t = 2 (d) Figure 5.5: Evolution in time of the probability density functions in the case of a diffusing domain Kt of virtual particles for: (a) the lateral distance between pairs of particles fHKt computed in the discrete case using (5.10) and (5.11) with n = 3721 and N = 100 at times t = 0 (black), t = 1 (blue) and t = 2 (red); (b) the streamwise distance between pairs of particles fVKt computed in the discrete case using (5.10) and (5.11) with n = 3721 and N = 100 at times t = 0 (black), t = 1 (blue) and t = 2 (red); (c) the distance between pairs of particles fDKt computed in the discrete case using (5.10) and (5.11) with n = 3721 and N = 100 at times t = 0 (black), t = 1 (blue) and t = 2 (red); and (d) the ratio of the lateral distance to the streamwise distance between pairs of particles fMKt computed in the discrete case using (5.10) and (5.11) with n = 3721 and N = 100 at times t = 0 (black), t = 1 (blue) and t = 2 (red), note that, in this case, the distribution is in a log–log plot. 5.3.4 Conclusion of the test studies We have analysed the probability distributions of two-point properties for: a circu- lar domain in axisymmetric expansion, an elliptical domain in non-axisymmetric expansion and a square domain expanding due to a diffusion-like process. From these studies, we have learnt that: 129 5 Two-point statistics for turbulent relative dispersion • the p.d.f. of the lateral and streamwise distance between two points, H and V respectively, characterizes the extent of the domain along each specific direction. The p.d.f. of H or V do not seem to depend on the actual shape of the distribution. In normalized form, the results of fH or fV were iden- tical between the disc and the ellipse (see solid curves in figures 5.1a,b and figures 5.3a,b), and were very similar to the results for the square domain and the diffusing domain (see black lines in figures 5.5a,b, and blue and red curves in figures 5.5a,b). • Regardless of the orientation, the p.d.f. of D characterizes the average distance between particles, and thus the shape of the domain. The p.d.f. fD of an axisymmetric domain (i.e. a disc in two dimensions) has its maximum value the furthest away from d = 0 (see figure 5.1c). The more elongated the distribution (e.g. an ellipse with large aspect ratio), the closer the peak of fD is to 0 (see figure 5.3c). • The ratio between the lateral and the streamwise distance between two points, M characterizes the symmetry between the x-direction and the z- direction. fM characterizes the aspect ratio of the extension of the domain along these two directions. Furthermore, through the evolution in time of these probability distributions we can note that • stretching or shrinking of the domain along the specific x- and z-directions can be quantified with fH and fV . This is also characterized by the shifting of the maximum of fM away from m = 1 (see figure 5.3d). • If fD remains self-similar in time, then the transformation seems to preserve the form (see the evolution of fD for the circular domain in figure 5.1c compared with the elliptical domain in figure 5.3c). • In a diffusion process (see figure 5.4), the distribution of particles tends to become more axisymmetric. In figures 5.5(a,b) the p.d.f. fHKt and fVKt are linear at t = 0 and then become smoother at each time step, similarly to fHLt and fVLt . 130 5.4 Analysis of the virtual particles in the jet structures 5.4 Analysis of the virtual particles in the jet structures 5.4.1 Virtual particles: time-dependent versus time-ave- raged velocity fields In the light of the preliminary study presented above, we analyse the statistical properties of particles seeded in the core and eddy structures of quasi-two-dimen- sional jets. We use the time-dependent velocity field of the jet shown previously in figure 4.3(b), where we seed a square cluster of n = 3721 virtual particles in an eddy (shown in light grey), a rectangular cluster of n = 7381 virtual particles and aspect ratio 2 at the interface between the core and the eddy (shown in grey), and a square cluster of n = 3721 virtual particles in the middle of the core (shown in dark grey). We have reproduced the time evolution of the particle clusters seeded in this time-dependent velocity field (previously shown in figures 4.4a–c) in figures 5.6(a–c). We now repeat the same process in the time-averaged velocity field (with an average duration time of 21.8 s, as explained in § 2.2.2) of the jet used in fig- ures 5.6(a–c). We show in figures 5.6(d–f ) the time evolution (the colour-scale used is the same to that used in figures 5.6a–c) of three clusters of virtual particles seeded in this time-averaged velocity field. The clusters in figures 5.6(d–f ) have the same size and are initially located at the same position in the velocity field as the clusters in figures 5.6(a–c), respectively. The comparison between the evolution of the particle clusters in the time- dependent and the time-averaged velocity fields reveals crucial information about the dynamics of the core and eddy structures. The cluster seeded at the location of an eddy in the time-averaged velocity field (see figure 5.6d) stretches in the streamwise direction and rotates counter-clockwise. The evolution of the corre- sponding cluster in the time-dependent velocity field (see figure 5.6a) does not display the same streamwise stretching, but rather expands isotropically. Strong stirring and turbulent mixing at the location of the eddy also seem to be fea- tures of the time-dependent velocity field only. At the interface between the eddy and the core, the time-averaged velocity field (see figure 5.6e) merely stretches the particle cluster in the streamwise direction. On the other hand, the cluster 131 5 Two-point statistics for turbulent relative dispersion seeded in the time-dependent velocity field (see figure 5.6b) not only experiences streamwise dispersion (clear at early times, shown with dark and red colours), but also divides into two as some particles are drawn into the neighbouring eddy. The particles drawn into the eddy experience the same isotropic dispersion and strong turbulent stirring as observed for the cluster in figure 5.6(a), whereas the particles that remain in the core are rapidly transported away. The particle clus- ter seeded in the core of the time-dependent velocity field (see figure 5.6c) has a similar evolution as the corresponding particle cluster seeded in the time-averaged velocity field (see figure 5.6f ). We can observe slightly more streamwise disper- sion in the time-averaged velocity field, whereas the time-dependent velocity field rather seems to stretch the particles in the cross-jet direction. 5.4.2 Two-point statistics: time-dependent versus time-a- veraged velocity fields We now study the time evolution of the two-point statistics of the three particle clusters evolving in the time-dependent velocity field (presented in figures 5.6a– c), as well as the three particle clusters evolving in the time-averaged velocity field of the same jet (presented in figures 5.6d–f ). For every cluster of both velocity fields, we compute the p.d.f., using the discrete formulation (5.10) and (5.11) (with N = 100, the number of bins), for the lateral distance H (except at t˜ = t/(d2/Q0) = 0 where we use the theoretical prediction defined in (5.22) for a rectangular domain), the streamwise distance V (except at t˜ = 0 where we use the theoretical prediction defined in (5.22) for a rectangular domain), the (Euclidean) distance D , and the ratio of the lateral distance to the streamwise distance between pairs of particles M , as defined in (5.3), (5.4), (5.5) and (5.6), respectively. We present the distributions of the p.d.f.s at three or four different times, linearly distributed from t˜ = 0 (the time we initially seed the particle cluster in the velocity field) to the time instant a particle of the cluster reaches the top boundary of the velocity field (this time varies between each cluster). In the eddy In figures 5.7(a–d), we present the non-dimensional p.d.f. of the lateral dis- tance fHEt d (where d = 0.5 cm is the nozzle width of the experimental appara- 132 5.4 Analysis of the virtual particles in the jet structures (a) 0 t (s)1 2 3 (b) (c) (d) (e) (f ) Figure 5.6: Comparison of the evolution in time of the virtual particles seeded in the velocity field shown in figure 4.3(b): (a–c) correspond to the time-dependent velocity field, while (d–f ) correspond to the time-averaged velocity field. (a,d) Particle cluster initially distributed at the centre of an eddy and shown in light grey in figure 4.3(b). (b,e) Particle cluster initially distributed between the eddy and the core and shown in grey in figure 4.3(b). (c,f ) Particle cluster initially distributed in the core of the jet and shown in dark grey in figure 4.3(b). Each colour corresponds to a time period of ∆t = 0.2 s (or ∆t˜ = 33 in dimensionless time), the colour scale shown at the bottom of (b) is the same to that used in figure 4.3(b). 133 5 Two-point statistics for turbulent relative dispersion tus presented in figure 4.1), the streamwise distance fVEt d, the ratio between the lateral and streamwise distances between pairs of virtual particles fMEt and the distance fDEt d for the cluster initially seeded at the location of an eddy in the time-dependent velocity field (see figure 5.6a). We plot the p.d.f. against the non-dimensional variables h /d, v /d, m and d /d in figures 5.7(a–d), respectively. The p.d.f.s are plotted at four different times, from t˜ = 0 to 392, using different colours. The colour scale used here is the same as shown in figure 5.6(b). Sim- ilarly, we show in figures 5.7(e–h) the evolution in time of the non-dimensional p.d.f.s fHEt d, fVEt d, fDEt d and fMEt for the cluster initially seeded at the location of an eddy in the time-averaged velocity field (see figure 5.6d). As we can see in figure 5.7(a), the range of fHEt d increases slightly in time from approximately 0 ≤ h /d ≤ 5 at t˜ = 0 to 0 ≤ h /d ≤ 7 at t˜ = 392. In the streamwise direction, the distribution of particles stretches more than in the lateral direction, as the range of fVEt d (shown in figure 5.7b) increases from approximately 0 ≤ v /d ≤ 5 at t˜ = 0 to 0 ≤ v /d ≤ 13 at t˜ = 392. The small change in aspect ratio can also be observed in figure 5.7(c), where fMEt is no longer exactly symmetric with m = 1 for t˜ > 0. Moreover, the smooth profile of fMEt at t˜ = 0 seems to be disturbed near m = 1 for t˜ > 0. This disturbance could suggest changes in the distribution of the particles with time. The evolution in time of the p.d.f. of the distance between two points fDEt , shown in figure 5.7(d), also reveals important changes in the distribution of the particles. The increase in the range of fDEt (from approximately 0 ≤ d /d ≤ 6 at t˜ = 0 to 0 ≤ d /d ≤ 13 at t˜ = 392) means that the particles have been spread over a larger domain. Moreover, the characteristic profile of fDEt at t˜ = 0 (which corresponds to a square domain) quickly vanishes, thus suggesting a radical change in the shape of the domain. Finally, for t˜ > 0, fDEt displays large fluctuations and peaks (different from the small fluctuations at t˜ = 0 which are due to the resolution problem mentioned previously), which vary in amplitude and location with time. In the light of the observations made in § 5.4.1, we can notice some major differences between the p.d.f.s for the time-dependent flow field (shown in figures 5.7a–d) and the p.d.f.s for the time-averaged flow field (shown in figures 5.7e– h). Firstly, the aspect ratio of the distribution in the time-averaged flow field deviates considerably in time from the aspect ratio of the distribution in the time-dependent flow field. In figure 5.7(e), the range of fHEt d decreases slightly in 134 5.4 Analysis of the virtual particles in the jet structures time from approximately 0 ≤ h /d ≤ 5 at t˜ = 0 to 0 ≤ h /d ≤ 3 at t˜ = 196. In figure 5.7(f ), the range of fVEt d increases considerably in time from approximately 0 ≤ h /d ≤ 5 at t˜ = 0 to 0 ≤ v /d ≤ 35 at t˜ = 196. The aspect ratio (between the lateral and streamwise extent) drops from 1 to less than 0.1, as clearly shown by fMEt plotted in figure 5.7(g). Secondly, the distribution of the p.d.f. of the distance fDEt for the time-averaged velocity field (plotted in figure 5.7h) is much smoother than for the time-dependent velocity field (plotted in figure 5.7d). We believe these continuous and rapid variations in time of the profile of fDEt for the time-dependent flow field can be related to the intense stirring effect of the turbulent eddy. The chaotic dynamics of the turbulent flow in the eddy perturbs the distribution of particles. This manifests itself in the rapid displacement of the peaks in the distribution of fDEt , for the time-dependent flow field. At the interface between the core and the eddy In figures 5.8(a–d), we present the non-dimensional p.d.f. of the lateral distance fHItd, the streamwise distance fVItd, the ratio between the lateral and streamwise distances between pairs of virtual particles fMIt and the distance fDItd for the clus- ter initially seeded at the interface between the eddy and the core (see figure 5.6b). We plot the p.d.f. against the non-dimensional variables h /d, v /d, m and d /d in figures 5.8(a–d), respectively. The p.d.f.s are plotted at four different times, from t˜ = 0 to 98, using different colours. The colour scale used here is the same as shown in figure 5.6(b). Similarly, we show in figures 5.8(e–h) the evolution in time of the non-dimensional p.d.f.s fHItd, fVItd, fDItd and fMIt for the cluster ini- tially seeded at the interface between the eddy and the core in the time-averaged velocity field (see figure 5.6e). In figure 5.8(a), the range of fHItd first decreases in time from approximately 0 ≤ h /d ≤ 10 at t˜ = 0 to 0 ≤ h /d ≤ 5 at t˜ = 33 before increasing to 0 ≤ h /d ≤ 13 at t˜ = 98. On the other hand, the distribution of particles steadily stretches in the streamwise direction, as the range of fVItd (shown in figure 5.8b) increases from approximately 0 ≤ v /d ≤ 5 at t˜ = 0 to 0 ≤ v /d ≤ 55 at t˜ = 98. This considerable change in the aspect ratio of the distribution, from 2 to less than 1/4 is also clearly revealed in figure 5.8(c), where the peak of fMIt rapidly moves from the right-hand side of m = 1 to the left-hand side. The evolution in time of the p.d.f. of the distance between two points fDIt , shown in figure 5.8(d), is different 135 5 Two-point statistics for turbulent relative dispersion Time-dependent flow field 0 2 4 6 8 10 h /d 0 0.1 0.2 0.3 0.4 0.5 0.6 f H E t d t˜ = 0 t˜ = 131 t˜ = 262 t˜ = 392 (a) Time-averaged mean flow field 0 1 2 3 4 5 h /d 0 0.2 0.4 0.6 0.8 1.0 1.2 f H E t d t˜ = 0 t˜ = 66 t˜ = 131 t˜ = 196 (e) Time-dependent flow field 0 2 4 6 8 10 12 14 v /d 0 0.1 0.2 0.3 0.4 f V E t d t˜ = 0 t˜ = 131 t˜ = 262 t˜ = 392 (b) Time-averaged mean flow field 0 5 10 15 20 25 30 35 v /d 0 0.1 0.2 0.3 0.4 f V E t d t˜ = 0 t˜ = 66 t˜ = 131 t˜ = 196 (f ) Time-dependent flow field 10−4 10−2 1 102 104 m 10−4 10−3 10−2 0.1 1 f M E t t˜ = 0 t˜ = 131 t˜ = 262 t˜ = 392 (c) Time-averaged mean flow field 10−4 10−2 1 102 104 m 10−4 10−3 10−2 0.1 1.0 f M E t t˜ = 0 t˜ = 66 t˜ = 131 t˜ = 196 (g) Time-dependent flow field 0 2 4 6 8 10 12 14 d /d 0 0.1 0.2 0.3 f D E t d t˜ = 0 t˜ = 131 t˜ = 262 t˜ = 392 (d) Time-averaged mean flow field 0 5 10 15 20 25 30 35 d /d 0 0.1 0.2 0.3 f D E t d t˜ = 0 t˜ = 66 t˜ = 131 t˜ = 196 (h) Figure 5.7: Comparison between the time evolutions (the colour-scale used is the same to that used in figure 5.6) of the p.d.f.s in the case of the cluster of virtual particles initially seeded in the eddy of the time-dependent (a–d) (see figure 5.6a) and the time-averaged (e–h) (see figure 5.6d) velocity fields, for: (a,e) the dimensionless lateral distance between pairs of particles fHEt d computed using (5.10) and (5.11) with n = 3721 and N = 100; (b,f ) the dimensionless streamwise distance between pairs of particles fVEt d computed using (5.10) and (5.11) with n = 3721 and N = 100; (c,g) the ratio of the lateral distance to the streamwise distance between pairs of particles fMEt computed using (5.10) and (5.11) with n = 3721 and N = 100 (in a log–log plot); and (d,h) the dimensionless distance between pairs of particles fDEt d computed using (5.10) and (5.11) with n = 3721 and N = 100. 136 5.4 Analysis of the virtual particles in the jet structures from the evolution of fDEt for the cluster seeded in the eddy (shown in figure 5.7d). The increase in the range of fDIt (from approximately 0 ≤ d /d ≤ 10 at t˜ = 0 to 0 ≤ d /d ≤ 55 at t˜ = 98) means that the particles have been spread over a larger domain very rapidly. Contrary to fDEt , the characteristic profile of fDIt at t˜ = 0 (which corresponds to a square domain) does not completely change. The main peak recedes towards d → 0 as time increases, thus suggesting a significant thinning of the distribution, probably owing to the intense lateral shear at the interface between the core and the eddy. Note that, in this case, the statistical study stops at t˜ ≈ 98, which corresponds to the time when the first particle reaches the top of the visualization window (see figure 5.6b). We can notice one major difference between the p.d.f.s for the time-dependent flow field (shown in figures 5.8a–d) and the p.d.f.s for the time-averaged flow field (shown in figures 5.8e–h). The distribution of the p.d.f.s fHIt , fVIt and fDIt are smoother for the time-averaged velocity field (plotted in figures 5.8e,f,h, respec- tively) than for the time-dependent flow field (plotted in figures 5.8a,b,c, respec- tively). We believe that the jaggedness observed for the time-dependent flow field is, similarly to the case of the eddy, related to the unstable and turbulent flow of the shear layer at the interface between the eddy and the core. In the core In figures 5.9(a–d), we present the non-dimensional p.d.f. of the lateral distance fHCt d, the streamwise distance fVCt d, the ratio between the lateral and streamwise distances between pairs of virtual particles fMCt and the distance fDCtd for the cluster initially seeded in the core (see figure 5.6c). We plot the p.d.f. against the non-dimensional variables h /d, v /d, m and d /d in figures 5.9(a–d), respectively. The p.d.f.s are plotted at four different times, from t˜ = 0 to 98, using different colours. The colour scale used here is the same as shown in figure 5.6(b). Similarly, we show in figures 5.9(e–h) the evolution in time of the non-dimensional p.d.f.s fHCt d, fVCt d, fDCtd and fMCt for the cluster initially seeded at the location of the core in the time-averaged velocity field (see figure 5.6f ). The p.d.f. of the lateral distance fHCt , presented in figure 5.9(a), seems to remain linear until approximately t˜ = 66, and then becomes bimodal at t˜ = 98 with a peak close to h = 0 and the other one near h = 13. The p.d.f. of the streamwise distance fVCt , shown in figure 5.9(b), decreases approximately linearly 137 5 Two-point statistics for turbulent relative dispersion Time-dependent flow field 0 2 4 6 8 10 12 14 h /d 0 0.1 0.2 0.3 0.4 f H I t d t˜ = 0 t˜ = 33 t˜ = 66 t˜ = 98 (a) Time-averaged mean flow field 0 2 4 6 8 10 h /d 0 0.05 0.10 0.15 0.20 f H I t d t˜ = 0 t˜ = 33 t˜ = 66 (e) Time-dependent flow field 0 10 20 30 40 50 v /d 0 0.1 0.2 0.3 0.4 f V I t d t˜ = 0 t˜ = 33 t˜ = 66 t˜ = 98 (b) Time-averaged mean flow field 0 5 10 15 20 25 30 v /d 0 0.1 0.2 0.3 0.4 f V I t d t˜ = 0 t˜ = 33 t˜ = 66 (f ) Time-dependent flow field 10−4 10−2 1 102 104 m 10−4 10−3 10−2 0.1 1 f M I t t˜ = 0 t˜ = 33 t˜ = 66 t˜ = 98 (c) Time-averaged mean flow field 10−4 10−2 1 10−2 10−4 m 10−4 10−3 10−2 0.1 1 f M I t t˜ = 0 t˜ = 33 t˜ = 66 (g) Time-dependent flow field 0 10 20 30 40 50 d /d 0 0.05 0.10 0.15 0.20 f D I t d t˜ = 0 t˜ = 33 t˜ = 66 t˜ = 98 (d) Time-averaged mean flow field 0 5 10 15 20 25 30 d /d 0 0.05 0.10 0.15 0.20 f D I t d t˜ = 0 t˜ = 33 t˜ = 66 (h) Figure 5.8: Comparison between the time evolutions (the colour-scale used is the same to that used in figure 5.6) of the p.d.f.s in the case of the cluster of virtual particles initially seeded between an eddy and the core of the time-dependent (a–d) (see figure 5.6b) and the time-averaged (e–h) (see figure 5.6e) velocity fields, for: (a,e) the dimensionless lateral distance between pairs of particles fHItd computed using (5.10) and (5.11) with n = 7381 and N = 100; (b,f ) the dimensionless streamwise distance between pairs of particles fVItd computed using (5.10) and (5.11) with n = 7381 and N = 100; (c,g) the ratio of the lateral distance to the streamwise distance between pairs of particles fMIt computed using (5.10) and (5.11) with n = 7381 and N = 100 (in a log–log plot); and (d,h) the dimensionless distance between pairs of particles fDItd computed using (5.10) and (5.11) with n = 7381 and N = 100. 138 5.5 Discussion and Conclusion at all time, with its range increasing only slightly (from 0 ≤ v /d ≤ 5 at t˜ = 0 to approximately 0 ≤ v /d ≤ 8 at t˜ = 98). The p.d.f. of the ratio between the lateral and streamwise distances fMCt , shown in figure 5.9(c), moves away to the right of the axis of symmetry m = 1. Therefore, the distribution stretches in the cross-jet or lateral direction, conversely to the cluster of particles seeded between the core and the eddy discussed above. This lateral stretching is probably due to the linear time-averaged lateral spreading of the jet velocity field with z. Finally, we can notice that, similarly to fHCt , the p.d.f. of the distance fDCt , shown in figure 5.9(d), also becomes more and more bimodal with time. The bimodality can be related to the gradual splitting of the cluster of virtual particles, as it becomes thinner along the centreline of the jet (see figure 5.6c). This effect must originate from the divergence of the lateral mean flow along the jet axis. We point out two minor differences between the p.d.f.s for the time-dependent flow field (shown in figures 5.9a–d) and the p.d.f.s for the time-averaged flow field (shown in figures 5.9e–h). Firstly, the range of the p.d.f. fHCt for the time-averaged velocity field (plotted in figure 5.9e) increases less (from 0 ≤ h /d ≤ 5 at t˜ = 0 to approximately 0 ≤ h /d ≤ 7 at t˜ = 66) than for the time-dependent flow field (plotted in figure 5.9a). The spurious fluctuations, which can be noticed in the distribution of fHCt in figure 5.9(e) at t˜ = 33 and 66 (and also, to some extent, in figures 5.9d,h at t˜ = 0), are due to the discretization issue mentioned previously (the problem does not occur at t˜ = 0 where we plot the theoretical prediction defined in (5.22)). Secondly, the bimodality of the distribution of the p.d.f. fDCt for the time-dependent flow field (shown in figure 5.9d) is not clear for the time- averaged flow field (shown in figure 5.9h), though it may develop at later time due to the time-averaged mean diverging lateral velocity near the centreline of the jet. 5.5 Discussion and Conclusion In this chapter, we investigate turbulent relative dispersion in the flow field of quasi-two-dimensional jets using two-point statistics. To obtain the data neces- sary to compute these two-point statistics, we have developed what we believe to be a new method which allows us to perform effectively Lagrangian particle tracking in the turbulent flow of the jets. We use virtual particle tracking, which 139 5 Two-point statistics for turbulent relative dispersion Time-dependent flow field 0 2 4 6 8 10 12 14 16 18 h /d 0 0.1 0.2 0.3 0.4 f H C t d t˜ = 0 t˜ = 33 t˜ = 66 t˜ = 98 (a) Time-averaged mean flow field 0 2 4 6 8 10 h /d 0 0.1 0.2 0.3 0.4 f H C t d t˜ = 0 t˜ = 33 t˜ = 66 (e) Time-dependent flow field 0 1 2 3 4 5 6 7 8 9 v /d 0 0.1 0.2 0.3 0.4 f V C t d t˜ = 0 t˜ = 33 t˜ = 66 t˜ = 98 (b) Time-averaged mean flow field 0 2 4 6 8 10 v /d 0 0.1 0.2 0.3 0.4 f V C t d t˜ = 0 t˜ = 33 t˜ = 66 (f ) Time-dependent flow field 10−4 10−2 1 102 104 m 10−4 10−3 10−2 0.1 1 f M C t t˜ = 0 t˜ = 33 t˜ = 66 t˜ = 98 (c) Time-averaged mean flow field 10−4 10−2 1 102 104 m 10−4 10−3 10−2 0.1 1 f M C t t˜ = 0 t˜ = 33 t˜ = 66 (g) Time-dependent flow field 0 2 4 6 8 10 12 14 16 18 d /d 0 0.1 0.2 0.3 f D C t d t˜ = 0 t˜ = 33 t˜ = 66 t˜ = 98 (d) Time-averaged mean flow field 0 2 4 6 8 10 12 d /d 0 0.1 0.2 0.3 f D C t d t˜ = 0 t˜ = 33 t˜ = 66 (h) Figure 5.9: Comparison between the time evolutions (the colour-scale used is the same to that used in figure 5.6) of the p.d.f.s in the case of the cluster of virtual particles initially seeded in the core of the time-dependent (a–d) (see figure 5.6c) and the time-averaged (e–h) (see figure 5.6f ) velocity fields, for: (a,e) the dimensionless lateral distance between pairs of particles fHCt d computed using (5.10) and (5.11) with n = 3721 and N = 100; (b,f ) the dimensionless streamwise distance between pairs of particles fVCt d computed using (5.10) and (5.11) with n = 3721 and N = 100; (c,g) the ratio of the lateral distance to the streamwise distance between pairs of particles fMCt computed using (5.10) and (5.11) with n = 3721 and N = 100 (in a log–log plot); and (d,h) the dimensionless distance between pairs of particles fDCtd computed using (5.10) and (5.11) with n = 3721 and N = 100. 140 5.5 Discussion and Conclusion consists of seeding and tracking clusters of virtual particles (or passive tracers) in experimentally-measured velocity fields. As we discussed in the previous chapter (see § 4.1.2), there are numerous advantages to using this technique. The spatial and temporal resolutions are only limited by the resolution of the acquisition tech- nique used to measure the velocity field.1 Virtual particle tracking can potentially be applied to any laboratory flows, with a possible range of Reynolds numbers, Schmidt numbers or Prandtl numbers far exceeding the capabilities of numerical simulations. A large quantity of virtual particles can be seeded instantaneously in the flow field, with any arbitrary initial distribution, and then tracked over a spatial range only limited by the size of the measured velocity field. One could ar- gue that virtual particle tracking is not adapted to the study of three-dimensional flow fields. With only a two-dimensional velocity field of a three-dimensional flow field, it is true that virtual particle tracking cannot give meaningful information, because the trajectories of real Lagrangian particles are also three-dimensional. We believe that the recent development of volumetric particle image velocimetry to measure the three components of the velocity in three-dimensional domains (see e.g. Kitzhofer et al., 2011; Cierpka & Kaehler, 2012, for recent reviews) can address this shortcoming. The flow in quasi-two-dimensional jet is appropriate for the application of par- ticle tracking velocimetry because the three-dimensionality of the flow can be considered insignificant in the first order. In § 4.1.2 we report that the mean di- vergence of the flow is small compared with the mean vorticity. Moreover, Dracos et al. (1992) found that the flow of quasi-two-dimensional jets is primarily gov- erned by a two-dimensional inverse cascade of turbulence, except at scales of the order of (or less than) the gap width of the tankW . Therefore, we believe that par- ticle tracking velocimetry can give physically meaningful information about the dispersion in quasi-two-dimensional jets. However, the three-dimensional small- scale turbulence, typically of the order of W = 1cm or less, cannot be adequately resolved in this study, with only a two-dimensional velocity field. Bearing in mind the limited spatial resolution of our data, we have probed the large-scale dispersion of the (large-scale) eddy and core structures of the flow. 1It can be noted that particle image velocimetry, a common technique to measure veloc- ity fields, is considered technically less demanding than experimental Lagrangian particles tracking techniques, such as particle tracking velocimetry or other optical particle tracking techniques (Kitzhofer, Nonn & Bru¨cker, 2011) 141 5 Two-point statistics for turbulent relative dispersion The time evolution of the probability distributions of key two-point properties (such as the lateral distance, the streamwise distance, the Euclidean distance and the ratio of the lateral distance to the streamwise distance between two points) in the main structures of quasi-two-dimensional jets has shown different behaviours for the different parts of the flow. We compare the results of the two-point statis- tics obtained in the time-dependent velocity field with results obtained in the time-averaged velocity field of the same jet and results obtained with simple ge- ometrical distributions of points (a circle, an ellipse and a square). From the study of these simple geometrical distributions, we have been able to understand how the variation in time of general shape characteristics of the distribution af- fects the p.d.f.s of the two-point properties. In particular, we have been able to measure that, in the eddy, the distribution of particles disperses slowly and in a rather axisymmetric manner. At the interface between the core and the eddy, the distribution of particles stretches considerably in the streamwise direction at a high rate. This is accompanied by thinning of the particle cluster. In the core of the jet, the particle distribution disperses slowly in the cross-jet direction and splits along the jet axis. Finally, we believe that the comparison between the p.d.f.s for the time-averaged flow field and the p.d.f.s for the time-dependent flow field demonstrates the intense stirring (and potentially the resulting vigorous tur- bulent mixing) occurring within the eddy and, to some extent, at the interface between the eddy and the core. This aspect is revealed by the rapid displacement through time of the peaks in the distribution of fDEt (t) (the time evolution of the p.d.f. of the distance between two particles initially seeded in the eddy) for the time-dependent velocity field of the eddy. The chaotic dynamics of the turbulent flow in the eddy strongly perturbs the distribution of the virtual particles, which manifests itself in the time evolution of the p.d.f. for the separation distance between particles. Future research about the turbulent relative dispersion of the flow of quasi-two- dimensional jets could investigate the ideas of Richardson (1926) and Batchelor (1952) to describe the relative dispersion in the jet by a differential equation based on the p.d.f.s of two-point properties. In Chapter 3, we propose a model for the transport and streamwise dispersion in the jet, based on the Eulerian description of the flow. Forming the connection between the Eulerian and the Lagrangian descriptions of the turbulent dispersion could provide invaluable insight in the 142 5.5 Discussion and Conclusion physics of anisotropic turbulent processes. One particular question of interest is to relate the streamwise turbulent eddy diffusivity KdM1/20 z1/2 in the general effective advection–diffusion equation (3.15) (obtained using a mixing length hy- pothesis) to the p.d.f. of the streamwise distance between two points obtained directly from virtual particle tracking, in an effort to identify and parameterize the cumulative quantitative effect of the complex time-dependent flow on streamwise dispersion. Another possible avenue of research would be to improve the spatial resolution of the velocity field, and perhaps to measure a truly three-dimensional velocity field of the flow. With a fully resolved velocity field in time (i.e. resolving Kol- mogorov time scale τηK ≈ 40 ms) and in space (i.e. resolving Kolmogorov length scale ηK ≈ 0.2 mm), we could explore, for instance, the two-point dispersion model of Batchelor (1950). As Bourgoin et al. (2006) pointed out, there is a need for more experimental evidence. A comparison between the results of two- point statistics for the flow field of quasi-two-dimensional jets with the results for three-dimensional turbulent flows could shed new light on the physics of turbulent relative dispersion. 143 Chapter 6 Flow induced by a quasi-two-dimensional jet in a confined rectangular domain 6.1 Introduction A turbulent momentum jet induces a flow towards the jet in the surrounding ambient fluid. The entrainment of the ambient fluid, which is the result of a complex turbulent dynamics at the boundary of the jet, was modelled by Morton et al. (1956). Using dimensional analysis, they related the lateral velocity of the entrained fluid at the boundary of the jet as simply being proportional to the time-averaged maximum axial (or streamwise) velocity in the jet. Contrary to round axisymmetric jets, the velocity of the fluid entrained by a plane jet does not decay with distance away from the jet axis. The flow induced by plane 145 6 Flow induced by a jet in a confined domain jets is important in industrial applications such as chemical reactors and mixing chambers (Jirka & Harleman, 1979). As we discuss in the introduction of Chapter 2 (see § 2.1) rivers flowing into lakes or oceans can be modelled as quasi-two-di- mensional turbulent jets (Giger et al., 1991; Dracos et al., 1992; Rowland et al., 2009). Studying the flow induced by rivers emerging into lakes or oceans, Joshi & Taylor (1983) revealed the impact on the local sediment transport and the possible coastal erosion. Taylor (1958) calculated the stream function of the flow in the ambient of plane jets and axisymmetric jets, when emerging from either a plane wall into a semi-infinite domain or directly into unbounded space, and for both buoyant and non-buoyant jets. Assuming an inviscid incompressible potential flow in the am- bient, using a slip boundary condition at the wall (if present) and modelling the jet as a distribution of sinks, he solved Laplace’s equation to obtain the stream function. However, Schneider (1981) demonstrated that the hypothesis of an in- viscid fluid and the use of a slip boundary condition at the wall gave an incorrect result for the streamlines in the ambient of axisymmetric jets. Comparing Tay- lor (1958)’s analytical solution and Schneider (1981)’s numerical solution in the case of a laminar axisymmetric jet, Zauner (1985) confirmed experimentally the importance of viscosity in the ambient flow and the need to satisfy the condition of zero tangential velocity at the wall. Nevertheless, inviscid potential theory using slip boundary conditions at the walls is still valid in the case of the flow induced by plane turbulent jets, because the Reynolds number in the ambient flow is comparable to the jet Reynolds number (Schneider, 1981). The flow induced by a turbulent jet can also influence the axial momentum flux of the jet (Kotsovinos, 1978; Schneider, 1985; Kotsovinos & Angelidis, 1991). According to Kotsovinos & Angelidis (1991), this influence occurs through two factors. The momentum flux of the induced flow can contribute positively or neg- atively to the jet momentum flux depending on the angle between the streamlines of the induced flow and the direction of the jet flow at the jet boundary. The pres- sure field at the boundary of the jet contributes negatively to the jet momentum flux. Therefore, predicting the streamlines of the induced flow is important to determine the rate of change of the jet momentum flux. In the case of a plane jet emerging from a wall into a semi-infinite domain, the streamlines of the induced potential flow form two sets of confocal parabolas with axes perpendicular to the 146 6.1 Introduction jet axis (Taylor, 1958). Thus, the streamlines are opposed to the jet flow and the jet momentum flux slowly decays with distance. However, the more realistic case of a plane jet emerging from a wall into a confined domain does not seem to have been solved in the literature. The case of a plane jet emerging from a wall into a domain confined in the axial, lateral and spanwise directions is a common problem, because, in practice, semi-infinite domains or fully unbounded domains (as assumed by Taylor, 1958; Schneider, 1981) do not exist. In his model, Schneider (1981) analysed how the angle between the jet axis and the wall (from which the jet emerged) influences the streamlines of the induced flow. Revuelta, Sa´nchez & Lin˜a´n (2002) investigated numerically the case of an axisymmetric laminar jet confined in an axisymmet- ric domain. They predicted the size and the induced pressure drop of a long recirculating region surrounding the jet, before the jet expands across the whole domain. Jirka & Harleman (1979) studied experimentally and theoretically the stability and mixing of plane jets confined in the axial direction, but unconfined in the lateral (or cross-stream) direction. For non-buoyant jets, they observed on both sides of the jet the formation of alternating recirculation cells. The cell closer to the jet is driven by two distinct mechanisms. As the vertical upward jet impinges on the free surface at the top, the flow spreads laterally outwards along the free surface. Along the bottom boundary, the flow is driven inwards by the jet entrainment process. The size and the total mass flow of the cell are controlled by the growth characteristics of the jet and the associated entrainment mechanism. Moreover, Jirka & Harleman (1979) noted that if passive tracers are injected at the source of the jet, their concentration in the jet increases due to the recirculation in the cell. In this chapter, we are interested in the flow induced by a quasi-two-dimen- sional turbulent jet emerging from a plane wall into a fully confined domain (see experimental apparatus presented in figure 2.1). In this domain, the distance between the source and the lateral or axial boundaries is much larger than the nozzle width, d = 0.5 cm. As we discuss in Chapter 2, the flow in the jet does not seem to be affected by the streamwise confinement for 0 ≤ z/d ≤ hi/d ≈ 120, where hi is the height at which the impingement region starts (see § 2.4). For hi/d ≤ z/d ≤ hf/d = 183 (where hf is the height of the free surface), the flow experiences a transition as it impinges on the free surface. The vertical upward 147 6 Flow induced by a jet in a confined domain flow from the jet spreads laterally outwards, symmetrically with respect to the jet axis. Two overflows, located close to the lateral walls, maintain the free surface at a constant hf = 91.5 cm. A large portion of the flow spreading along the free surface recirculates inside the tank as it reaches the lateral boundaries and produces the counterflow mentioned in § 2.4 (with volume flux Qr). Somewhat similarly to the recirculation cells observed by Jirka & Harleman (1979), the flow in our experimental apparatus also displays a recirculation cell on either side of the jet, but in our case, the recirculation cells are confined laterally by rigid walls. In this study, we do not model the flow in the impingement region located directly above the jet nor the recirculation flow at the lateral boundaries near the free surface (i.e. the region ranging hi ≤ z ≤ hf in the streamwise direction and spanning the entire domain in the lateral (or x-) direction and the spanwise (or y-) direction). We only model the flow on the left-hand side of the jet axis (the flow on the right-hand side can be obtained by symmetry), before the transition from a jet flow to an impingement flow. We assume that the jet is a distribution of sinks. The domain of study, which we designate as Ds, ranges 0 ≤ x/d ≤ xj/d = 90 in the lateral or cross jet direction (where xj represents the lateral coordinate of the jet nozzle, considering the origin of the domain (x = 0, y = 0, z = 0) at the bottom left-hand-side corner of the tank), 0 ≤ z ≤ hi in the streamwise direction, and spans the entire domain in the spanwise (or y-) direction, W/2 ≤ y ≤ W/2 (where W = 1 cm is the gap width). We distinguish two aspect ratios in this study: the aspect ratio of the inner dimensions of the experimental apparatus, (2xj)/hf ≈ 1; and the aspect ratio of the domain Ds, ζ = xj/hi = 3/4, or the jet aspect ratio. In § 6.2, we develop a model of the ambient flow field in the domain Ds using two-dimensional potential theory. We present the results for the potential field, the stream function, and the velocity field. In § 6.3, we compare the theoretical results with results from dyed quasi-two-dimensional turbulent jets and particle image velocimetry experiments for the stream function, the velocity field, the volume flux and the momentum flux of the induced flow. We draw our conclusions in § 6.4. 148 6.2 Potential flow model 6.2 Potential flow model 6.2.1 Description of the entrainment problem We consider a similar experimental apparatus to that which is depicted in figure 2.1. We model the ambient flow field at the left-hand side of a quasi-two-dimen- sional turbulent jet using potential theory. We assume a steady laminar plane flow in the rectangular domain Ds. In Cartesian coordinates (x, z) with the origin at the bottom left-hand corner of the inside of the experimental apparatus,1, the domain is bounded at the bottom (z = 0m) and on the left-hand side (x = 0m) by rigid walls. For the top boundary, we do not wish to consider the impingement region observed by Jirka & Harleman (1979) near the free surface (located at z = hf = 0.915 m). So, instead of choosing the free surface as the top boundary, we choose the height of transition between the jet region and the impingement region, which is at z = hi ≈ 0.6 m (see § 2.4 and figure 2.4). On the right-hand side, the domain is delimited by the jet boundary, which we assume to be along the jet axis (Taylor, 1958) at x = xj ≈ 0.45 m. Although the distance between the jet boundary and the axis increases with z, we believe that this assumption is valid because the jet velocity spread rate b(z) (which is of the same order of magnitude as the jet boundary) is much smaller than the lateral dimension of the domain at all height: b/xj ≤ 0.2 for 0 ≤ z ≤ hi according to (2.5a) using an entrainment coefficient α = 0.068. Since we assume a two-dimensional plane flow in the domain (similarly to the model in § 2.4), we do not consider the boundaries in the spanwise (or y-) direction in this model. The domain Ds is delimited by: 0 ≤ x ≤ xj and 0 ≤ z ≤ hi. We use the following boundary conditions for the velocity field u = (u, w). The normal velocity vanishes at the walls, u(x = 0, z) = 0 and w(x, z = 0) = 0, and the slip condition applies for the tangential velocity at the walls (Taylor, 1958). At z = hi, we assume a uniform constant line source, the flux per unit length is w(x, z = hi) = −ℓ, (6.1) where ℓ ≥ 0 is a constant, which is determined below. The boundary condition (6.1) represents the recirculation of the flow in the experimental apparatus. The 1Note that in the model developed in § 2.4 the origin of the domain is in the middle of the bottom wall of the experimental apparatus. 149 6 Flow induced by a jet in a confined domain last boundary condition corresponds to the jet at x = xj. Similarly to Taylor (1958), we assume that, in the ambient, the influence of the jet can be considered as a line sink along the z-axis of strength j(z), varying with height, j(z) = u(x = xj, z) = αwm, (6.2) where we use the entrainment assumption of Morton et al. (1956), and where wm is defined by (2.5b) (which assumes a constant momentum flux). Hence, the strength of the line sink is j(z) = Kj √ d√z − z0 , with Kj = ( αM0 2 √ 2d )1/2 , (6.3) where M0 is the initial momentum of the jet at z = 0, and z0 the space virtual origin defined in (2.6). By continuity, the volume flux of the line source must equal the volume flux of the line sink ∫ xj 0 ℓ dx = ∫ hi 0 j(z) dz. (6.4) Finally, we assume that the flow is irrotational and incompressible in the do- main. Therefore, the velocity field u = (u, w) derives from a potential ϕ, such that u = ∇ϕ (where ∇ is the gradient operator), which must satisfy Laplace’s equation in the domain: ∇2ϕ = 0 for 0 ≤ x ≤ xj , 0 ≤ z ≤ hi. (6.5) We scale all spatial variables with the height of the impingement region hi, such that x˜ = x/hi and z˜ = z/hi (where wide tildes denote non-dimensional variables). We define ζ = xj/hi = 3/4, the jet aspect ratio of the domain Ds corresponding to our particular experimental problem. (As mentioned previously, we distinguish the jet aspect ratio ζ from the aspect ratio of the apparatus 2xj/hf ≈ 1.) We scale velocities with Kj (which is proportional to the streamwise velocity at the nozzle) defined in (6.3), such that u˜ = u/Kj and w˜ = w/Kj. We summarize the entrainment problem in figure 6.1. The non-dimensional 150 6.2 Potential flow model potential ϕ˜ is the solution to Laplace’s equation in the domain Ds: ∇˜ 2ϕ˜ = 0 for 0 ≤ x˜ ≤ ζ, 0 ≤ z˜ ≤ 1, (6.6) subject to the Neumann boundary conditions    ∂ϕ˜ ∂x˜ = 0 for x˜ = 0, 0 ≤ z˜ ≤ 1, ∂ϕ˜ ∂x˜ = j˜ for x˜ = ζ, 0 ≤ z˜ ≤ 1, ∂ϕ˜ ∂z˜ = 0 for 0 ≤ x˜ ≤ ζ, z˜ = 0, ∂ϕ˜ ∂z˜ = −ℓ˜ for 0 ≤ x˜ ≤ ζ, z˜ = 1 , (6.7) and with the continuity condition (from (6.4)) ∫ 1 0 j˜(z˜) dz˜ = ζℓ˜, (6.8) where j˜(z˜) = ( d˜ z˜ − z˜0 )1/2 , (6.9) according to (6.3). Therefore, the strength of the line source is, in non-dimensional form, ℓ˜ = 2 √ d˜ ζ (√ 1− z˜0 − √ −z˜0 ) . (6.10) 6.2.2 Decomposition of the problem To simplify our problem and to eventually improve the convergence of the numeri- cal calculation of our analytical solution, we split the non-dimensional potential ϕ˜, by virtue of the superposition principle for linear problems, into two components: ϕ˜ = ϕ˜u + ϕ˜p, (6.11) 151 6 Flow induced by a jet in a confined domain x z 1 ∂ϕ˜ ∂x˜ = 0 0 ∂ϕ˜ ∂z˜ = 0 ζ ∂ϕ˜ ∂z˜ = −ℓ˜ ∂ϕ˜ ∂x˜ = j˜∇˜ 2ϕ˜ = 0 Figure 6.1: Description of the entrainment problem of which ϕ˜ is solution. where ϕ˜u is the solution of a ‘uniform problem’ in the domain Ds, with a non- dimensional uniform line source of strength ℓ˜ at z˜ = 1 and a non-dimensional uniform line sink of strength ζℓ˜ at x˜ = ζ, as described in figure 6.2(a), and where ϕ˜p is defined as a perturbation to this uniform problem. The ‘perturbation problem’ is represented in figure 6.2(b). In the perturbation problem, we have no-flux boundary conditions at x˜ = 0, z˜ = 0 and z˜ = 1, and a varying flux at x˜ = ζ such that ∂ϕ˜p ∂x˜ (x˜ = ζ, z˜) = j˜ (z˜)− ζℓ˜. (6.12) 6.2.3 Solution to the uniform problem ϕ˜u Solving Laplace’s equation in the domain described in figure 6.2(a), the solution to the uniform problem, with a uniform line source at z˜ = 1 and a uniform line sink at x˜ = ζ, is ϕ˜u = 1 2 ( x˜2 − z˜2 ) for 0 ≤ x˜ ≤ ζ, 0 ≤ z˜ ≤ 1. (6.13) As a solution of Laplace’s equation under Neumann boundary conditions, ϕ˜u is the unique solution, to within a constant, to the uniform problem. In a two-dimensional inviscid and incompressible flow, we can also define a 152 6.2 Potential flow model (a) x z 1 ∂ϕ˜u ∂x˜ = 0 0 ∂ϕ˜u ∂z˜ = 0 ζ ∂ϕ˜u ∂z˜ = −ℓ˜ ∂ϕ˜u ∂x˜ = ζℓ˜∇˜ 2ϕ˜u = 0 (b) x z 1 ∂ϕ˜p ∂x˜ = 0 0 ∂ϕ˜p ∂z˜ = 0 ζ ∂ϕ˜p ∂z˜ = 0 ∂ϕ˜p ∂x˜ = j˜ − ζℓ˜∇˜ 2ϕ˜p = 0 Figure 6.2: We decompose the entrainment problem into two problems (see equation (6.11)): (a) a uniform problem of which ϕ˜u is solution; (b) a perturbation problem of which ϕ˜p is solution. stream function ψ such that ∇ψ ·∇ϕ = 0, (6.14) i.e. the streamlines are orthogonal to the equipotential lines in the domain. The corresponding non-dimensional stream function ψ˜u for the uniform problem de- scribed in figure 6.2(a) is ψ˜u = x˜z˜ for 0 ≤ x˜ ≤ ζ, 0 ≤ z˜ ≤ 1. (6.15) In figures 6.3(a,b), we show the non-dimensional potential ϕ˜u and the non- dimensional stream function ψ˜u, respectively, for the uniform problem described in figure 6.2(a). For the aspect ratio, we use ζ = xj/hi = 3/4, the jet aspect ratio of our particular case. The flow field in the right-hand-side half of the tank can be found by symmetry with respect to the jet axis. As we can see in figure 6.3(b), the streamlines are hyperbolas. By definition we have u = ∇ϕ, so the velocity field of the uniform problem u˜u = (u˜u, w˜u) can be derived from the potential ϕ˜u described in (6.13). We find u˜u = x˜, w˜u = −z˜ for 0 ≤ x˜ ≤ ζ, 0 ≤ z˜ ≤ 1. (6.16a,b) 153 6 Flow induced by a jet in a confined domain x˜ z˜ 0 0.2 0.4 0.6 0 0.2 0.4 0.6 0.8 1.0 (a) x˜ z˜ 0 0.2 0.4 0.6 0 0.2 0.4 0.6 0.8 1.0 (b) Figure 6.3: (a) Non-dimensional potential ϕ˜u (defined by (6.13)), and (b) non- dimensional stream function ψ˜u (defined by (6.15)) for the uniform problem described in figure 6.2(a), using ζ = xj/hi = 3/4. The velocity components are linear in their coordinate direction and constant in the orthogonal direction. The non-dimensional velocity field (u˜u, w˜u) is presented in figures 6.4(a,b), respectively. x˜ z˜ 0 0.2 0.4 0.6 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0 0.2 0.4 0.6 0.8 1.0(a) x˜ z˜ 0 0.2 0.4 0.6 -0.9 -0.8 -0.7 -0.6 -0.5 -0.4 -0.3 -0.2 -0.1 0 0.2 0.4 0.6 0.8 1.0 (b) Figure 6.4: (a) Non-dimensional lateral velocity u˜u (defined by (6.16a)), and (b) non-dimensional streamwise velocity w˜u (defined by (6.16b)) for the uniform problem described in figure 6.2(a), using ζ = xj/hi = 3/4. 154 6.2 Potential flow model 6.2.4 Solution to the perturbation problem ϕ˜p A solution to the perturbation problem can be found by the method of separation of variables. The solution ϕ˜p consists of an infinite linear combination of the product of hyperbolic cosines and cosines, i.e. ϕ˜p = ∞∑ n=1 An cosh (nπx˜) cos (nπz˜) for 0 ≤ x˜ ≤ ζ, 0 ≤ z˜ ≤ 1, (6.17) where An are coefficients which can be determined using the non-homogeneous boundary condition at x˜ = ζ, ∂ϕ˜p ∂x˜ ∣∣∣∣ x˜=ζ = ∞∑ n=1 Annπ sinh (nπζ) cos (nπz˜) = j˜(z˜)− ζℓ˜ for 0 ≤ z˜ ≤ 1, (6.18) according to (6.12). We define the coefficients Bn such that Bn = Annπ sinh (nπζ) for n ≥ 1. (6.19) Re-writing equation (6.18), ∞∑ n=1 Bn cos (nπz˜) = j˜(z˜)− ζℓ˜ for 0 ≤ z˜ ≤ 1, (6.20) we can clearly see that the coefficients Bn are the Fourier coefficients of an even function E˜ defined as E˜(z˜) = j˜(|z˜|)− ζℓ˜ for − 1 ≤ z˜ ≤ 1. (6.21) Therefore, we can calculate the coefficients Bn as follows Bn = ∫ 1 −1 E˜(z˜) cos (nπz˜) dz˜ for n ≥ 1, (6.22) which simplifies to Bn = 2 ∫ 1 0 j˜(z˜) cos (nπz˜) dz˜, (6.23) because j˜(|z˜|) and cos (nπz˜) are even functions for −1 ≤ z˜ ≤ 1 and ζℓ˜ cos (nπz˜) integrates to zero in the interval −1 ≤ z˜ ≤ 1 for n ≥ 1. Using equation (6.9), we 155 6 Flow induced by a jet in a confined domain have Bn = 2 √ d˜ ∫ 1 0 cos (nπz˜)√ z˜ − z˜0 dz˜. (6.24) Then, applying the transformation q = nπ(z˜ − z˜0) we find Bn = 2 √ d˜√nπ ( cos (nπz˜0) ∫ npi(1−z˜0) −npiz˜0 cos q√q dq − sin (nπz˜0) ∫ npi(1−z˜0) −npiz˜0 sin q√q dq ) . (6.25) Finally, we can apply another transformation s = √ 2q/π, which gives Bn = 2 √ 2d˜√n ( cos (nπz˜0) [ C(y) ]√2n(1−z˜0) √ −2nz˜0 − sin (nπz˜0) [ S(y) ]√2n(1−z˜0) √ −2nz˜0 ) (6.26) for n ≥ 1, where we have introduced the Fresnel C and S integrals (see e.g. Abramowitz & Stegun, 1972) defined as C(y) = ∫ y 0 cos (π 2 s2 ) ds and S(y) = ∫ y 0 sin (π 2 s2 ) ds. (6.27) It can be noted that the coefficient B0 equals zero from the condition of continuity stated in equation (6.8). Therefore, we have found a unique solution ϕ˜p (defined by equations (6.17), (6.19) and (6.26)) to the perturbation problem described in figure 6.2(b). According to (6.14), the corresponding non-dimensional stream function ψ˜p for the perturbation problem described in figure 6.2(b) is ψ˜p = ∞∑ n=1 An sinh (nπx˜) sin (nπz˜) for 0 ≤ x˜ ≤ ζ, 0 ≤ z˜ ≤ 1, (6.28) where the coefficients An are given by (6.19) and (6.26). We present in figures 6.5(a,b) the non-dimensional potential ϕ˜p and the non- dimensional stream function ψ˜p, respectively, for the perturbation problem de- scribed in figure 6.2(b). We compute the series ϕ˜p (defined by equations (6.17), (6.19) and (6.26)) and ψ˜p (defined by equations (6.28), (6.19) and (6.26)) numer- ically for nmax = 100, the (finite) number of terms of both series. We use the jet aspect ratio ζ = xj/hi = 3/4 of our particular case, and the space virtual origin z0 = −4.7 d (computed from (2.6) using α = 0.068, < M > / ( Q02/d ) = 0.55). Similarly to u˜u, the velocity field of the perturbation problem u˜p = (u˜p, w˜p) can 156 6.2 Potential flow model x˜ z˜ 0 0.2 0.4 0.6 0 0.2 0.4 0.6 0.8 1.0 (a) x˜ z˜ 0 0.2 0.4 0.6 0 0.2 0.4 0.6 0.8 1.0 (b) Figure 6.5: (a) Non-dimensional potential ϕ˜p (defined by equations (6.17), (6.19) and (6.26) with nmax = 100 the number of terms of the series, and z0 = −4.7 d the space virtual origin), and (b) non-dimensional stream function ψ˜p (defined by (6.28), (6.19) and (6.26) with nmax = 100, the number of terms of the series, and z0 = −4.7 d the space virtual origin) for the perturbation problem described in figure 6.2(b), using ζ = xj/hi = 3/4. be derived from the potential ϕ˜p defined by equations (6.17), (6.19) and (6.26). We find for 0 ≤ x˜ ≤ ζ, 0 ≤ z˜ ≤ 1, u˜p = ∞∑ n=1 nπAn sinh (nπx˜) cos (nπz˜), (6.29) w˜p = − ∞∑ n=1 nπAn cosh (nπx˜) sin (nπz˜), (6.30) where the coefficients An are given by (6.19) and (6.26). We present the non- dimensional velocities u˜p and w˜p for the perturbation problem in figures 6.6(a,b), respectively. Similarly to ϕ˜p and ψ˜p, we compute the series u˜p (defined by equa- tions (6.29), (6.19) and (6.26)) and w˜p (defined by equations (6.30), (6.19) and (6.26)) numerically for nmax = 100, the number of terms of both series. We use the aspect ratio ζ = xj/hi = 3/4, and the space virtual origin z0 = −4.7 d (com- puted from (2.6) using α = 0.068, < M > / ( Q02/d ) = 0.55). We note that both the lateral and the streamwise velocities are strongly affected by the origin of the 157 6 Flow induced by a jet in a confined domain x˜ z˜ 0 0.2 0.4 0.6 -0.05 0 0.05 0.1 0.15 0.2 0.25 0 0.2 0.4 0.6 0.8 1.0 (a) x˜ z˜ 0 0.2 0.4 0.6 -0.1 -0.08 -0.06 -0.04 -0.02 0 0.2 0.4 0.6 0.8 1.0 (b) Figure 6.6: (a) Non-dimensional lateral velocity u˜p (defined by (6.29), (6.19) and (6.26) with nmax = 100 the number of terms of the series, and z0 = −4.7 d the space virtual origin), and (b) non-dimensional streamwise velocity w˜p (defined by (6.30), (6.19) and (6.26) with nmax = 100 the number of terms of the series, and z0 = −4.7 d the space virtual origin) for the perturbation problem described in figure 6.2(b), using ζ = xj/hi = 3/4. line sink at (x˜ = ζ, z˜ = 0). This is due to the singularity at the virtual origin z˜ = z˜0 < 0. We can see in figure 6.6(a) that the lateral velocity u˜p is maximum at the bottom right-hand corner, (x˜ = ζ, z˜ = 0). In figure 6.6(b), we can see that, along the right-hand side boundary x˜ = ζ, the streamwise velocity w˜p sharply decreases from the bottom boundary z˜ = 0 to approximately z˜ ≈ 0.1. Then, w˜p slowly increases again for z˜ > 0.1, to eventually vanish at the top boundary z˜ = 1. Since there is no flux inwards or outwards at the boundaries x˜ = 0, z˜ = 0 and z˜ = 1, by continuity, the total integrated flux along the right-hand side boundary at x˜ = ζ must also be zero, ∫ 1 0 ∂ϕ˜p ∂x˜ (x˜ = ζ, z˜) dz˜ = 0, (6.31) according to equation (6.8) and (6.12), thus ∂ϕ˜p ∂x˜ (x˜ = ζ, z˜) = √ d˜ ( 1√ z˜ − z˜0 − 2 (√ 1− z˜0 − √ −z˜0 )) , (6.32) according to (6.12), with j˜(z˜) defined by (6.9) and ℓ˜ defined by (6.10). There is 158 6.2 Potential flow model a local non-uniform flux along the right-hand side boundary at x˜ = ζ oriented inwards for z ≥ z˜c and outwards for z ≤ z˜c, where z˜c is defined by ∂ϕ˜p ∂x˜ (x˜ = ζ, z˜c) = 0. (6.33) Hence, the non-dimensional height z˜c is z˜c = 1 4 (√ 1− z˜0 − √ −z˜0 )2 + z˜0. (6.34) We show in figure 6.7 the distribution of the non-dimensional perturbation flux ∂ϕ˜p/∂x˜(x˜ = ζ, z˜) = j˜(z˜)− ζℓ˜ along the right-hand side boundary. The analytical formula (6.32), for the flux along the right-hand side boundary, is plotted with a solid curve. We can see the relative importance of the line sink and the line source. The varying line sink j˜(z˜) is stronger than the line source ζℓ˜ and oriented outwards (i.e. in the positive direction) for z˜ < z˜c (where z˜c ≈ 0.33, defined in (6.34), is marked with dashed lines). The flux ∂ϕ˜p/∂x˜(x˜ = ζ, z˜) increases steeply approaching the origin, i.e. as z˜ → 0, as expected near the singularity at z˜ = z˜0 < 0. For z˜ > z˜c, the uniform line source dominates and the local flux is oriented inwards (i.e. in the negative direction). The exchange flow at the right- hand-side boundary is also depicted clearly by the distribution of the streamlines shown in figure 6.5(b). We compute the potential ϕ˜p (defined in (6.17)), the stream function ψ˜p (defined in (6.28)), and the velocity field u˜p = (u˜p, w˜p) (defined in (6.29) and (6.30), respectively) numerically and we truncate their infinite series to a finite number of terms nmax. In order to test the accuracy of the numerical computation of these truncated series, we compare the numerical computation of the truncated series of the lateral velocity along the right-hand side boundary, designated by u˜nmaxp (x˜ = ζ, z˜) for the first nmax terms of the series (6.29), with the analytical formula (6.32) of the flux imposed at the same boundary, the perturbation flux ∂ϕ˜p/∂x˜(x˜ = ζ, z˜). We measure the mismatch (introduced by the truncation) between the truncated series and the analytical formula by calculating their standard deviation, as a function of the number of terms of the series nmax. The standard deviation is 159 6 Flow induced by a jet in a confined domain z˜ ϕ˜p ∂x˜ (z˜) Analytical formula (6.32) Numerical truncated series (6.36) z˜c = 0.33 -0.1 0 0.1 0.2 0.3 0 0.2 0.4 0.6 0.8 1.0 Figure 6.7: Distribution of the non-dimensional flux along the right-hand side bound- ary. The analytical formula of the imposed boundary condition, ∂ϕ˜p/∂x˜(x˜ = ζ, z˜) defined by (6.32), is plotted with a solid curve. We plot the numerical truncated se- ries of the flux u˜nmaxp (defined in (6.36), with nmax = 100 the number of terms of the truncated series, and z0 = −4.7 d the space virtual origin) with pluses. The location z˜c (defined by (6.34)), where the flux vanishes and changes sign, is marked with dashed lines. defined, in discrete form, by σp(nmax) = ( 1 Nz Nz∑ i=1 ( u˜nmaxp (ζ, z˜i)− ∂ϕ˜p ∂x˜ (ζ, z˜i) )2)1/2 , (6.35) where z˜i are linearly distributed from z˜1 = 0 to z˜Nz = 1, with Nz = 1001 the discretization number, and u˜nmaxp (ζ, z˜i) = nmax∑ n=1 Bn cos (nπz˜i), for all 1 ≤ i ≤ Nz, (6.36) according to (6.29) and (6.19), where the coefficients Bn are described by (6.26) for n ≥ 1. We plot in figure 6.8(a) the standard deviation σp (defined in (6.35)) against 160 6.2 Potential flow model nmax. We can see that σp decreases rapidly and monotonically as nmax increases. Thus, the numerical truncated series u˜nmaxp (ζ, z˜i) converges rapidly towards the analytical formula for ∂ϕ˜p/∂x˜(ζ, z˜i). We find that for nmax = 100 the standard deviation is very small, σp ≤ 0.1 %. As a comparison with the analytical formula (6.32) for the perturbation flux ∂ϕ˜p/∂x˜(x˜ = ζ, z˜), we plot in figure 6.7 the numer- ical truncated series u˜nmaxp (ζ, z˜) (defined in (6.36), for nmax = 100) with pluses. As expected, the match is excellent. To study the smoothness of the (non-truncated) Fourier series u˜∞p (ζ, z˜), we present in a log–log plot in figure 6.8(b) the coefficients of the series Bn (described by (6.26) and shown with pluses), for 1 ≤ n ≤ 200. As we can observe, the coefficients Bn appear to fall off like O(1/n2) (the function 1/n2 is plotted with a red line) rather than O(1/n) (the function 1/n is plotted with a black line). A decrease of O(1/n2) means that the Fourier series u˜∞p (ζ, z˜) is continuous while its first derivative (with respect to z˜) is discontinuous over the interval 0 ≤ z˜ ≤ 1. The Fourier series u˜∞p (ζ, z˜) (defined in (6.36), with nmax = ∞) converges precisely to the even continuous function E(z˜) = j˜(|z˜|) − ζℓ˜ (except perhaps on a set of measure zero, see e.g. Ko¨rner, 1988) defined in the periodic interval −1 ≤ z˜ ≤ 1, according to (6.20) and (6.21). The function E is continuous over this interval, but its first derivative is discontinuous at z˜ = 0 mod Tp = 2 (where Tp is the period of the function E) owing to the absolute value, and at z˜ = 1 mod Tp by construction of the periodic function E. Since the Fourier series u˜∞p (ζ, z˜) also satisfies Dirichlet’s conditions (see e.g. Kahane & Lemarie´-Rieusset, 1998), then it converges to the analytical formula (6.32) for the right-hand side boundary condition ∂ϕ˜p/∂x˜(x˜ = ζ, z˜), for all point 0 ≤ z˜ ≤ 1, with a smoothness of order 2. We believe that the numerical computations of the truncated series for the potential ϕ˜p (defined in (6.17)), the stream function ψ˜p (defined in (6.28)), and the velocity field u˜p = (u˜p, w˜p) (defined in (6.29) and (6.30), respectively) should all be sufficiently accurate with nmax = 100. 6.2.5 Solution to the entrainment problem According to the superposition principle (6.11), we can combine the potential for the uniform solution ϕ˜u (defined by (6.13)) with the potential for the perturbation solution ϕ˜p (defined by (6.17), (6.19) and (6.26)). We find a unique analytical 161 6 Flow induced by a jet in a confined domain nmax σ p 0 50 100 150 200 0 0.01 0.02 0.03 0.04 0.05 (a) n B n Coefficients Bn 1/n2 1/n 1 10 100 10−4 10−2 1 (b) Figure 6.8: (a) Standard deviation σp (defined in (6.35)) between the truncated se- ries u˜nmaxp (ζ, z˜) (defined in (6.36)) and the analytical formula ∂ϕ˜p/∂x˜(ζ, z˜i) (defined in (6.32)). (b) Log-log plot of the coefficients Bn (defined in (6.26) and plotted with pluses) of the series u˜nmaxp (ζ, z˜) (defined in (6.36)) versus n. The function 1/n is plotted with a black curve and the function 1/n2 is plotted with a red curve. solution, to within a constant, for the potential of the entrainment problem ϕ˜ = 1 2 ( x˜2 − z˜2 ) + ∞∑ n=1 An cosh (nπx˜) cos (nπz˜) for 0 ≤ x˜ ≤ ζ, 0 ≤ z˜ ≤ 1. (6.37) where An are given by (6.19) and (6.26). Again, applying the superposition principle, the corresponding stream function is ψ˜ = x˜z˜ + ∞∑ n=1 An sinh (nπx˜) sin (nπz˜) for 0 ≤ x˜ ≤ ζ, 0 ≤ z˜ ≤ 1, (6.38) where the coefficients An are given by (6.19) and (6.26). We show in figures 6.9(a,b) the non-dimensional potential ϕ˜ (defined by (6.37), (6.19) and (6.26)) and the non-dimensional stream function ψ˜ (defined by (6.38), (6.19) and (6.26)) for the entrainment problem described in figure 6.1. We com- pute the series ϕ˜ and ψ˜ numerically for nmax = 100, the number of terms of both series. Again, we use the jet aspect ratio ζ = xj/hi = 3/4 of our particular case, and the space virtual origin z0 = −4.7 d (computed from (2.6) using α = 0.068, < M > / ( Q02/d ) = 0.55). The streamlines are very similar to those found in the uniform problem in figure 6.3(b). The difference is that they are slightly steeper along the right-hand-side boundary, probably due to the singularity at the virtual origin z˜0 of the line sink. In figure 6.9(b), we also plot with red curves the Taylor’s 162 6.2 Potential flow model solution for the streamlines ψ˜T (Taylor, 1958) ψ˜T = (√ x˜2 + z˜2 − z˜ )1/2 for 0 ≤ x˜ ≤ ζ, 0 ≤ z˜ ≤ 1. (6.39) The stream function ψ˜T corresponds to the two-dimensional incompressible and irrotational flow induced by a plane jet emerging from a plane wall (at x˜ = ζ) into a semi-infinite domain. As we can observe in figure 6.9(b), the streamlines pre- dicted by Taylor (1958) are qualitatively different from our solution. The stream function ψ˜T produces concave streamlines, whereas our solution ψ˜ produces con- vex streamlines. The discrepancy between Taylor’s stream function ψ˜T and our stream function ψ˜ is due to the fact that we consider a fully confined domain, which induces a recirculation flow on either side of the jet, whereas Taylor (1958) considered fully unbounded domains or the case of a jet emerging from a wall into a semi-infinite domain, thus ignoring the possibility of recirculation in the ambient flow. Nevertheless, the streamlines of both stream functions are pointing downwards, i.e. opposite to the jet direction. The velocity field for the entrainment problem is, for 0 ≤ x˜ ≤ ζ, 0 ≤ z˜ ≤ 1, u˜ = x˜+ ∞∑ n=1 nπAn sinh (nπx˜) cos (nπz˜), (6.40) w˜ = −z˜ − ∞∑ n=1 nπAn cosh (nπx˜) sin (nπz˜), (6.41) where the coefficients An are given by (6.19) and (6.26). We show in figures 6.10(a,b) the non-dimensional lateral velocity u˜ (defined by (6.40), (6.19) and (6.26)) and the non-dimensional streamwise velocity w˜ (defined by (6.41), (6.19) and (6.26)) for the entrainment problem described in figure 6.1. We compute the series u˜ and w˜ numerically for nmax = 100, the number of terms of both series. We use the aspect ratio ζ = xj/hi = 3/4, and the space virtual origin z0 = −4.7 d (computed from (2.6) using α = 0.068, < M > / ( Q02/d ) = 0.55). As we can see, the velocity field u˜ is very similar to the velocity field of the uniform problem u˜u presented in figure 6.4(a,b). The influence of the varying line sink mainly appears in its vicinity (i.e. along the right-hand-side boundary at x˜ = ζ). The perturbation of the velocity field is stronger near the source of the jet due to the singularity at the virtual origin (x˜ = ζ, z˜ = z˜0). 163 6 Flow induced by a jet in a confined domain x˜ z˜ 0 0.2 0.4 0.6 0 0.2 0.4 0.6 0.8 1.0 (a) x˜ z˜ 0 0.2 0.4 0.6 0 0.2 0.4 0.6 0.8 1.0 (b) Figure 6.9: (a) Non-dimensional potential ϕ˜ (defined by equations (6.37), (6.19) and (6.26) with nmax = 100 the number of terms of the series, and z0 = −4.7 d the space virtual origin), and (b) non-dimensional stream function ψ˜ (defined by (6.38), (6.19) and (6.26) with nmax = 100, the number of terms of the series, and z0 = −4.7 d the space virtual origin) for the entrainment problem described in figure 6.1, using ζ = xj/hi = 3/4. We also plot with red curves in (b) the Taylor’s solution (defined by (6.39)) for the streamlines of a flow induced by a plane jet emerging from a plane wall (at x˜ = ζ) into a semi-infinite domain (Taylor, 1958). The aspect ratio ζ and the virtual origin z0 influence the flow field of the full problem only through the perturbation part of the problem. The aspect ratio ζ and the virtual origin z0 appear only in the coefficients An of the series, in (6.19) and in (6.26) respectively. We find that, qualitatively, increasing the aspect ratio (i.e. stretching the domain Ds in the x-direction) tends to ‘stretch’ the streamlines in the lateral direction throughout the domain and decrease the angle (with respect to the x-axis) made by the streamlines at the right-hand side boundary (i.e. the ambient fluid enters the jet more perpendicular to the jet axis). Decreasing the aspect ratio produces the opposite effects. Changing the virtual origin has only a local impact along the right-hand side boundary. An increase in |z0| (i.e. the virtual origin is further away below the right-hand side bottom corner) diminishes the influence of the singularity on the flow field, decreases the strength of the sink line, and thus decreases the angle (with respect to the x-axis) made by the streamlines at the right-hand side boundary. Decreasing |z0| (i.e. 164 6.3 Experimental results x˜ z˜ 0 0.2 0.4 0.6 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 0 0.2 0.4 0.6 0.8 1.0(a) x˜ z˜ 0 0.2 0.4 0.6 -0.9 -0.8 -0.7 -0.6 -0.5 -0.4 -0.3 -0.2 -0.1 0 0.2 0.4 0.6 0.8 1.0 (b) Figure 6.10: (a) Non-dimensional lateral velocity u˜ (defined by (6.40), (6.19) and (6.26) with nmax = 100 the number of terms of the series, and z0 = −4.7 d the space virtual origin), and (b) non-dimensional streamwise velocity w˜ (defined by (6.41), (6.19) and (6.26) with nmax = 100 the number of terms of the series, and z0 = −4.7 d the space virtual origin) for the entrainment problem described in figure 6.1, using ζ = xj/hi = 3/4. the virtual origin is closer to the right-hand side bottom corner), has the opposite effect. Our model, and in particular the assumption of a uniform flux at the top boundary, cannot apply for all ranges of aspect ratios. In the case of a very large aspect ratio, Jirka & Harleman (1979) observed alternating recirculation cells on either side of the jet. Thus, the assumption of a uniform flux at the top boundary seems to be incorrect for approximately ζ > 3, from the measurements of the size of the primary recirculation cells made by Jirka & Harleman (1979). In the case of a small aspect ratio, we believe that the width of the jet (whose boundary expands at a rate of approximately 0.22 from the z-axis) could also affect the flux at the top boundary for ζ < 2/3. 6.3 Experimental results 6.3.1 Experimental procedure The experimental procedure is almost identical to the experimental procedure described in § 2.2. The main difference concerns the location of the PIV study areas. To investigate the flow induced by the jet, we choose two PIV study areas 165 6 Flow induced by a jet in a confined domain x z d = 5 mm Constant-head tank Overflow H = 1 m L = 1 m W = 0.01 m u w 2b(z) CCD camera Red filter (dyed jets) Study area 4 (PIV) Study area 5 (PIV) Figure 6.11: Schematic diagram of the experimental apparatus. The two PIV study areas are shown with dashed lines along the left-hand-side inner wall. located on the left-hand-side of the jet axis along the left-hand side inner wall, as shown in figure 6.11. Study area 4 covers a height from z = 0.425 to 0.85 m, while study area 5 covers a height from z = 0 to 0.425m. Both study areas cover a width from x = 0 to 0.425 m. We analyse three steady turbulent jets in each study area. In study area 4, the jet flow rates are 32.2, 36.8 and 40.3 cm3 s−1 . The corresponding jet Reynolds number (based on the jet source characteristics, such that Rej = dws/ν) are in the range 3220 ≤ Rej ≤ 4030. In study area 5, the jet flow rates are 33.5, 37.5 and 40.3 cm3 s−1 . The corresponding jet Reynolds number are in the range 3350 ≤ Rej ≤ 4030. The frequency of image acquisition is set at 60 frames per second for both study areas. The duration of every experiment is approximately 91 s. 6.3.2 Qualitative observations In figure 6.12(a), a superposition of 20 images (i.e. a duration of 0.33 s) of the filming of two experiments, where passive tracers (0.23 mm Pliolite VTAC particles) were mixed with a quasi-two-dimensional jet (Rej = 4030), depicts the tracers as streaks. One experiment is filmed in the PIV study area 5 (see figure 166 6.3 Experimental results 6.11) located at 0 ≤ x/d ≤ 85, 0 ≤ z/d ≤ 85. The other experiment is filmed in the PIV study area 4 (see figure 6.11) located at 0 ≤ x/d ≤ 85, 85 ≤ z/d ≤ 170. In figure 6.12(a), we can visualize the recirculation of the flow on the left-hand side of the jet. Along the right-hand side border of the picture (x/d ≈ 80), we can see three characteristic eddies (located at approximately z/d = 45, z/d = 75 and z/d = 110) of the flow in a quasi-two-dimensional jet (the axis of the jet, not visible in the experiment, is plotted with a dot-dashed line at x = xj.). The top eddy (z/d = 110) is close to the boundary between the jet region and the impingement region, shown with a red dashed line at z = hi = 120 d. As it approaches the free surface (not visible in this experiment, but identified with a black dashed line at z = hf ), the flow spreads laterally. The flow is eventually redirected downwards along the wall at x = 0, before being re-entrained by the jet. The location of the transition height hi = 120 d, between the jet region and the impingement region, was chosen in § 2.4 because we found that the jet was no longer self-similar beyond this height. In figure 6.12(a), we can see that hi also corresponds approximately to the location of the last eddy of the quasi- two-dimensional jet. We actually have not seen eddies beyond this height. The size of the eddies approaching the height hi is around 30 d. This size is still small compared with the distance between the left-hand-side wall and the jet axis, xj/d = 90, but the eddy may start becoming affected by the left-hand-side wall. Jirka & Harleman (1979) found a transition height between the jet region and the impingement region at approximately 85 % of the total depth hf . In our experiment, we find that the ratio between the transition height and the free surface is hi/hf ≈ 65 %. We believe that this discrepancy is due to the lateral confinement, which was insignificant in the experiments of Jirka & Harleman (1979). In figure 6.12(b), we show a picture of a quasi-two-dimensional turbulent dyed jet (Rej ≈ 4000) in the region delimited by −40 ≤ (x − xj)/d ≤ 40 and 0 ≤ z/d ≤ 100. The dye was injected at a constant rate for approximately 3 s in a steady jet, following the experimental procedure detailed in § 2.2.1. We took the picture approximately 4.2 s after the injection of the dye, hence the brighter intensity of the non-dyed fluid in the jet compared with the ambient flow. The dye streaks outside the jet reveal the re-entrainment process of the flow owing to the recirculation in the domain. As Jirka & Harleman (1979) noted, the re- 167 6 Flow induced by a jet in a confined domain entrainment process leads to an increase in dye concentration in the jet. The boundary between the jet and the ambient fluid can be identified by the sharp contrast between the light intensity in the induced ambient flow and the light intensity in the turbulent jet flow. We find in figure 2.4 that the location of the average dye edge of quasi-two-dimensional jets is a linear function of height with a slope equal to 0.22 from the z-axis (calculated for 20 ≤ z/d ≤ 120), this corresponds to an average angle of approximately 12◦ from the z-axis. Kotsovinos (1978) also defined the jet boundary as the separation between the turbulent jet flow and the ambient flow field. He too used the sharp contrast between dyed jets and the ambient fluid to determine the boundaries of plane turbulent jets. In figure 2.14, we can observe that the location of the jet boundary is almost at the location x0(z), where the time-averaged streamwise velocity vanishes and changes sign. It means that, in average, the flow velocity is purely lateral at the jet boundary. Therefore, we choose x0(z) ≈ 0.22z as the location of the jet boundary in this study. We can also see that, at the jet boundary, the streamlines of the induced ambient flow are oriented in the opposite direction to the jet flow, thus producing a counterflow which can affect the momentum flux of the jet (Kotsovinos & Angelidis, 1991). 6.3.3 Quantitative results In figure 6.13, we compare the streamlines predicted theoretically in equations (6.38), (6.19) and (6.26) (plotted with solid curves) with the streamlines of an ensemble-averaged experimental flow field (plotted with dotted curves). The ex- perimental flow field is the ensemble average of the time-averaged flow fields of the three jets studied (see § 6.3.1). Both the experimental and theoretical stream- lines start at the same locations along the top boundary of the PIV study area 5, i.e. at z/d = 85. As we can see, the theory and the data agree in the far field away from the jet axis. The streamlines of the experimental data, which point downwards in the ambient fluid, change direction at the boundary of the jet to point upwards. In figure 6.14(a), we compare the time-averaged lateral velocity predicted the- oretically in equations (6.40), (6.19) and (6.26) (plotted with solid curves) with the time-averaged lateral velocity (plotted with dotted curves) of the ensemble- averaged experimental flow field shown in figure 6.13. We show the experimen- 168 6.3 Experimental results 0 20 40 60 80 100 120 140 160 hf d z/ d 0 20 40 60 80 xj dx/d (a) (b) Figure 6.12: (a) Passive tracers (Pliolite particles) shown as streaks in a typical jet (Rej = 4030). The axis of the jet is plotted with a dot-dashed line at x = xj . The free surface is plotted with a black dashed line at z = hf . The transition between the jet region and the impingement region is shown with a red dashed line at z = hi. (b) Grey-scale picture of a turbulent quasi-two-dimensional dyed jet (Rej ≈ 4000) and the induced flow in the region delimited by −40 ≤ (x− xj)/d ≤ 40 and 0 ≤ z/d ≤ 100. tal and theoretical lateral distributions of the lateral velocity at four different heights (indicated by dashed lines) in the PIV study area 5 (0 ≤ x/d ≤ 85, 0 ≤ z/d ≤ 85). We normalize the time-averaged lateral velocity u with the maxi- mum time-averaged streamwise velocity wmax measured in the domain (at height z/d ≈ 75, see figure 6.14b). The normalized velocity u/wmax is then scaled so that the maximum amplitude (i.e. wmax/wmax = 1) matches the z-separation between two heights of measurement (shown with dashed lines in figures 6.14a). Similarly to the streamlines, the theory and the data agree in the far field away from the jet boundary, except for the lowest curve at z/d = 10. We believe that the difference at z/d = 10 is due to the fact that the jet is very close to the source, and thus the flow of the quasi-two-dimensional jet (as well as its induced flow) is not yet established (see § 2.4). 169 6 Flow induced by a jet in a confined domain 0 10 20 30 40 50 60 70 80 x/d 0 20 40 60 80 100 z/ d Data Theory Figure 6.13: Experimental (dotted curves) and theoretical (solid curves) distributions of the time-averaged streamlines of the flow induced by quasi-two-dimensional jets in the PIV study area 5, 0 ≤ x/d ≤ 85, 0 ≤ z/d ≤ 85 (see equations (6.38), (6.19) and (6.26) for the theoretical curve). In figure 6.14(b), we compare the time-averaged streamwise velocity predicted theoretically in equations (6.41), (6.19) and (6.26) (plotted with solid curves) with the time-averaged streamwise velocity (plotted with dotted curves) of the ensemble-averaged experimental flow field shown in figure 6.13. We show the ex- perimental and theoretical lateral distributions of the lateral velocity at four dif- ferent heights (indicated by dashed lines) in the PIV study area 5 (0 ≤ x/d ≤ 85, 0 ≤ z/d ≤ 85). Similarly to the time-averaged lateral velocity, we normal- ize the time-averaged streamwise velocity w with the maximum time-averaged streamwise velocity wmax measured in the domain (at height z/d ≈ 75). The normalized velocity w/wmax is then scaled so that the maximum amplitude (i.e. wmax/wmax = 1) matches the z-separation between two heights of measurement (shown with dashed lines in figures 6.14b). The theory seems to agree with the data in the far field away from the jet boundary at least to leading order. How- ever, we can see that the experimental time-averaged streamwise velocity is not exactly uniform across the study area. In particular, the experimental streamwise velocity vanishes along the right-hand-side boundary at x = 0, contrary to the theoretical streamwise velocity which assumes a slip-boundary condition. In figure 6.15(a), we plot the non-dimensional time-averaged volume fluxQr/Q0 170 6.3 Experimental results 0 1 20 30 40 50 60 70 80 x/d 0 20 40 60 80 100 z/ d Data Theory (a) 0 10 20 30 40 50 60 70 80 x/d 0 20 40 60 80 100 z/ d Data Theory (b) Figure 6.14: Experimental (dotted curves) and theoretical (solid curves) lateral distri- butions of the time-averaged velocity field of the flow induced by quasi-two-dimensional jets in the PIV study area 5 (0 ≤ x/d ≤ 85, 0 ≤ z/d ≤ 85) at four different heights (plot- ted with dashed lines) for: (a) the normalized time-averaged lateral velocity u/wmax (see equations (6.40), (6.19) and (6.26) for the theoretical curve), with wmax the max- imum time-averaged streamwise velocity measured in the domain (at height z/d ≈ 75, see b); (b) the normalized time-averaged streamwise velocity w/wmax (see equations (6.41), (6.19) and (6.26) for the theoretical curve). of the return flow (combining the two sides of the jet) versus non-dimensional height z/d. The experimental data, plotted with a dotted curve, are computed from the streamwise velocity of the ensemble-averaged experimental flow field shown in figure 6.14(b), such that Qr,exp(z) = −2 ∫ xj−x0(z) 0 w(x, z) dx, (6.42) where x0(z) ≈ 0.22z defines the location where w = 0, as discussed in § 6.3.2. The first theoretical prediction, plotted with a solid curve, is computed using the streamwise velocity predicted by potential theory and defined in equations (6.41), (6.19) and (6.26), such that Qr,pot(z) = −2 ∫ xj 0 w(x, z) dx. (6.43) Note the difference between the top boundaries of the integral (6.42), where we choose the boundary of the jet, and the integral (6.43), where we choose the jet axis (because the jet is modelled as a line sink). Using equation (6.41), we find that the volume flux Qr,pot increases linearly with distance z. However, equation 171 6 Flow induced by a jet in a confined domain (6.43) is valid only in the region where the jet is self-similar and the momentum flux of the jet is conserved, i.e. for z ≤ hi. The second theoretical prediction, plotted with a dashed curve, is computed using conservation of volume flux at every height. By continuity, the downward volume flux of the return flow on both sides of the jet Qr,cont is equal to the upward volume flux of the jet Q (defined in (2.4b)) minus the source volume flux Q0, at every height: Qr,cont(z) = Q(z)−Q0. We find Qr,cont(z) = Q0 ( 4 √ 2αM0zQ02 + 1 )1/2 −Q0. (6.44) The volume flux Qr,cont increases like z1/2 with distance. However, similarly to Qr,pot, Qr,cont can only model the return flow for z ≤ hi. The theoretical predic- tions Qr,pot and Qr,cont have different growth rates because of the difference in the boundary conditions. Qr,pot is computed with fixed boundary conditions (the jet is modelled as a line sink located at x = xj), whereas the boundary conditions for Qr,cont change linearly with distance (the jet boundary is a function of height, x0(z) ∝ z). We can see in figure 6.15(a) that the experimental return-flow volume flux is increasing between the two theoretical curves. The first prediction (6.43), based on potential theory, underestimates the volume flux, whereas the second prediction (6.44), based on continuity, overestimates the volume flux. We can also note that the volume flux of the return flow becomes rapidly larger than the initial volume flux, Qr/Q0 ≈ 3 at mid-height in the tank z = hf/2. Therefore, the volume flux of the return flow is of the order of magnitude of the jet volume flux away from the source, i.e. Qr ≈ Q for z ≥ hf/2. As predicted, Qr,pot increases linearly while Qr,cont increases like z1/2 with distance. We believe that the assumption we make to model the jet as a fixed line sink is valid only if the distance between the jet axis and the lateral wall (at x = 0) is large. The trend of the experimental data Qr,exp is not accurate enough in figure 6.15(a) to distinguish a linear growth rate, as predicted by Qr,pot, or a growth rate of z1/2, as predicted by Qr,cont. Similarly to the volume flux, we plot in figure 6.15(b) the non-dimensional time-averaged momentum flux M r/ ( Q02/d ) of the return flow (combining the two sides of the jet) versus non-dimensional height z/d. The experimental data, plotted with a dotted curve, are computed from the time-averaged streamwise 172 6.3 Experimental results velocity of the ensemble-averaged experimental flow field shown in figure 6.14(b), such that M r,exp(z) = 2 ∫ xj−x0(z) 0 (w)2(x, z) dx. (6.45) The first theoretical prediction, plotted with a solid curve, is computed using the streamwise velocity predicted by potential theory and defined in equations (6.41), (6.19) and (6.26), such that M r,pot(z) = 2 ∫ xj 0 (w)2(x, z) dx. (6.46) Using equation (6.41), we find that the momentum flux M r,pot increases like z2 with distance. However, as we mentioned for the volume flux, equation (6.47) is valid only in the region where the jet is self-similar and the momentum flux of the jet is conserved, i.e. for z ≤ hi. The second theoretical prediction, plotted with a dashed curve, is computed using the volume flux Qr,cont defined in (6.44) and assuming a uniform velocity outside the jet (with slip boundary condition at the walls). We find M r,cont(z) = ( Qr,cont )2 (z) 2 (xj − x0(z)) . (6.47) The momentum flux M r,cont increases like z2 with distance for x0(z)/xj ≪ 1. But, unlike M r,pot, we find that in the limit x0(z)/xj → 1, M r,cont increases like 1/(1 − x0(z)/xj) with distance z. Nevertheless, and similarly to M r,pot, M r,cont can only model the return flow for z ≤ hi, so that x0(z)/xj < 1 and M r,cont only increases like z2. This discrepancy between the asymptotic behaviours of the two theoretical predictions M r,pot and M r,cont is, again, due to the difference between the boundary conditions. Similarly to the volume flux, we can see in figure 6.15(b) that the experimental return-flow momentum flux is increasing between the two theoretical curves. The first prediction (6.46), based on potential theory, underestimates the momentum flux, whereas the second prediction (6.47), based on continuity, overestimates the momentum flux. As predicted, we can observe that both M r,pot and M r,cont increases like z2, which seems to be also the case for the experimental momentum flux M r,exp. From consideration of figure 2.6, we find that the time-averaged momentum flux of the jet is approximately constant with height at an average 173 6 Flow induced by a jet in a confined domain 0 10 20 30 40 50 60 70 80 z/d 0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 Q r/ Q 0 Data: Qr,exp Theory: Qr,pot Theory: Qr,cont (a) 0 10 20 30 40 50 60 70 80 z/d 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 M r/ ( Q 0 2 /d ) Data: Mr,expTheory: Mr,pot Theory: Mr,cont (b) Figure 6.15: Experimental and theoretical distributions against non-dimensional height z/d of: (a) the experimental normalized time-averaged volume flux Qr,exp/Q0 (dotted curve) computed using (6.42), the theoretical prediction based on potential the- ory Qr,pot/Q0 (solid curve) and computed using (6.43), the theoretical prediction based on continuity Qr,pot/Q0 (dashed curve) and computed using (6.44); (b) the experi- mental normalized time-averaged momentum flux M r,exp/ ( Q02/d ) (dotted curve) com- puted using (6.45), the theoretical prediction based on potential theoryM r,pot/ ( Q02/d ) (solid curve) and computed using (6.46), the theoretical prediction based on continuity M r,cont/ ( Q02/d ) (dashed curve) and computed using (6.47). value of < M > / ( Q02/d ) = 0.55. The non-dimensional momentum flux of the return flow increases from M r/ ( Q02/d ) = 0 at z/d = 0 to approximately 0.06 at z/d = 80. Therefore, in our domain, the momentum flux of the return flow is rather insignificant compared with the jet momentum flux. This finding is completely different from the results for the volume flux of the return flow, which is not insignificant because it has to balance the volume flux of the flow. This crucial difference, which enables us to neglect the influence of the momentum flux of the return flow on the jet flow, is related to the distance between the jet and the lateral boundaries (i.e. to the lateral confinement of the jet) and justifies the assumptions we make in Chapter 2 that the return flow has a weak effect on the dynamics of the evolving jet. 6.4 Conclusion In this chapter, we study the flow induced by a steady quasi-two-dimensional turbulent jet in a confined rectangular domain. Using two-dimensional potential theory, we determine the induced flow in a representative domain Ds of aspect 174 6.4 Conclusion ratio ζ = xj/hi = 3/4 (corresponding to our particular case) between the stream- wise dimension and the cross-stream dimension. The domain is delimited by the jet axis at the right-hand-side boundary (at x = xj), the walls of the experimental apparatus at the left-hand-side and bottom boundaries and the transition height hi between the jet flow region and the impingement region, at the top boundary. The jet is modelled as a line sink (located on the jet axis) with a lateral flux per unit length varying with height in a similar way to the entrainment velocity due to a quasi-two-dimensional jet. The transition height hi is modelled as a uniform line source, whose total inwards flux matches the total outwards flux of the line sink. To solve Laplace’s equation in the domain Ds, we decompose the problem into a uniform problem with a uniform line source and a uniform line sink, and a perturbation problem accounting for the varying line sink condition at the jet boundary. We find an analytical solution for the potential field, the stream function and the velocity field in the domain Ds. It appears that in the far field, away from the jet, the results are dominated by the uniform problem with uniform boundary conditions. The influence of the varying line sink (i.e. the entrainment process of the jet) is strong near the source of the jet, because of the singularity at the virtual origin of the jet (located outside the domain below the bottom boundary). We observe qualitative discrepancies between our analytical solution for the streamlines of the induced flow compared with the solutions of Taylor (1958) or Schneider (1981). The second derivative of the streamlines with respect to the lateral or cross-jet coordinate (x) have a different sign. Our streamlines are convex, whereas the streamlines of Taylor (1958) or Schneider (1981) are concave. This difference is due to the fact that we consider a fully confined domain, which induces a recirculation flow on either side of the jet, whereas Taylor (1958) or Schneider (1981) considered fully unbounded domains or the case of a jet emerging from a wall into a semi-infinite domain, thus ignoring the possibility of recirculation in the ambient flow. We compare our theoretical flow field with experimental data from quasi-two- dimensional turbulent jets in a confined experimental apparatus of aspect ratio 1 (the ratio between the inner dimensions of the tank). The theoretical streamlines agree with the data in the far-field, away from the boundary of the jet. We find 175 6 Flow induced by a jet in a confined domain that the boundary of the jet, defined as the boundary between the turbulent jet flow and the ambient flow (Kotsovinos, 1978), also corresponds to the location x0(z) ≈ 0.22z where the flow is, in average, purely lateral because the time- averaged streamwise velocity vanishes and changes sign at x0(z). In our model, we assume that the jet boundary coincides exactly with the jet axis, instead of being at an angle of approximately 12◦. We find that this assumption is valid in the far-field away from the jet boundary and for z ≤ hi. We find that, to the leading order, the experimental velocity field agrees with the model. Differences are seen near the rigid boundaries, where the experimental time-averaged tangential velocity vanishes at the walls, contrary to the theoretical tangential velocity which is assumed to satisfy a slip boundary condition. Also, near the jet source, the experimental data differ from the model because the flow of the jet is not yet fully established. Finally, the experimental measurements for the volume flux and the momentum flux of the return flow agree to leading order with the model based on potential theory and a model based on volume conservation. In particular, we find that the time-averaged momentum flux of the return flow increases like z2 to approximately 10 % of the jet momentum flux at mid-height in the experimental apparatus. We believe that a jet emerging from a wall into a fully confined domain is a more realistic case than the case of a jet in an unbounded or semi-infinite domain. The streamlines of the induced flow are strongly modified by the recirculation cells observed on either side of the jet. This phenomenon is important in mixing problems because the re-entrainment process tends to increase the concentration in the jet of passive tracers injected in the fluid. The momentum flux of the jet can also become negatively affected by the counter-flow after a certain distance. The core and eddy structures also become affected by the confinement at a distance hi approximately equal to 65 % of the depth of the flow, for an experimental apparatus of aspect ratio 1 (i.e. the ratio between the inner dimensions of the tank) or a jet aspect ratio ζ = xj/hi = 3/4 (i.e. the ratio of the distance between the jet and the lateral boundary to the transition height of the impingement region). We believe that our model, and in particular the assumption of a uniform flux at the top boundary and the assumption of a jet boundary parallel to the z-axis on the right-hand side of the domain, is valid for a range of jet aspect ratios 2/3 < ζ = xj/hi < 3. At higher aspect ratios, secondary recirculation cells could 176 6.4 Conclusion form on either side of the jet (Jirka & Harleman, 1979), thus affecting the flux at the top boundary. On the other hand, at lower aspect ratios, the expansion of the jet boundary becomes significant compared with the size of the domain, and thus can influence the flux at the top boundary. 177 Chapter 7 Dynamics of particle-laden jets in quasi-two-dimensional environments 7.1 Introduction Two-phase flows involving mixtures of solid particles and liquids are common in industrial applications. One example can be found in the coking process of the residue, or heavy-tar, from the refinement of crude oil, which serves to produce graphite electrodes for smelting applications (see e.g. Lee et al., 1997). During the coking, chemical reactions normally provoke a gradual phase transition of the heavy-tar into solid sponge coke; however, this process can also occasionally lead to the formation of a less valuable product called shot coke (Eser et al., 1986). Eser et al. (1986) reproduced the chemical reactions that can cause the production of shot coke, but much less is known about the dynamics of the flow and its mixing properties, when the heavy-tar is injected into the reactor. Another important 179 7 Dynamics of particle-laden jets in quasi-two-dimensional environments industrial application of particle-laden jets is for fluidized beds, which are used in chemical reactors or in the transport of granular material (see e.g. Zoueshtiagh & Merlen, 2007). The study of particle-laden jets is also relevant to geophysical applications such as volcanic eruptions (see e.g. Sparks, 1986; Ernst et al., 1996; Veitch & Woods, 2000; Walters et al., 2006), and the transport and resuspension of sediments by jets (see e.g. Neves & Fernando, 1995; Colomer & Fernando, 1996; Colomer, Casamitjana & Fernando, 1998; Cardoso & Zarrebini, 2002; Jiang, Law & Cheng, 2005). We have conducted different experiments in which a vertical water jet is dis- charged below a flat bed of particles immersed in water. Similarly to the ex- periments described in § 2.2 and § 4.1, these experiments are performed in a quasi-two-dimensional environment. The jet and the bed are constrained be- tween two close walls in the spanwise (y-) direction. Typically, the dimension of the gap is two orders of magnitude smaller than the other dimensions. This particular geometry allows us to visualize and study the evolution of the system inside the bed of particles. Rich dynamical behaviours, characteristically differ- ent from the three-dimensional case, appear in two dimensions. For instance, the quasi-two-dimensional particle-laden jet (Q2DPL jet) presents an unreported instability occurring at intermediate flow rates. The objective is to understand and analyze the succession of regimes shown by the evolution of the system while the jet flow rate is changed. We are interested in the interaction between the jet and the bed of particles. The entrainment and recirculation of the particles in the jet reveal an interesting coupling with the ge- ometry of the eroded bed. The maximum height reached by the particles is a key parameter to understand the dynamics of the whole system. As a model of the Q2DPL jet, we draw a comparison with the non-buoyant quasi-two-dimensional momentum jet studied in the previous chapters. We discuss the assumptions and conditions under which the model holds on the basis of experimental and theo- retical results. Future work will be to compare Q2DPL jets with heavy fountains in a quasi-two-dimensional environment, in order to model the regimes where the density difference between the Q2DPL jet and the ambient fluid is important. The rest of this chapter is organized as follows. In § 7.2, we describe the experimental procedure. In § 7.3, we describe the different phenomenological regimes observed in the experiment, as the jet flow rate increases. In § 7.4, we 180 7.2 Experimental procedure discuss a model to predict the maximum height reached by the particles in the final dilute regime, based on the model for the time-averaged mean momentum jet (presented in § 2.4). We draw our conclusions in § 7.5 and suggest new avenues of research. 7.2 Experimental procedure The 0.5 m (L) × 0.01 m (W ) × 0.5 m (H) quasi-two-dimensional experimental apparatus is presented in figure 7.1.1 The (heavy) particles we use are 0.5 mm glass beads (density: ρp = 2.5 g cm−3 ). The initial thickness of the bed ranges from h0 = 1.75 to 8 cm. At the beginning of an experiment, we lay the bed flat at the bottom of the tank, which is filled with water. The water jet, injected through a circular nozzle of diameter d = 6 mm located at the middle of the bottom of the tank, has a source flow rate ranging from Q0 = 0 to 33 cm3 s−1 . We generate the flow either by gravity or using a peristaltic pump (the pulsing of the pump had no influence on the system for Q0 > 4 cm3 s−1 ). We increase the flow rate in a stepwise manner, with typical step 1 cm3 s−1 . After each increase in the flow rate, we allow the system to reach steady state (characterized by a fixed shape of the bed and by an approximately constant amount of particles in circulation in the jet, i.e. particles no longer sediment in the far field). This typically takes 5 to 30 minutes. At steady state, we take (with a ruler) some characteristic geometric measurements of the shape of the bed: the size of the cone formed by the erosion of the bed, the thickness of the bed above the nozzle (hsource) and the angle of the slopes of the cone (with accuracy of approximately ±5 mm for the lengths and ±5◦ for the angle). Moreover, we measure the maximum height reached by the particles in the jet hmax (accuracy of ±5 to ±20mm) and the oscillation frequency of the Q2DPL jet (accuracy of ±15 %). 1Note that the length L and the height H of the experimental apparatus used in this chapter are half the length and half the height of the experimental apparatus used in Chapters 2 to 6. 181 7 Dynamics of particle-laden jets in quasi-two-dimensional environments x z d = 0.6 cm Liquid jet overflow H = 50 cm L = 50 cm W = 1 cm Bed of particles h0 Liquid medium Figure 7.1: Sketch of the experimental apparatus. The evolution of the system is analyzed when a water jet of variable flow rate is injected at the bottom of the bed of particles (0.5 mm glass beads). 7.3 Phenomenological description 7.3.1 Regime diagram In figure 7.2, we present a schematic diagram of the successive regimes displayed by the system (bed of particles and jet) as the jet flow rate Q0 increases, and for various initial bed thicknesses h0. We start with Q0 = 0 and a flat bed of particles. At very low flow rates, the bed remains motionless because the pressure of the flow is insignificant compared with the weight of the bed of particles per unit area. In this pre-regime (not represented in figure 7.2), we have a porous medium flow, which can be modelled using Darcy’s law (Zoueshtiagh & Merlen, 2007) W = −κµ∇P, (7.1) where W is the superficial velocity, κ is the permeability of the bed, µ is the dynamic viscosity of the liquid and P the pressure. From this model, Zoueshtiagh 182 7.3 Phenomenological description Source volume flux Q0 h0 In it ia l be d th ic kn es s I Fluidized bed II Oscillatory flow III Core and eddy flow Figure 7.2: Schematic diagram showing the boundaries between the three different phenomenological regimes observed as the source flow rate is increased from Q0 = 0 to 33 cm3 s−1 . & Merlen (2007) deduced the flow velocity at the surface of the bed. Their experimental measurements of the flow velocity at the bed surface agreed with the theoretical prediction as long as the configuration could be considered as a point source in an infinite domain (i.e. d/h0 ≤ 0.2). As we increase the source flow rate (for a given bed thickness), the system displays very different regimes, which we describe in detail below. At low flow rates, we observe a fluidization regime after the Darcy flow regime (see regime I in figure 7.2). Then, the jet starts eroding the bed if we increase Q0 further. In this “oscillatory flow” regime (see regime II in figure 7.2), the Q2DPL jet is unstable and oscillates in the eroded bed with respect to the vertical axis. Finally, at large Q0 the bed is fully eroded and the jet can lift particles higher in the water above the bed. The motion of the particles in the jet flow shows the same core and eddy structures as we observe in particle-free quasi-two-dimensional jets. So, in reference to the flow in particle-free quasi-two-dimensional jets, we name this regime the “core and eddy flow” regime (see regime III in figure 7.2). 183 7 Dynamics of particle-laden jets in quasi-two-dimensional environments 7.3.2 Regime I: fluidized bed At low flow rates, the jet fluidizes the bed, which maintains a flat surface (see figure 7.3). We observe a strong recirculation of the particles inside a chimney, or narrow cone, located above the nozzle (see the dashed blue lines in regime I, fluidized bed, in figure 7.2). In fluidization models, the basic hypothesis states that the pressure gradient of the flow inside the chimney is balanced by the weight of the particles in the chimney (see e.g. Zoueshtiagh & Merlen, 2007), i.e. ∆p = φb∆ρgh0, (7.2) where ∆p = (p(0) − p(h0)) is the pressure difference between the bottom of the bed (at z = 0) and the top surface of the bed (at z = h0), φb is the volume fraction of the bed,2 ∆ρ = ρp−ρ is the difference between the density of the particles and the density of the liquid, and g is the constant of gravity. In the rest of the bed, the flow follows Darcy’s law (described in (7.1)). Therefore, the total flow rate at the nozzle exit is equal to the sum of the flow rate in the fluidized chimney and the flow rate in the unfluidized part of the bed. Zoueshtiagh & Merlen (2007) conducted a similar experiment in three dimen- sions. It is interesting to note that our observations about the fluidization of the bed in the quasi-two-dimensional case agree qualitatively with their report. They also proposed a model for the fluidization process and the formation of the chimney. However, they could not obtain any experimental evidence to validate or invalidate the model because their experimental apparatus did not allow them to visualize and measure the interior of the bed where the fluidization process occurred. We believe that our quasi-two-dimensional experiment, which gives a clear picture of the dynamics inside the bed, could provide quantitative data to verify the fluidization model of Zoueshtiagh & Merlen (2007). 7.3.3 Regime II: oscillatory flow As we increase the flow rate, the opening angle of the cone increases by an erosion process, as depicted in figure 7.4. At an intermediate range of flow rates the 2Ojha, Menon & Durian (2000) determined experimentally the volume fraction of a slowly defluidized bed of particles. They found that the volume fraction φb is independent of the size or shape of the particles and is approximately equal to φb = 0.59. 184 7.3 Phenomenological description Figure 7.3: Illustration of regime I, Q0 = 2.5 cm3 s−1 and the initial height h0 = 4cm. At very low flow rates the jet coming through the bed at the bottom center of the picture fluidizes the particles above it. The surface of the bed remains almost flat (only a small hump is observed), showing almost no activity. Q2DPL jet builds a mound of particles on each side of the cone. For every increase in the flow rate, the mounds grow and move away from the cone until they reach a steady state (i.e. a fixed position). In addition, we can observe a surprising behaviour in this regime: the Q2DPL jet does not maintain a vertical axis, but oscillates in the (x, z) plane about the z-axis (the origin of the domain (0, 0, 0) is at the centre of the nozzle). As we can see in figure 7.5, a large vortical structure develops alternately on each side of the z-axis at each semi-oscillation. As an explanation of this oscillating phenomenon, we believe that the Q2DPL jet becomes unstable as it emerges out of the bed and is no longer bounded by the steep walls of the cone. An oscillation can start when some disturbance breaks the symmetry of the sedimentation process. This leads to an asymmetry of the avalanching process occurring on the slopes of the cone. The stronger avalanche deflects the jet to the opposite side of the cone. This has the effect of reversing the asymmetric avalanching process, which can then deflect the jet back to the initial side of the oscillation, thus completing a whole cycle. The oscillation of the jet is sustained by the kinetic energy of the jet. The oscillation frequency of the jet appears to be steady for a given source flow rate and to decrease with increasing source flow rate. 185 7 Dynamics of particle-laden jets in quasi-two-dimensional environments Figure 7.4: Illustration of regime II, the jet flow rate is Q0 = 15 cm3 s−1 and the frequency of image acquisition is 500 frames per second. At medium-high flow rates, a mound of particles forms on each side of the cone. The opening angle of the cone increases rapidly with the flow rate. The jet oscillates from side to side in the crater. A steady state is reached when the jet cannot eject any particles above the mounds. Figure 7.5: Illustration of the vortical structure in regime II, Q0 = 18.5 cm3 s−1 and h0 = 4 cm. The Q2DPL jet displays a large vortex as it oscillates alternately on each side of the z-axis (several images are superimposed on this picture to show particles as streaks). 186 7.4 Core and eddy flow model 7.3.4 Regime III: core and eddy flow As we can see in figure 7.6(a), the final regime of the system is characterized by a fixed final shape of the bed. The bed forms a crater, whose slopes are at the angle of repose (approximately 26◦ for our particles). Moreover, it is flanked by two pyramidal mounds, whose added volumes (computed above the initial height of the bed, i.e. for z > h0) account for the volume of the crater. In this regime, the jet still entrains some particles, but the volume fraction is much lower. Unlike in regime II, the flow in the jet does not seem to be very strongly affected by the particles in this regime. We observe a sharp decrease in the frequency and amplitude of the oscillations of the Q2DPL jet, and also an increase in the rate of change of the particle maximum height hmax (plotted with triangles in figure 7.7) with the source flow rate Q0. The trajectory of the particles in the jet flow (shown as streaks in figure 7.6a, where the velocity is large) reveals three large eddies (identified with yellow cir- cular arrows) and a high speed core (identified with a yellow arrowed curve). The resemblance to the core and eddy structures in a quasi-two-dimensional turbulent jet (depicted with yellow circular arrows and a yellow arrowed curve, respectively, in the picture of a dyed quasi-two-dimensional jet in figure 7.6b) is strong. From our qualitative observations, we can report that the size of the eddies in the Q2DPL jet also grows with distance z. Therefore, it appears that in the case of a dilute concentration of particles, the momentum of the jet is not strongly affected by the negative bulk density of the two-phase flow. 7.4 Core and eddy flow model The experiment described above reveals the complexity of the interaction between the jet and the bed of particles. In this section, we are interested in the final regime, or core and eddy flow regime, which has strong similarities with the flow in a quasi-two-dimensional jet, described in the previous chapters. Assuming that the density of the particles does not affect the flow of the jet we propose a model for the maximum height reached by the particles transported by the jet. From the study of non-buoyant quasi-two-dimensional momentum jets pre- sented in Chapter 2, we know the velocity field of a quasi-two-dimensional jet in our apparatus. If we assume that, in the final regime (i.e. regime III), the con- 187 7 Dynamics of particle-laden jets in quasi-two-dimensional environments (a) (b) Figure 7.6: (a) Illustration of the core and eddy structures in regime III, Q0 = 23 cm3 s−1 and h0 = 4 cm. At very high flow rates, the slopes of the cone are at the angle of repose and the bed has a fixed shape. The Q2DPL jet displays three large growing eddies (identified with yellow circular arrows), which are advected upwards by the flow, as well as a high speed core (identified with a yellow arrowed curve). In this picture, the maximum height reached by the particles is approximately 20 cm. (b) Grey-scale picture of a dyed steady quasi-two-dimensional turbulent jet rising in the tank over a height of approximately 40 cm (this dyed jet was produced in the appara- tus depicted in figure 7.1, following the experimental procedure described in § 2.2.1). Similarly to (a), the eddies are identified with yellow circular arrows and the core is identified with a yellow arrowed curve. centration of the particles in the Q2DPL jets does not strongly affect the liquid phase of the jet, the particles are passively advected by the flow. Therefore, we can theoretically compute the maximum height reached by the particles htmax. We make the simplifying assumption that the theoretical particle maximum height htmax is equal to the height at which the maximum vertical velocity of the pure momentum jet matches the particle settling velocity. For our particles, the set- tling velocity is vs ≈ 7.2±0.4 cm s−1 in the tank at rest. We further assume that the maximum vertical velocity of the jet is approximately equal to its maximum time-averaged vertical velocity wm, described by (2.5b). Solving wm(z = htmax) = vs, (7.3) 188 7.4 Core and eddy flow model we find htmax = Q02 4 √ 2α ( 2 ( M0 vsQ0 )2 − 1 ) , (7.4) where α = 0.068 is the entrainment coefficient (Morton et al., 1956) measured in § 2.4, and M0 is the source momentum flux. As discussed in § 2.4, we have the relationship M0/ ( Q02/d ) ≈ 0.55. In figure 7.7, we present the experimental results for the evolution of the non- dimensional particle maximum height hmax/d against the non-dimensional source flow rate Q0/(vsd). The experimental data are obtained for an initial bed thick- ness h0 = 1.85 cm. The data plotted with triangles for hmax are obtained with increasing Q0 (in the stepwise manner described in § 7.2). The data plotted with pluses for hmax are obtained with decreasing Q0. We conduct this second phase directly after the Q0-increasing phase, after having reached the maximum source flow rate (which can lift the particles to the free surface) and the experiment has reached a steady state (as described in § 7.2). We also plot in figure 7.7 the thickness of the bed of particles above the nozzle hsource (in the Q0-increasing phase of the experiment) with blue squares (multi- plied by a factor 15, for clarity, and non-dimensionalized by d). The evolution of hsource with Q0 indicates the transitions between the three regimes I, II and III described in § 7.3. From Q0/(vsd) ≈ 0 to 0.8, hsource remains constant and the bed is fluidized by the jet (regime I, on the left-hand side of the first dotted line plotted in figure 7.7). In regime II (within the two dotted lines), or from Q0/(vsd) ≈ 0.8 to 3.6, hsource decreases because of the erosion process above the jet nozzle. In regime II, we can also notice that the Q0-increasing data for hmax are increasing slowly, i.e. the particles rise only slightly higher than the initial bed thickness. For Q0 > 3.6, or regime III (on the right-hand side of the second dotted line in figure 7.7), the erosion process is finished. In regime III, the Q0- increasing data for hmax are in a new regime: the rate of increase of hmax with Q0 is larger than in regime II. Moreover, we show in figure 7.7 the impingement transition height hi = 3/4H (plotted with a black dashed line), which we discuss in Chapter 6. The transport of the particles by the jet is perturbed beyond this height because the flow changes from a jet flow to an impingement flow. We can see that, for z > hi, hmax no longer shows the same increasing trend, but approximately plateaus. 189 7 Dynamics of particle-laden jets in quasi-two-dimensional environments We plot in figure 7.7 the range of the non-dimensional theoretical prediction htmax/d ± 12 % (computed using (7.4)) between two red dashed curves. We al- low a variation of ±12 % in the calculation of htmax to account for the typical 0.4/7.2 = 5.5 % standard deviation in the measurements of the particle settling velocity vs. As we can see, the theoretical curves lie above the Q0-increasing data for hmax (plotted with triangles) and slightly below the Q0-decreasing data for hmax (plotted with pluses). Moreover, we can observe a strong hysteresis be- tween the two data sets: the particles rise lower in the Q0-increasing phase than in the Q0-decreasing phase of the experiment. We believe that the main reason for this hysteresis is because during the Q0-increasing phase the bulk density of the Q2DPL jet is larger than during the Q0-decreasing phase. During the Q0- increasing phase, the Q2DPL jet loses particles because, at each increase of the source flow rate, particles in the jet can settle outside the cone of recirculation until the steady state is reached. On the other hand, once the system has reached a steady state at the maximum flow rate, no more particles can settle out of the recirculation cone as the source flow rate is reduced. Therefore, the assumption that the particle concentration does not affect the jet flow appears to be incor- rect in the Q0-increasing phase of the experiment. In the Q0-decreasing phase, the ‘dilute’ assumption seems to be valid because the theoretical prediction un- derestimates the experimental data only slightly. This small mismatch could be related to the (second) assumption that the time-dependent vertical velocity is approximately equal to the time-averaged vertical velocity. Indeed, we find in Chapter 2 that the vertical velocity in the high-speed core of the jet is different from the Gaussian profile of the time-averaged vertical velocity, and we are un- able to investigate whether the maximum height hmax reached by the particles is attained exclusively with flow in the high-speed core. 7.5 Conclusion 7.5.1 Summary We have studied the dynamics of quasi-two-dimensional particle-laden jets in the case of a vertical jet injected below a bed of particles confined in a quasi- two-dimensional environment. We have observed several regimes as we increase the source flow rate and vary the initial bed thickness. Initially, we find the well- 190 7.5 Conclusion 0 1 2 3 4 5 6 7 Q0/(vsd) 0 10 20 30 40 50 60 70 80 h /d Increasing Q0: hmax Decreasing Q0: hmax Data: 15hsource Theory: htmax ± 12% hi I II III Regime separation lines Figure 7.7: Evolution against the non-dimensional source flow rate Q0/(vsd) of: the non-dimensional experimental particle maximum height hmax/d for increasing source flow rate (black triangles), the non-dimensional experimental particle maximum height for decreasing source flow rate (black pluses), the non-dimensional theoretical particle maximum height plus or minus 12 percent htmax/d± 12% (red dashed curves), and the non-dimensional experimental bed thickness (measured above the nozzle) 15hsource/d (plotted with blue squares). We plot the non-dimensional impingement transition height hi/d with a black dashed line. The different regimes (I, II and III), presented in the regime diagram shown in figure 7.2 and discussed in § 7.3, are delimited with dotted lines. known Darcy flow regime and fluidization regime for low flow rates (or large initial bed thicknesses). Then, the bed of particles evolves towards a triangular shape because the jet erodes the bed of particles gradually. The jet entrains particles above the bed, which settle and avalanche on the slope of the triangular eroded bed. In this regime, we observe an instability characterized by the oscillation of the Q2DPL jet. Finally, at large flow rates the particles are transported higher by the jet and their bulk concentration in the jet decreases. We propose a model for the final regime, in which the flow of the Q2DPL jet displays the same characteristic core and eddy structure as the flow of the non- buoyant quasi-two-dimensional momentum jet. Assuming that the concentration 191 7 Dynamics of particle-laden jets in quasi-two-dimensional environments of the particles does not affect the flow of the jet and that the time-dependent vertical velocity is approximately equal to the time-averaged vertical velocity, we calculate the maximum height reached by the particles by equating their set- tling velocity with the maximum time-averaged vertical velocity of non-buoyant quasi-two-dimensional jets (as described in Chapter 2). The comparison with ex- perimental results shows that the order of magnitude and the trend (with source flow rate) of the particle maximum height are predicted by the model. However, in the phase where the source flow rate increases, the assumption of a dilute suspension of particles in the jet appears to be incorrect because the theoreti- cal prediction overestimates the maximum height reached by the particles. On the other hand, in the (hysteretic) phase where the source flow rate decreases, the dilute assumption appears to be correct because the theoretical prediction only slightly underestimates the maximum height reached by the particles. The small mismatch is thought to be due to the difference between the time-dependent vertical velocity in the high-speed core of the jet and the time-averaged vertical velocity, which we assume in the simplified model presented here. 7.5.2 Future work Study of heavy fountains The next step in the understanding of the different regimes is to model the Q2DPL jet as a heavy fountain in a quasi-two-dimensional environment. We believe that a heavy fountain can account for the non-dilute regimes of the flow: the oscillatory flow regime, or regime II, and the Q0-increasing phase of regime III. A heavy fountain is a vertical upward jet with negative buoyancy (see e.g. Baines, Turner & Campbell, 1990, for an introduction to the theory of heavy fountains). From dimensional analysis, its maximum height is related to its initial momentum flux M0 and its initial buoyancy flux B0 zmax = A M0 |B0|2/3 , (7.5) with M0 = 2b0w02, B0 = 2b0 (ρf − ρh) ρf gw0, (7.6) and where A is a constant of proportionality which can be determined experimen- 192 7.5 Conclusion tally (Bower et al., 2008), ρf is the density of the ambient fluid, ρh is the density of the heavy fountain, b0 is the initial half-width of the fountain, and w0 is the initial time-averaged vertical velocity of the fountain. Then, we can re-write the density of the fountain in terms of the densities of the fluid and the particles, and the initial volume fraction of the particles φ0 in the fountain: ρh = φ0 ρp + (1− φ0)ρf . (7.7) Therefore, the particle maximum height is zmax = A ( φ0 (ρf − ρp) ρf g )−2/3 (2b0)1/3w04/3. (7.8) A future aim could be to apply this model to regime II and the Q0-increasing phase of regime III of the Q2DPL jet and to verify it experimentally. We could inject a homogenous heavy buoyant fluid in the quasi-two-dimensional tank in order to investigate the maximum height reached by the particles as a function of M0 and B0, and to measure the experimental constant A. However, relating the heavy-fountain model to the Q2DPL jet is particularly challenging because it requires an estimation of the initial buoyancy flux of the Q2DPL jet, and in particular the initial volume fraction φ0. The rate of entrainment of the particles at the source is one of the most critical and intriguing points: the recirculation of the particles denotes the coupling between the solid phase and the flow of the jet. Both the shape of the crater and the deposition of the particles determine the avalanching flux feeding the jet. The jet entrains these particles, which in turn affect the momentum of the jet by changing its bulk density. Thereafter, the particles rise to the height where their settling velocity matches the vertical velocity of the fluid. In order to close the recirculation problem, we must relate the sedimentation rate of the particles to both the settling time of the particles and their rising time inside the jet. In conclusion, a robust model accounting for the recirculation of the particles is needed to understand the full dynamics of the system. 193 7 Dynamics of particle-laden jets in quasi-two-dimensional environments Changing the viscosity of the medium To improve our model of the motivating industrial application (the problem of shot-coke formation in a late-stage oil-refining process), the viscosity of the medium in which the bed of particles is prepared could be different from the viscosity of the jet. This experiment is expected to have very rich dynamics be- cause instabilities such as finger-like structures can occur when viscous forces and gravity forces play a key role (Sto¨hr & Khalili, 2006). Study of the oscillation of the Q2DPL jet Another puzzling and interesting issue in this study is our discovery of an insta- bility at intermediate flow rates. The Q2DPL jet oscillates steadily at a fixed flow rate and about a vertical axis. It also produces a large vortex as it tilts sideways at every half oscillation. From these observations, we can wonder what sets the frequency of the oscillation and why the frequency decreases with Q0. There are different time scales that can influence the frequency: the rising time of the particles in the jet, the settling time of the particles and the avalanching time of the particles. Even the vortex recirculation time could be considered part of the problem; however, the vorticity tends to increase with the flow rate, which seems to be in contradiction with the fact that the oscillation frequency actually decreases with it. Finally, we also noticed that the frequency rapidly drops at the transition between regimes II and III, thus suggesting a different model for the evolution of the frequency in these two regimes. Injecting water and particles through the nozzle As a further step in the understanding of the experiments, we could change the experimental procedure by injecting both liquid and particles through the nozzle. The experimental results should be closer to real applications, such as industrial two-phase flows and volcanic eruptions. The complexity of this problem is likely to increase: for example, there cannot be a steady state at fixed flow rate because of the continuous injection of particles in the tank. We might find that the shape of the bed evolves in the reverse order from that which occurs in the present ex- periment as we increase the flow rate, i.e. passing through regime III, then regime II and finally regime I as shown on figure 7.2. From preliminary experiments, we 194 7.5 Conclusion observe the formation of an open and flatter crater from the sedimentation of the particles (Jiang et al., 2005; Neves & Fernando, 1995). Moreover, we find that the crater grows in size and its slopes become steeper, thus blocking the rise of the Q2DPL jet. The final stages could also show a fluidization regime and eventu- ally a porous medium flow. An interesting issue is to determine the mechanisms accounting for the transition from one regime to the next. Study of the three-dimensional case Three-dimensional particle-laden jets have been studied by many scientists (see e.g. Cardoso & Zarrebini, 2002; Colomer et al., 1998; Colomer & Fernando, 1996; Ernst et al., 1996; Jiang et al., 2005; Neves & Fernando, 1995; Walters et al., 2006; Zoueshtiagh & Merlen, 2007). However, our understanding of the quasi-two-di- mensional experiment gives us the opportunity to analyze the three-dimensional case from a different perspective. When comparing the two cases, it is possible to consider a wide variety of interesting issues: the dynamics of the particle-laden jet, the interaction between the particle-laden jet and the bed, the mechanisms of recirculation of the particles, their mixing properties, and the three-dimensional manifestation of the periodic oscillation observed in the quasi-two-dimensional environment. 195 Chapter 8 Conclusion and future work 8.1 Review In this thesis, we have studied experimentally and theoretically the dynamics of steady quasi-two-dimensional turbulent jets. In Chapter 1, we present a brief summary of past studies on quasi-two-dimensional jets, as well as some motiva- tions for this study. Giger et al. (1991) and Dracos et al. (1992) gave the first clear description of this particular type of jets, which occurs in the far field of a plane turbulent jet confined between two close boundaries separated by a gap widthW (i.e. for z > 10W , where z is the streamwise coordinate). They observed that, in the far field, the unstable flow develops into a meandering core with large counter-rotating eddies developing on alternate sides of the core. They found an inverse cascade of quasi-two-dimensional turbulence, which affects not only the structure of the flow but also the transport, dispersion and mixing properties. 197 8 Conclusion and future work One particular application relevant to this study concerns the flow of rivers dis- charging into lakes or oceans. Various phenomena are related to this type of flow: sediment transport, coastal erosion, and the transport and dispersion of passive tracers such as pollutants. Understanding the physics of the flow is crucial to the prediction and assessment of the environmental impact. In Chapter 2, we describe the phenomenology of the core and eddy structure of the jet using detailed experimental measurements of the velocity field, obtained with particle image velocimetry. We observe an inverse cascade typical of quasi- two-dimensional turbulence where both the core and the eddies grow linearly with z and travel at an average speed proportional to z−1/2. We find that quasi-two- dimensional jets are self-similar and their mean properties are consistent with both experimental results and theoretical models of the time-averaged properties of fully unconfined planar two-dimensional jets. The experimental results for the spatial statistical distribution of the core and eddy structure led us to believe that the dynamics of the interacting core and large eddies accounts for the Gaussian profile of the mean streamwise velocity. The lateral excursions (caused by the propagating eddies) of the high-speed central core produce a Gaussian distribution for the time-averaged streamwise velocity. In addition, we find that approximately 75% of the total momentum flux of the jet is contained within the core. The eddies travel substantially slower (at approximately 25 % of the maximum speed of the core) at each height and their growth is primarily attributed to entrainment of ambient fluid. The frequency of occurrence of the eddies decreases in a stepwise manner due to merging, with a well-defined minimum value of the corresponding Strouhal number St = fb/wm ≥ 0.07 (where f is the eddy frequency, b is the velocity spread rate of the jet and wm is the maximum time-averaged streamwise velocity in the jet). In Chapter 3, we investigate theoretically the streamwise transport and dis- persion properties of quasi-two-dimensional jets. We model the evolution in time and space of the concentration of passive tracers released in these jets using a one-dimensional time-dependent effective advection–diffusion equation. Based on the study of the flow field presented in Chapter 2, we make a mixing length hy- pothesis to model the streamwise turbulent eddy diffusivity Dzz ∝ bwm, where b is the jet velocity spread rate, wm is the maximum time-averaged streamwise velocity, and Dzz is the streamwise component of the turbulent eddy diffusive 198 8.1 Review tensor. Under these assumptions, the effective advection–diffusion equation for φ(z, t), the cross-stream integral of the ensemble-averaged concentration, is of the form: ∂tφ+KaM1/20 ∂z ( φ/z1/2 ) = KdM1/20 ∂z ( z1/2∂zφ ) , (8.1) where t is time, Ka (the advection parameter) and Kd (the dispersion parameter) are empirical dimensionless parameters which quantify the importance of advec- tion and dispersion, respectively, and M0 is the source momentum flux. We find analytical solutions to this equation for φ in the cases of a constant-flux release and an instantaneous finite-volume release. We also give an integral formulation for the more general case of a time-dependent release, which we solve analytically when tracers are released at a constant flux over a finite period of time. In Chapter 4, we compare the theoretical predictions of the streamwise ad- vection and dispersion model, derived in Chapter 3, with experimental evidence. From our experimental results, whose concentration distributions agree with the model, we find that Ka = 1.65±0.10 and Kd = 0.09±0.02, for both finite-volume releases and constant-flux releases using either dye or virtual passive tracers. The experiments also show that streamwise dispersion increases in time as t2/3. As a result, in the case of finite-volume releases, more than 50% of the total volume of tracers is transported ahead of the purely advective front (i.e. the front location of the tracer distribution if all dispersion mechanisms are ignored, corresponding formally to the assumption of ‘top-hat’ velocity profiles in the jet); and in the case of constant-flux releases, at each instant in time, approximately 10 % of the total volume of tracers is transported ahead of the advective front. Finally, we assess the statistical significance of our results. We find that experimental or real concentrations are more likely to differ from the concentrations predicted by the model at large concentration levels than at low concentration levels. These find- ings are important in problems of pollution control in rivers because they show that pollutants can travel faster than expected and their concentration may be higher than predicted. In Chapter 5, we investigate turbulent relative dispersion in quasi-two-dimen- sional turbulent jets. Following the seminal paper of Richardson (1926), we use two-point statistics to describe the dispersion properties of the core and eddy structure of the jet. The experimental data are obtained using what we believe to be a novel Lagrangian-particle-tracking technique, which we refer to as vir- 199 8 Conclusion and future work tual particle tracking. Virtual particle tracking, first introduced in Chapter 4, consists of tracking (numerically) virtual passive tracers seeded in the experimen- tally measured velocity field of a flow. We demonstrate that this technique can yield valuable experimental data to compare with turbulent relative dispersion models. We calculate the time evolution of the probability distributions of key two-point properties (such as the lateral distance, the streamwise distance, the Euclidean distance and the ratio of the lateral distance to the streamwise distance between two points) in three different parts of the flow of quasi-two-dimensional jets. We find that in the eddy, the distribution of particles disperses slowly and in a rather axisymmetric manner. At the interface between the core and the eddy, the distribution of particles stretches considerably in the streamwise direction at a high rate. In the core of the jet, the particle distribution disperses slowly in the cross-jet direction and splits along the jet axis. Finally, we believe that the rapid change in time of the jagged distribution of the p.d.f. for the distance between two points in the eddy reveals the intense stirring (and potentially the resulting vigorous turbulent mixing) occurring within the eddy. In Chapter 6, we use potential theory to describe the ambient flow induced by a quasi-two-dimensional jet discharged vertically upwards in a fully confined rectangular domain. In our experimental apparatus (of aspect ratio 1), we can observe that at a height hi ≈ 0.65hf (where hf is the distance of the free surface from the source) the jet flow becomes an impingement flow which spreads laterally along the free surface, recirculates downwards along the lateral boundaries of the apparatus, and is eventually re-entrained by the jet. In the domain, spanning from the lateral rigid boundary to the jet axis in the x-direction and from the bottom rigid boundary to the impingement transition height hi in the streamwise direction, we solve Laplace’s equation. We assume slip boundary conditions at the rigid boundaries, a sink link with varying strength proportional to (z−z0)−1/2 (where z0 is the space virtual origin of the jet) at the jet axis, and a uniform source line (whose integrated volume flux matches the integrated volume flux of the sink line) at the impingement transition height. The analytical stream function and velocity field agree with our experimental measurements, except near the boundary of the jet. We also find that (contrary to the volume flux) the time- averaged momentum flux of the induced return flow is insignificant compared with the time-averaged momentum flux of the jet, typically less than 10 %. We believe 200 8.2 Future work that this means that the induced return flow in our experimental apparatus has little impact on the flow structures of the quasi-two-dimensional jet studied in the previous chapters. In Chapter 7, we study the case of a momentum jet discharged below a bed of particles in a quasi-two-dimensional environment. As the jet flow rate increases, the interaction between the jet and the bed of particles evolves through three main different regimes. At low flow rates or large initial bed thicknesses, the jet fluidizes the bed. At intermediate flow rates, the jet erodes the bed and form a pyramidal mound on either side of the jet axis. The particle-laden jet is also unstable and oscillates about a vertical axis. At large flow-rates, the bed is fully eroded and the flow of the particle-laden jet shows the same core and eddy structure as the particle-free quasi-two-dimensional jet observed in Chapter 2. We propose a model to predict the maximum height of rise reached by the particles in the jet based on the time-averaged vertical velocity of a particle-free quasi- two-dimensional jet. We find that the model agrees with experimental data for a dilute suspension of particles in the jet. 8.2 Future work This study has raised questions for future research. We highlight below the vari- ous possible directions already discussed throughout the thesis. For instance, the streamwise advection and dispersion model (developed in Chapter 3) could be ex- tended to include advection and dispersion in the cross-jet direction of quasi-two- dimensional jets. With a two-dimensional time-dependent model, the distribution of the concentration of passive tracers in quasi-two-dimensional jets would be fully resolved. Such a model would provide more accurate predictions for dispersion and transport in river flows. A relationship between the two-point statistics in the jet and our streamwise advection and dispersion model (i.e. connecting Chapters 3 and 4 with Chap- ter 5) could improve our understanding of relative dispersion in turbulent flows. Furthermore, the spatial resolution of the results obtained with the technique of virtual particle tracking (described in § 4.1.2 and used in Chapters 4 and 5) could be enhanced to resolve the finest scale of turbulence. This would provide crucial experimental data for comparison with the vast number of turbulent dispersion 201 8 Conclusion and future work models. We also believe that the technique of virtual particle tracking, developed in this study, can be successfully applied to other flow problems. Virtual particle tracking can resolve Lagrangian particle tracking (as shown in Chapter 5), as well as identify Eulerian features in the flow (as performed in § 2.5 to study quantitatively the core and eddy structures of quasi-two-dimensional jets). The study of particle-laden jets in quasi-two-dimensional environments has also opened many avenues of research. An analogy with heavy fountains could give a basis to model the flow regimes with large concentrations of particles, and potentially explain the physics of the oscillatory instability displayed by quasi- two-dimensional particle-laden jets. Changing the viscosity of the liquid phase, or injecting solid particles with the liquid phase at the source would show very rich dynamics relevant to many industrial applications, such as coking, and geophysical applications, such as volcanic eruptions. The influence of quasi-two-dimensional confinement on buoyant jets or plumes could also be studied. This problem is relevant to the study of natural ventilation in buildings with line sources of heat. The question of the stability of the flow and the conditions of emergence of the core and eddy structure can be raised when buoyancy forces play an important role. Moreover, one might investigate whether the entrainment, transport, dispersion and mixing mechanisms in quasi-two-di- mensional buoyant jets or quasi-two-dimensional plumes are analogous to those in the non-buoyant case studied in this thesis. Finally, the fundamental modelling of the turbulence in the flow of quasi-two-di- mensional jets could be investigated. Dracos et al. (1992) found an inverse cascade of turbulence at scales larger than the gap width W and a three-dimensional cascade of turbulence at smaller scales. Thus, the turbulence in quasi-two-dimen- sional jets is neither purely two-dimensional nor exactly three-dimensional. The study of the transfer of energy in this quasi-two-dimensional cascade of turbulence could provide some insight about the general problem of turbulence. 202 Appendix A Advection–diffusion model for quasi-two-dimensional jets A.1 Proof of equation (3.89) If t > T0, we have, according to (3.88), φT0(z, t) = KT0 T0 z1/2 (∫ +∞ s(t) ha−1e−h dh− ∫ +∞ s(t−T0) ha−1e−h dh ) , (A.1) with KT0 = 2B 3KdM01/2Γ [a+ 1] , a = 2 3 (Ka Kd − 1 2 ) and s(t) = 4z 3/2 9KdM01/2t . (A.2a–c) 203 A Advection–diffusion model for quasi-two-dimensional jets Combining the two integrals, in the limit t≫ T0, (A.1) becomes φT0(z, t) ∼ KT0 T0 z1/2 (s(t))a−1 e−s(t) (s(t− T0)− s(t)) (A.3) ∼ KT0 z1/2 t (s(t)) a e−s(t) (A.4) Using η = z/ ( t2/3M01/3 ) , (A.2b) and (A.2c), we obtain φT0(z, t) ∼ t−2/3KT0M01/6 ( 4 9Kd )a ηKa/Kd exp [ − 4 9Kd η3/2 ] . (A.5) Finally, using (A.2a), we find φT0(z, t) ∼t−2/3 B ( 3 2 )2a+1 (Kd)a+1 Γ [a+ 1]M01/3 ηKa/Kd exp [ − 4 9Kd η3/2 ] (A.6) =t−2/3yδ(η), (A.7) where yδ is defined in (3.73), and hence (3.89) follows. 204 Appendix B Two-point statistics in circular distributions B.1 Conditional probability for the x-distance be- tween two points in a disc In equation (5.15), the conditional probability for the x- or lateral distance be- tween two points H (x1,x2) (with 0 ≤ x1 ≤ R(t) fixed and x12 + z12 ≤ R2(t)) is, using Cartesian coordinates, PDt(H (x1,x2) ≤ h |x1) = 1 πR2(t) ∫∫ Dt∩H (x1,h ) dz2dx2, (B.1) where the domain H (x1, h ) is defined such that x2 ∈ H (x1, h ) if |x1 − x2| ≤ h . Since H (x1, h ) does not depend on z2 (as long as x22 + z22 ≤ R2(t)) and 205 B Two-point statistics in circular distributions Dt ∩ H (x1, h ) is symmetric with respect to the x-axis, (B.1) becomes, for 0 ≤ h ≤ R(t), PDt(H (x1,x2) ≤ h |x1) = 2 πR2(t) ∫ x1+h x1−h √ R2(t)− x22 dx2, 0 ≤ x1 ≤ R(t)− h (B.2) and PDt(H (x1,x2) ≤ h |x1) = 2 πR2(t) ∫ R(t) x1−h √ R2(t)− x22dx2, R(t)−h ≤ x1 ≤ R(t); (B.3) for R(t) ≤ h ≤ 2R(t), PDt(H (x1,x2) ≤ h |x1) = 2 πR2(t) ∫ R(t) −R(t) √ R2(t)− x22 dx2, 0 ≤ x1 ≤ −R(t) + h , (B.4) and PDt(H (x1,x2) ≤ h |x1) = 2 πR2(t) × ∫ R(t) x1−h √ R2(t)− x22 dx2, −R(t) + h ≤ x1 ≤ R(t); (B.5) and for 2R(t) ≤ h , PDt(H (x1,x2) ≤ h |x1) = 2 πR2(t) ∫ R(t) −R(t) √ R2(t)− x22 dx2, 0 ≤ x1 ≤ R(t). (B.6) Solving the integrals in (B.2), (B.3), (B.4), (B.5) and (B.6), we obtain the results described in (5.16), (5.17), (5.18), (5.19) and (5.20), respectively. 206 B Two-point statistics in circular distributions B.2 Value at the origin for the p.d.f. of the lateral distance between two points in a disc The value at the origin for the p.d.f. of the lateral distance between two points in a disc fHDt is defined as fHDt (0) = lim δh→0 fHDt (δh ) δh (B.7) = lim δh→0 4 πR2(t) ∫ R(t) 0 PDt(H (x1,x2) ≤ δh |x1) δh √ R2(t)− x12 dx1 (B.8) ≈ 4πR2(t) ∫ R(t) 0 lim δh→0 PDt(H (x1,x2) ≤ δh |x1) δh √ R2(t)− x12 dx1, (B.9) where we use (5.15). According to (5.16), the conditional probability is, for δh → 0, lim δh→0 PDt(H (x1,x2) ≤ δh |x1) δh = limδh→0 1piδh ( arcsin ( x1+δh R(t) ) +( x1+δh ) R(t) √ 1− ( x1+δh R(t) )2 − arcsin ( x1−δh R(t) ) − (x1−δh )R(t) √ 1− ( x1−δh R(t) )2 ) (B.10) = 2piR(t) dd( x1R(t)) [ arcsin( x1R(t))+( x1 R(t)) √ 1−( x1R(t)) 2 ] (B.11) = 4 πR(t) √ 1− ( x1 R(t) )2 . (B.12) We can replace (B.12) in (B.9), and integrate to find fHDt (0) = 32 3π2R(t) . (B.13) 207 B Two-point statistics in circular distributions B.3 Value at 2R(t) for the p.d.f. of the lateral distance between two points in a disc The value at h = 2R(t) for the p.d.f. of the lateral distance between two points in a disc, fHDt is defined as fHDt (2R(t)) = lim δh→0 fHDt (2R(t))− fHDt ( 2R(t)− δh ) δh (B.14) = lim δh→0 1− fHDt ( 2R(t)− δh ) δh , (B.15) where we use (5.20), then using to (5.18) and (5.19) we can show that for δh ≪ 1 fHDt ( 2R(t)− δh ) = 1 +O [ δh ]3/2 . (B.16) Therefore, replacing (B.16) in (B.15), we obtain fHDt (2R(t)) = 0. (B.17) B.4 P.d.f of the x-distance between two points in a square domain The c.d.f. of the x− or lateral distance between two points in a square domain St (described in § 5.3.1) is, using Cartesian coordinates, FHSt (h ) = PSt(H (x1,x2) ≤ h ) = 1 R(t) ∫ R(t) 0 PSt(H (x1,x2) ≤ h |x1) dx1, (B.18) based on the general definition (5.7), and where we use the fact that the domain St is symmetric with respect to the z-axis, the density is uniform over the whole domain Dt and the conditional probability does not depend on z1 (as long as −R(t) ≤ z1 ≤ R(t)). Similarly to the conditional probability for the disc described 208 B Two-point statistics in circular distributions in § B.1, the conditional probability for the domain St is PSt(H (x1,x2) ≤ h |x1) = 1 4R2(t) ∫ St∩H (x1,h ) dz2dx2, (B.19) where the domain H (x1, h ) is defined such that x2 ∈ H (x1, h ) if |x1− x2| ≤ h . We can integrate (B.19) with respect to both x1 and z1. We find, for 0 ≤ h ≤ R(t), PSt(H (x1,x2) ≤ h |x1) =    h R(t) 0 ≤ x1 ≤ R(t)− h R(t)− (x1 − h ) 2R(t) R(t)− h ≤ x1 ≤ R(t) ; (B.20) for R(t) ≤ h ≤ 2R(t), PSt(H (x1,x2) ≤ h |x1) =    1 2R(t) 0 ≤ x1 ≤ −R(t) + h R(t)− (x1 − h ) 2R(t) −R(t) + h ≤ x1 ≤ R(t) ; (B.21) and for 2R(t) ≤ h , PSt(H (x1,x2) ≤ h |x1) = 1 2R(t) . (B.22) Using (B.20), (B.21) and (B.22) we can integrate (B.18) to find FHSt (h ) = 4R(t)h − h 2 4R2(t) . (B.23) Finally, we can differentiate (B.23) with respect to h to obtain the p.d.f. fHSt described in (5.22). B.5 Conditional probability for the Euclidean di- stance between two points in a disc In equation (5.23), the conditional probability for the distance between two points D (x1,x2) (with 0 ≤ x1 ≤ R(t) fixed and x12 + z12 ≤ R2(t)) is, using Cartesian 209 B Two-point statistics in circular distributions coordinates, PDt(D (x1,x2) ≤ d |x1) = 1 πR2(t) ∫∫ Dt∩J (x1,d ) dz2dx2, (B.24) where the domain J (x1, h ) is defined such that x2 ∈ J (x1, d ) if (x1 − x2)2 + z22 ≤ d 2 (we can fix z1 = 0 by axisymmetry). (B.24) becomes, for 0 ≤ d ≤ R(t), PDt(D (x1,x2) ≤ d |x1) = 1 πR2(t) ∫ d 0 ∫ 2pi 0 r2dr2dθ2, 0 ≤ r1 ≤ R(t)−d , (B.25) with r1 = √ x12 + z12 and r2 = √ x22 + z22, and PDt(D (x1,x2) ≤ d |x1) = 2 πR2(t) (∫ xI r1−d √ d 2 − (x2 − r1)2 dx2 + ∫ R(t) xI √ R2(t)− x22 dx2 ) , R(t)− d ≤ r1 ≤ R(t), (B.26) where xI is defined in (5.26); for R(t) ≤ d ≤ 2R(t), PDt(D (x1,x2) ≤ d |x1) = 1 πR2(t) ∫ R(t) 0 ∫ 2pi 0 r2 dr2dθ2, 0 ≤ r1 ≤ −R(t) + d , (B.27) and PDt(D (x1,x2) ≤ d |x1) = 2 πR2(t) (∫ xI r1−d √ d 2 − (x2 − r1)2 dx2 + ∫ R(t) xI √ R2(t)− x22 dx2 ) , −R(t) + d ≤ r1 ≤ R(t); (B.28) and for 2R(t) ≤ d , PDt(D (x1,x2) ≤ d |x1) = 1 πR2(t) ∫ R(t) 0 ∫ 2pi 0 r2 dr2dθ2, 0 ≤ r1 ≤ R(t). (B.29) 210 B Two-point statistics in circular distributions Solving the integrals in (B.25), (B.26), (B.27), (B.28) and (B.29), we obtain the results described in (5.24), (5.25), (5.27), (5.28) and (5.29), respectively. B.6 Conditional probability for the ratio of the lateral distance to the streamwise distance between two points in a disc In equation (5.31), the conditional probability for the ratio of the lateral distance to the streamwise distance between two points M (x1,x2) (with 0 ≤ r1 ≤ R(t) and 0 ≤ θ1 ≤ π/2 fixed) is, using polar coordinates, PDt(M (x1,x2) ≤ m |x1) = 1 πR2(t) ∫∫ Dt∩G (x1,m ) r2 dr2dθ2, (B.30) where the domain G (x1,m ) is defined such that x2 ∈ G (x1, d ) if |x1 − x2|/|z1 − z2| ≤ m . (B.30) becomes, for 0 ≤ m , PDt(M (x1,x2) ≤ m |x1) = 1 πR2(t) (∫ pi−υ υ ∫ l 0 r2 dr2dθ2 + ∫ −υ −pi+υ ∫ l 0 r2 dr2dθ2 ) , (B.31) with υ = arctan (1/m ), and where l = r1 cos (θ2 − θ1 − π) + √ R2(t)− r12 sin2 (θ2 − υ − π) (B.32) is the equation of the perimeter of Dt in polar coordinates with the origin at x1. We can then integrate (B.31) with respect to r2. We obtain PDt(M (x1,x2) ≤ m |x1) = 1 πR2(t) (∫ pi−υ υ l2 2 dθ2 + ∫ −υ −pi+υ l2 2 dθ2 ) . (B.33) Finally, replacing (B.32) into (B.33) and integrating (B.33) with respect to θ2, we find equations (5.32) and (5.33). 211 Bibliography Abramowitz, M. & Stegun, I. A. 1972 Handbook of Mathematical Func- tions with Formulas, Graphs, and Mathematical Tables, 10th Edition. Dover Publications. Albertson, M. L., Dai, Y. B., Jensen, R. A. & Rouse, H. 1950 Diffusion of submerged jets. Transactions of the American Society of Civil Engineers 115, 639–664. Antonia, R. A., Satyaprakash, B. R. & Hussain, A. K. M. F. 1980 Measurements of dissipation rate and some other characteristics of turbulent plane and circular jets. Physics of Fluids 23, 695–700. Baines, W. D., Turner, J. S. & Campbell, I. H. 1990 Turbulent fountains in an open chamber. Journal of Fluid Mechanics 212, 557–592. Batchelor, G. K. 1950 The application of the similarity theory of turbulence to atmospheric diffusion. Quarterly Journal of the Royal Meteorological Society 76, 133–146. Batchelor, G. K. 1952 Diffusion in a field of homogeneous turbulence. 2 The relative motion of particles. Proceedings of the Cambridge Philosophical Society 48, 345–362. 213 Bibliography Bird, R. B., Stewart, W. E. & Lightfoot, E. N. 2007 Transport Phenom- ena, 2nd Edition. John Wiley & Sons. Bourgoin, M., Ouellette, N. T., Xu, H. T., Berg, J. & Bodenschatz, E. 2006 The role of pair dispersion in turbulent flow. Science 311, 835–838. Bower, D. J., Caulfield, C. P., Fitzgerald, S. D. & Woods, A. W. 2008 Transient ventilation dynamics following a change in strength of a point source of heat. Journal of Fluid Mechanics 614, 15–37. Cardoso, S. S. S. & Zarrebini, M. 2002 Sedimentation from surface currents generated by particle-laden jets. Chemical Engineering Science 57, 1425–1437. Cenedese, C. & Dalziel, S. B. 1998 Concentration and depth fields deter- mined by the light transmitted through a dyed solution. In Proceedings of the 8th International Symposium on Flow Visualization. Chen, D. & Jirka, G. H. 1998 Linear stability analysis of turbulent mixing layers and jets in shallow water layers. Journal of Hydraulic Research 36, 815– 830. Chen, D. & Jirka, G. H. 1999 LIF study of plane jet bounded in shallow water layer. Journal of Hydraulic Engineering – ASCE 125, 817–826. Cierpka, C. & Kaehler, C. J. 2012 Particle imaging techniques for volumetric three-component (3D3C) velocity measurements in microfluidics. Journal of Visualization 15, 1–31. Colomer, J., Casamitjana, X. & Fernando, H. J. S. 1998 Resuspension and sedimentation of particles from a sediment bed by turbulent jets. Applied Scientific Research 59, 229–242. Colomer, J. & Fernando, H. J. S. 1996 Resuspension of particle bed by round vertical jet. Journal of Environmental Engineering – ASCE 122, 864– 869. Coomaraswamy, I. A. 2011 Natural ventilation of buildings: time-dependent phenomena. PhD thesis, University of Cambridge, Cambridge, UK. 214 Bibliography Dalziel, S. B. 1992 Decay of rotating turbulence - Some particle tracking ex- periments. Applied Scientific Research 49, 217–244. Dalziel, S. B., Patterson, M. D., Caulfield, C. P. & Coomaraswamy, I. A. 2008 Mixing efficiency in high-aspect-ratio Rayleigh–Taylor experiments. Physics of Fluids 20 (065106). De Young, D. S. 1997 Growth of large scale structures in two-dimensional mixing layers. Physics of Fluids 9, 2168–2170. Dimotakis, P. E., Miake-Lye, R. C. & Papantoniou, D. A. 1983 Structure and dynamics of round turbulent jets. Physics of Fluids 26, 3185–3192. Dracos, T., Giger, M. & Jirka, G. H. 1992 Plane turbulent jets in a bounded fluid layer. Journal of Fluid Mechanics 241, 587–614. Drayton, M. J. 1993 Eulerian and Lagrangian studies of inhomogeneous tur- bulence generated by an oscillating grid. PhD thesis, University of Cambridge, Cambridge, UK. Dutkiewicz, S., Griffa, A. & Olson, D. B. 1993 Particle diffusion in a meandering jet. Journal of Geophysical Research–Oceans 98, 16,487–16,500. Ernst, G. G. J., Sparks, R. S. J., Carey, S. N. & Bursik, M. I. 1996 Sedimentation from turbulent jets and plumes. Journal of Geophysical Research 101, 5575–5589. Eser, S., Jenkins, R. G., Derbyshire, F. J. & Malladi, M. 1986 Car- bonization of coker feedstocks and their fractions. Carbon 24, 77–82. Foss, J. F. & Jones, J. B. 1968 Secondary flow effects in a bounded rectangular jet. Transactions of the ASME: Journal of Basic Engineering 90, 241–248. Fung, J. C. H. 1990 Kinematic simulation of turbulent flow and particle motion. PhD thesis, University of Cambridge, Cambridge, UK. Giger, M., Dracos, T. & Jirka, G. H. 1991 Entrainment and mixing in plane turbulent jets in shallow water. Journal of Hydraulic Research 29, 615–642. 215 Bibliography Cervantes de Gortari, J. & Goldschmidt, V. W. 1981 The apparent flap- ping motion of a turbulent plane jet – further experimental results. Transactions of the ASME: Journal of Fluids Engineering 103, 119–126. Gradshteyn, I. S. & Ryzhik, I. M. 2007 Table of Integrals, Series, and Products, 7th Edition. Elsevier Academic Press. Holdeman, J. D. & Foss, J. F. 1975 Initiation, development, and decay of secondary flow in a bounded jet. Transactions of the ASME: Journal of Fluids Engineering 97, 342–352. Hussein, H. J., Capp, S. P. & George, W. K. 1994 Velocity measurements in a high-Reynolds number, momentum-conserving, axisymmetric, turbulent jet. Journal of Fluid Mechanics 258, 31–75. Itoˆ, S. 1992 Diffusion Equations . American Mathematical Society. Jiang, J. S., Law, A. W. K. & Cheng, N. S. 2005 Two-phase analysis of vertical sediment-laden jets. Journal of Engineering Mechanics – ASCE 131, 308–318. Jirka, G. H. 2001 Large scale flow structures and mixing processes in shallow flows. Journal of Hydraulic Research 39, 567–573. Jirka, G. H. 2006 Integral model for turbulent buoyant jets in unbounded strat- ified flows. Part 2. Plane jet dynamics resulting from multiport diffuser jets. Environmental Fluid Mechanics 6, 43–100. Jirka, G. H. & Harleman, D. R. F. 1979 Stability and mixing of a vertical plane buoyant jet in confined depth. Journal of Fluid Mechanics 94, 275–304. Jirka, G. H. & Uijttewaal, W. S. J. 2004 Shallow flows: a definition. In Shallow Flows (ed. G. H. Jirka & W. S. J. Uijttewaal), pp. 3–11. Taylor & Francis. Joshi, P. B. & Taylor, R. B. 1983 Circulation induced by tidal jets. Journal of Waterway Port Coastal and Ocean Engineering – ASCE 109, 445–464. Kahane, J.-P. & Lemarie´-Rieusset, P.-G. 1998 Se´ries de Fourier et On- delettes . Cassini. 216 Bibliography Kitzhofer, J., Nonn, T. & Bru¨cker, C. 2011 Generation and visualization of volumetric PIV data fields. Experiments in Fluids 51, 1471–1492. Ko¨rner, T. W. 1988 Fourier Analysis . Cambridge University Press. Kotsovinos, N. E. 1975 A study of the entrainment and turbulence in a plane buoyant jet. PhD thesis, California Institute of Technology, Pasadena, USA. Kotsovinos, N. E. 1976 A note on the spreading rate and virtual origin of a plane turbulent jet. Journal of Fluid Mechanics 77, 305–311. Kotsovinos, N. E. 1978 A note on the conservation of the axial momentum of a turbulent jet. Journal of Fluid Mechanics 87, 55–63. Kotsovinos, N. E. & Angelidis, P. B. 1991 The momentum flux in turbulent submerged jets. Journal of Fluid Mechanics 229, 453–470. Kotsovinos, N. E. & List, E. J. 1977 Plane turbulent buoyant jets. Part 1. Integral properties. Journal of Fluid Mechanics 81, 25–44. Kuang, J., Hsu, C.-T. & Qiu, H. 2001 Experiments on vertical turbulent plane jets in water of finite depth. Journal of Engineering Mechanics – ASCE 127, 18–26. Landel, J. R., Caulfield, C. P. & Woods, A. W. 2012a Meandering due to large eddies and the statistically self-similar dynamics of quasi-two-dimensional jets. Journal of Fluid Mechanics 692, 347–368. Landel, J. R., Caulfield, C. P. & Woods, A. W. 2012b, sub judice Streamwise dispersion and mixing in quasi-two-dimensional steady turbulent jets. Journal of Fluid Mechanics. . Law, A. W. K. 2006 Velocity and concentration distributions of round and plane turbulent jets. Journal of Engineering Mathematics 56, 69–78. Lee, J. M., Baker, J. J., Murray, D., Llerena, R. & Rolle, J. G. 1997 Quality analysis of petroleum cokes and coals for export specifications required in use of specialty products and utility fuels. In Preprints of Symposia, 214th National Meeting, American Chemical Society 42, pp. 844–853. 217 Bibliography List, E. J. 1982 Turbulent jets and plumes. Annual Review of Fluid Mechanics 14, 189–212. Luo, K., Klein, M., Fan, J.-R. & Cen, K.-F. 2006 Effects on particle dis- persion by turbulent transition in a jet. Physics Letters A 357, 345–350. Mathieu, J. & Scott, J. 2000 An Introduction to Turbulent Flow . Cambridge University Press. Monin, A. S. & Yaglom, A. M. 1975 Statistical Fluid Mechanics, Volume 2. MIT Press. Morton, B. R., Taylor, G. I. & Turner, J. S. 1956 Turbulent gravitational convection from maintained and instantaneous sources. Proceedings of the Royal Society of London. Series A 234, 1–23. National Institute of Standards and Technology 2011-08-29 Digital Library of Mathematical Functions. http://dlmf.nist.gov/. Neves, M. J. & Fernando, H. J. S. 1995 Sedimentation of particles from jets discharged by ocean outfalls: a theoretical and laboratory study. Water Science and Technology 32, 133–139. Ojha, R., Menon, N. & Durian, D. J. 2000 Hysteresis and packing in gas- fluidized beds. Physical Review E 62, 4442–4445. Paranthoe¨n, P., Fouari, A., Dupont, A. & Lecordier, J. C. 1988 Dis- persion measurements in turbulent flows (boundary layer and plane jet). Inter- national Journal of Heat and Mass Transfer 31, 153–165. Picano, F., Sardina, G., Gualtieri, P. & Casciola, C. M. 2010 Anoma- lous memory effects on transport of inertial particles in turbulent jets. Physics of Fluids 22 (051705). Pope, S. B. 1985 Pdf methods for turbulent reactive flows. Progress in Energy and Combustion Science 11, 119–192. Pope, S. B. 2000 Turbulent Flows . Cambridge University Press. 218 Bibliography Prandtl, L. 1925 A report on testing for built-up turbulence. Zeitschrift fu¨r Angewandte Mathematik und Mechanik 5, 136–139. Ramaprian, B. R. & Chandrasekhara, M. S. 1985 LDA measurements in plane turbulent jets. Transactions of the ASME: Journal of Fluids Engineering 107, 264–271. Revuelta, A., Sa´nchez, A. L. & Lin˜a´n, A. 2002 Confined axisymmetric laminar jets with large expansion ratios. Journal of Fluid Mechanics 456, 319– 352. Richardson, L. F. 1926 Atmospheric diffusion shown on a distance-neighbour graph. Proceedings of the Royal Society of London. Series A 110, 709–737. Robinson, S. K. 1991 Coherent motions in the turbulent boundary layer. Annual Review of Fluid Mechanics 23, 601–639. Rowland, J. C., Stacey, M. T. & Dietrich, W. E. 2009 Turbulent char- acteristics of a shallow wall-bounded plane jet: experimental implications for river mouth hydrodynamics. Journal of Fluid Mechanics 627, 423–449. Salazar, J. P. L. C. & Collins, L. R. 2009 Two-particle dispersion in isotropic turbulent flows. Annual Review of Fluid Mechanics 41, 405–432. Sawford, B. 2001 Turbulent relative dispersion. Annual Review of Fluid Me- chanics 33, 289–317. Schefer, R. W., Kerstein, A. R., Namazian, M. & Kelly, J. 1994 Role of large-scale structure in a nonreacting turbulent CH4 jet. Physics of Fluids 6, 652–661. Schneider, W. 1981 Flow induced by jets and plumes. Journal of Fluid Me- chanics 108, 55–65. Schneider, W. 1985 Decay of momentum flux in submerged jets. Journal of Fluid Mechanics 154, 91–110. Shinneeb, A.-M., Bugg, J. D. & Balachandar, R. 2011 Coherent struc- tures in shallow water jets. Transactions of the ASME: Journal of Fluids En- gineering 133, 011203. 219 Bibliography Sparks, R. S. J. 1986 The dimensions and dynamics of volcanic eruption columns. Bulletin of Volcanology 48, 3–15. Stanley, S. A., Sarkar, S. & Mellado, J. P. 2002 A study of the flow-field evolution and mixing in a planar turbulent jet using direct numerical simulation. Journal of Fluid Mechanics 450, 377–407. Sto¨hr, M. & Khalili, A. 2006 Dynamic regimes of buoyancy-affected two- phase flow in unconsolidated porous media. Physical Review E 73, 036301. Sveen, J. K. & Dalziel, S. B. 2005 A dynamic masking technique for com- bined measurements of PIV and synthetic schlieren applied to internal gravity waves. Measurement Science and Technology 16, 1954–1960. Taylor, G. I. 1953 Dispersion of soluble matter in solvent flowing slowly through a tube. Proceedings of the Royal Society of London. Series A 219, 186–203. Taylor, G. I. 1958 Flow induced by jets. Journal of the Aero/Space Sciences 25, 464–465. Thomas, F. O. & Brehob, E. G. 1986 An investigation of large-scale structure in the similarity region of a two-dimensional turbulent jet. Physics of Fluids 29, 1788–1795. Toschi, F. & Bodenschatz, E. 2009 Lagrangian properties of particles in turbulence. Annual Review of Fluid Mechanics 41, 375–404. Turner, J. S. 1986 Turbulent entrainment: the development of the entrain- ment assumption, and its application to geophysical flows. Journal of Fluid Mechanics. 173, 431–471. Uberoi, M. S. & Singh, P. I. 1975 Turbulent mixing in a two-dimensional jet. Physics of Fluids 18, 764–769. Veitch, G. & Woods, A. W. 2000 Particle recycling and oscillations of vol- canic eruption columns. Journal of Geophysical Research 105, 2829–2842. Walters, A. L., Phillips, J. C., Brown, R. J., Field, M., Gernon, T., Stripp, G. & Sparks, R. S. J. 2006 The role of fluidisation in the 220 Bibliography formation of volcaniclastic kimberlite: grain size observations and experimental investigation. Journal of Volcanology and Geothermal Research 155, 119–137. Wang, H. W. & Law, A. W.-K. 2002 Second-order integral model for a round turbulent buoyant jet. Journal of Fluid Mechanics 459, 397–428. Woods, A. W. & Caulfield, C. P. 1992 A laboratory study of explosive volcanic eruptions. Journal of Geophysical Research 97, 6699–6712. Xu, H. & Bodenschatz, E. 2008 Motion of inertial particles with size larger than Kolmogorov scale in turbulent flows. Physica D 237, 2095–2100. Yang, Y., Crowe, C. T., Chung, J. N. & Troutt, T. R. 2000 Experiments on particle dispersion in a plane wake. International Journal of Multiphase Flow 26, 1583–1607. Yannopoulos, P. & Noutsopoulos, G. 1990 The plane vertical turbulent buoyant jet. Journal of Hydraulic Research 28, 565–580. Yannopoulos, P. C. 2006 An improved integral model for plane and round turbulent buoyant jets. Journal of Fluid Mechanics 547, 267–296. Yeung, P. K. 2002 Lagrangian investigations of turbulence. Annual Review of Fluid Mechanics 34, 115–142. Zauner, E. 1985 Visualization of the viscous flow induced by a round jet. Journal of Fluid Mechanics 154, 111–119. Zoueshtiagh, F. & Merlen, A. 2007 Effect of a vertically flowing water jet underneath a granular bed. Physical Review E 75, 056313–1. 221