Hostname: page-component-745bb68f8f-5r2nc Total loading time: 0 Render date: 2025-02-11T17:15:13.253Z Has data issue: false hasContentIssue false

An empirical expression for $\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}$ on the axis of a slightly heated turbulent round jet

Published online by Cambridge University Press:  22 March 2019

J. Lemay*
Affiliation:
Department of Mechanical Engineering, Université Laval, 1065 avenue de la Médecine, Québec City, QC G1V 0A6, Canada
L. Djenidi
Affiliation:
Discipline of Mechanical Engineering, School of Engineering, University of Newcastle, Newcastle, 2308 NSW, Australia
R. A. Antonia
Affiliation:
Discipline of Mechanical Engineering, School of Engineering, University of Newcastle, Newcastle, 2308 NSW, Australia
A. Benaïssa
Affiliation:
Department of Mechanical and Aerospace Engineering, Royal Military College of Canada, PO Box 17000, Station Forces, Kingston, ON K7K 7B4, Canada
*
Email address for correspondence: jean.lemay@gmc.ulaval.ca

Abstract

Self-preservation analyses of the equations for the mean temperature and the second-order temperature structure function on the axis of a slightly heated turbulent round jet are exploited in an attempt to develop an analytical expression for $\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}$, the mean dissipation rate of $\overline{\unicode[STIX]{x1D703}^{2}}/2$, where $\overline{\unicode[STIX]{x1D703}^{2}}$ is the temperature variance. The analytical approach follows that of Thiesset et al. (J. Fluid Mech., vol. 748, 2014, R2) who developed an expression for $\unicode[STIX]{x1D716}_{k}$, the mean turbulent kinetic energy dissipation rate, using the transport equation for $\overline{(\unicode[STIX]{x1D6FF}u)^{2}}$, the second-order velocity structure function. Experimental data show that complete self-preservation for all scales of motion is very well satisfied along the jet axis for streamwise distances larger than approximately 30 times the nozzle diameter. This validation of the analytical results is of particular interest as it provides justification and confidence in the analytical derivation of power laws representing the streamwise evolution of different physical quantities along the axis, such as: $\unicode[STIX]{x1D702}$, $\unicode[STIX]{x1D706}$, $\unicode[STIX]{x1D706}_{\unicode[STIX]{x1D703}}$, $R_{U}$, $R_{\unicode[STIX]{x1D6E9}}$ (all representing characteristic length scales), the mean temperature excess $\unicode[STIX]{x1D6E9}_{0}$, the mixed velocity–temperature moments $\overline{u\unicode[STIX]{x1D703}^{2}}$, $\overline{v\unicode[STIX]{x1D703}^{2}}$ and $\overline{\unicode[STIX]{x1D703}^{2}}$ and $\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}$. Simple models are proposed for $\overline{u\unicode[STIX]{x1D703}^{2}}$ and $\overline{v\unicode[STIX]{x1D703}^{2}}$ in order to derive an analytical expression for $A_{\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}$, the prefactor of the power law describing the streamwise evolution of $\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}$. Further, expressions are also derived for the turbulent Péclet number and the thermal-to-mechanical time scale ratio. These expressions involve global parameters that are most likely to be influenced by the initial and/or boundary conditions and are therefore expected to be flow dependent.

Type
JFM Papers
Copyright
© 2019 Cambridge University Press 

1 Introduction

Turbulent jets involving a scalar continue to interest researchers for both fundamental reasons and because of many industrial applications. The latter can range from the dispersal of pollution to the design of devices optimizing the mixing of a scalar quantity. An important characteristic of the round jet is that all scales of motion can satisfy the requirements of self-preservation. Particularly along its axis, turbulence decays according to well-defined power laws. As a consequence, the turbulent and local Reynolds numbers (Djenidi et al. Reference Djenidi, Antonia, Lefeuvre and Lemay2016, hereafter referred to as DALL), defined as $Re_{\unicode[STIX]{x1D706}}\equiv \sqrt{\overline{u^{2}}}\unicode[STIX]{x1D706}/\unicode[STIX]{x1D708}$ and $Re_{0}\equiv U_{0}R_{U}/\unicode[STIX]{x1D708}$ , respectively are constant. Here, the overbar denotes time-averaged quantities defining one-point statistics, $\unicode[STIX]{x1D708}$ is the kinematic viscosity, $\unicode[STIX]{x1D706}\equiv (\overline{u^{2}}/\overline{(\unicode[STIX]{x2202}u/\unicode[STIX]{x2202}x)^{2}})^{1/2}$ is the Taylor microscale. In what follows, $\unicode[STIX]{x1D716}_{k}$ is the mean dissipation rate of $k$ , the turbulent kinetic energy, $k\equiv (\overline{u^{2}}+\overline{v^{2}}+\overline{w^{2}})/2$ ; $u$ , $v$ and $w$ are respectively the fluctuating velocity components in the streamwise ( $x$ ), radial ( $r$ ) and azimuthal ( $\unicode[STIX]{x1D711}$ ) directions, $U_{0}$ is the mean streamwise velocity on the jet axis and $R_{U}$ is the jet half-radius. The constancy of $Re_{\unicode[STIX]{x1D706}}$ with $x$ implies that self-preservation should apply irrespective of the scaling parameters used, provided they comply with self-preservation. The combinations ( $\unicode[STIX]{x1D702}\equiv (\unicode[STIX]{x1D708}^{3}/\unicode[STIX]{x1D716}_{k})^{1/4}$ and $\unicode[STIX]{x1D710}_{K}\equiv (\unicode[STIX]{x1D708}\unicode[STIX]{x1D716}_{k})^{1/4}$ ) or ( $\unicode[STIX]{x1D706}$ and $\sqrt{\overline{u^{2}}}$ ), ( $R_{U}$ and $U_{0}$ ) can therefore be employed interchangeably (see DALL).

It is well established from free shear flow similarity analysis (see for example Tennekes & Lumley Reference Tennekes and Lumley1972; Pope Reference Pope2000) that a turbulent round jet spreads linearly, viz. $R_{U}\propto x$ , and that $U_{0}\propto x^{-1}$ and $\overline{u^{2}}\propto x^{-2}$ . These results are traditionally derived from the mean flow equations written by assuming the thin shear-layer approximation. More recently, starting with the transport equation for $\overline{(\unicode[STIX]{x1D6FF}u)^{2}}$ or, equivalently, a scale-by-scale (sbs) kinetic energy budget equation (Burattini, Antonia & Danaila Reference Burattini, Antonia and Danaila2005a ), Thiesset, Antonia & Djenidi (Reference Thiesset, Antonia and Djenidi2014) (hereafter referred to as TAD) carried out a two-point self-preservation analysis on the axis of a turbulent round jet. They showed that $\unicode[STIX]{x1D716}_{k}$ evolves longitudinally as a $-4$ th power law, viz.

(1.1) $$\begin{eqnarray}\frac{\unicode[STIX]{x1D716}_{k}D}{U_{j}^{3}}=A_{\unicode[STIX]{x1D716}_{k}}\hat{x}^{-4},\end{eqnarray}$$

where $\hat{x}=(x-x_{0})/D$ ( $x_{0}$ is a virtual origin, $D$ is the nozzle diameter, $U_{j}$ is the jet exit velocity), and $A_{\unicode[STIX]{x1D716}_{k}}$ is a constant to be determined. The assumptions made by these authors (isotropy of dissipative scales and negligible contributions from turbulent and pressure diffusion terms in the $k$ budget) differ from the usual assumption ( $C_{\unicode[STIX]{x1D716}}=\unicode[STIX]{x1D716}_{k}L_{u}/u^{\prime 3}=\text{const.}$ ) which underpins the well-known $-4$ th power law for $\unicode[STIX]{x1D716}_{k}$ in the far field of a round jet (see for example Friehe, Van Atta & Gibson Reference Friehe, Van Atta and Gibson1971; Antonia, Satyaprakash & Hussain Reference Antonia, Satyaprakash and Hussain1980; Mi, Xu & Zhou Reference Mi, Xu and Zhou2013). Further, DALL derived the following power laws, $U_{0}\propto \hat{x}^{-1}$ , $k\propto \hat{x}^{-2}$ , $\unicode[STIX]{x1D716}_{k}\propto \hat{x}^{-4}$ and $l_{q}\propto \hat{x}$ (where $l_{q}$ is the characteristic length scale) from their analysis based on the application of self-preservation to the sbs energy budget. The main difference, relative to previous studies, relies on the fact that the conservation of (simplified) mean momentum and mean turbulent kinetic energy equations were not required. Of particular relevance to the present study, TAD obtained an expression for the prefactor of the $-4$ th power law for $\unicode[STIX]{x1D716}_{k}$ defined by (1.1):

(1.2) $$\begin{eqnarray}A_{\unicode[STIX]{x1D716}_{k}}=B_{U}^{3}A_{I}^{2}(2+{\mathcal{R}}),\end{eqnarray}$$

where $A_{I}=\sqrt{\overline{u^{2}}}/U_{0}$ and ${\mathcal{R}}=\overline{v^{2}}/\overline{u^{2}}$ , two constant parameters along the jet axis (as required by self-preservation), and $B_{U}$ is the coefficient in $U_{0}/U_{j}=B_{U}\hat{x}^{-1}$ . Considering $R_{U}$ and $U_{0}$ to be the relevant outer local scales for the $k$ budget, the normalized form of the mean dissipation rate can be written as

(1.3) $$\begin{eqnarray}\unicode[STIX]{x1D716}_{k}^{\ast }=\frac{\unicode[STIX]{x1D716}_{k}R_{U}}{U_{0}^{3}}=B_{R_{U}}A_{I}^{2}(2+{\mathcal{R}}).\end{eqnarray}$$

Hereafter, $^{\ast }$ stands for normalization with $R_{U}$ and $U_{0}$ . Based on DALL’s analysis, self-preservation on the jet axis requires that $\unicode[STIX]{x1D716}_{k}^{\ast }=\text{const.}$

Otherwise, from measurements of the one-point turbulent kinetic energy budget, Panchapakesan & Lumley (Reference Panchapakesan and Lumley1993a ) and Darisse, Lemay & Benaïssa (Reference Darisse, Lemay and Benaïssa2015) observed on the jet axis that the mean dissipation rate is essentially compensated by convection ( $C_{k}^{\ast }$ ) and production ( $P_{k}^{\ast }$ ), viz. $\unicode[STIX]{x1D716}_{k}^{\ast }\simeq C_{k}^{\ast }+P_{k}^{\ast }$ . Using self-preservation forms for $k$ , $\overline{U}$ and $\overline{V}$ , Darisse et al. (Reference Darisse, Lemay and Benaïssa2015) derived the explicit relations for $C_{k}^{\ast }$ and $P_{k}^{\ast }$ on the jet axis. Using axisymmetry, the continuity equation, the power-law expressions for $U_{0}$ and $R_{U}$ and the notation introduced by TAD, viz. $A_{I}$ and ${\mathcal{R}}$ , with $k/U_{0}^{2}=A_{I}^{2}({\mathcal{R}}+1/2)$ , the normalized convection and production terms can finally be written as

(1.4) $$\begin{eqnarray}\displaystyle & \displaystyle C_{k}^{\ast }=2B_{R_{U}}A_{I}^{2}({\mathcal{R}}+1/2), & \displaystyle\end{eqnarray}$$
(1.5) $$\begin{eqnarray}\displaystyle & \displaystyle P_{k}^{\ast }=B_{R_{U}}A_{I}^{2}({\mathcal{R}}-1). & \displaystyle\end{eqnarray}$$

The sum of these equations gives

(1.6) $$\begin{eqnarray}\unicode[STIX]{x1D716}_{k}^{\ast }\simeq C_{k}^{\ast }+P_{k}^{\ast }=B_{R_{U}}A_{I}^{2}({\mathcal{R}}+2).\end{eqnarray}$$

This indicates that relation (1.3) represents essentially the sum of the convection and production terms of the normalized form of the $k$ budget written on the jet axis. Darisse et al. (Reference Darisse, Lemay and Benaïssa2015) showed that this relation was in very good agreement with $\unicode[STIX]{x1D716}_{k}$ inferred as the remainder of their $k$ budget.

In this paper we extend the analysis of DALL and TAD to a slightly heated turbulent round jet with a view to developing an analytical expression for $\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}$ on the jet axis. The analysis is restricted to small values of $\unicode[STIX]{x1D6E9}_{j}$ , the mean temperature excess relative to ambient at the jet exit, in order to allow temperature to be treated as a passive scalar. The power law $\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}\propto \hat{x}^{-4}$ , describing the evolution of $\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}$ along the jet axis, is derived in § 2 by applying self-preservation to the sbs budget of $\overline{\unicode[STIX]{x1D703}^{2}}/2$ . This rigorous derivation is new and particularly relevant, because it does not involve any of the ad hoc assumptions usually made by classical and simpler approaches. The development of an expression for the prefactor $A_{\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}$ of this power law is presented in § 2.6. Models for the mixed velocity–temperature moments are then developed in § 3 in order to propose a relatively simple expression (§ 4) for $A_{\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}$ ; we will show that this parameter is not universal since it is influenced by the boundary conditions of the jet. Finally, using these results, simple expressions are also presented in § 4 for the turbulent Péclet number and the thermal-to-mechanical time scale ratio.

2 Outcomes of self-preservation for the scalar on the jet axis

2.1 Power-law evolution of $\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}$ derived from the scale-by-scale budget

The analytical approach supporting the development of a power law expressing the longitudinal evolution of $\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}$ is based on a self-preservation solution of the transport equation for $\overline{(\unicode[STIX]{x1D6FF}\unicode[STIX]{x1D703})^{2}}$ , the second-order temperature structure function; $\unicode[STIX]{x1D6FF}\unicode[STIX]{x1D703}=(\unicode[STIX]{x1D703}^{+}-\unicode[STIX]{x1D703})$ , where $\unicode[STIX]{x1D703}^{+}=\unicode[STIX]{x1D703}(x+s)$ and $\unicode[STIX]{x1D703}=\unicode[STIX]{x1D703}(x)$ depends on the separations $s$ between two points along the jet axis. Starting from the temperature equation written at two independent points in space, and assuming local isotropy and local homogeneity (see Burattini et al. Reference Burattini, Antonia and Danaila2005a , for a discussion on the homogeneity condition), the transport equation for $\overline{(\unicode[STIX]{x1D6FF}\unicode[STIX]{x1D703})^{2}}$ is written as:

(2.1) $$\begin{eqnarray}-\overline{(\unicode[STIX]{x1D6FF}u)(\unicode[STIX]{x1D6FF}\unicode[STIX]{x1D703})^{2}}+2\unicode[STIX]{x1D6FC}\frac{\text{d}\overline{(\unicode[STIX]{x1D6FF}\unicode[STIX]{x1D703})^{2}}}{\text{d}s}+LST(s)=\frac{4}{3}\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}s,\end{eqnarray}$$

where $LST(s)$ represents the contribution of the large scales to the budget of $\overline{(\unicode[STIX]{x1D6FF}\unicode[STIX]{x1D703})^{2}}$ . Here, similar developments to those reported by Antonia et al. (Reference Antonia, Ould-Rouis, Anselmet and Zhu1997), Hill (Reference Hill1997, Reference Hill2001), Danaila et al. (Reference Danaila, Anselmet, Zhou and Antonia1999), Shivamoggi & Antonia (Reference Shivamoggi and Antonia2000) and Burattini, Antonia & Danaila (Reference Burattini, Antonia and Danaila2005b ) are used to derive the large scale term on the jet axis:

(2.2) $$\begin{eqnarray}LST(s)=-C(s)-P(s)-D(s),\end{eqnarray}$$

where

(2.3) $$\begin{eqnarray}\displaystyle & \displaystyle C(s)=\frac{1}{s^{2}}U_{0}\int _{0}^{s}y^{2}\frac{\unicode[STIX]{x2202}\overline{(\unicode[STIX]{x1D6FF}\unicode[STIX]{x1D703})^{2}}}{\unicode[STIX]{x2202}x}\,\text{d}y, & \displaystyle\end{eqnarray}$$
(2.4) $$\begin{eqnarray}\displaystyle & \displaystyle P(s)=\frac{2}{s^{2}}\frac{\text{d}\unicode[STIX]{x1D6E9}_{0}}{\text{d}x}\int _{0}^{s}y^{2}\overline{(\unicode[STIX]{x1D6FF}u)(\unicode[STIX]{x1D6FF}\unicode[STIX]{x1D703})}\,\text{d}y, & \displaystyle\end{eqnarray}$$
(2.5) $$\begin{eqnarray}\displaystyle & \displaystyle D(s)=\frac{1}{2s^{2}}\int _{0}^{s}y^{2}\frac{\unicode[STIX]{x2202}}{\unicode[STIX]{x2202}\overline{x}}\left[\overline{(u^{+}+u)(\unicode[STIX]{x1D6FF}\unicode[STIX]{x1D703})^{2}}\right]\,\text{d}y+\frac{4}{s^{2}}\int _{0}^{s}y^{2}\frac{\unicode[STIX]{x2202}}{\unicode[STIX]{x2202}r}\overline{v(\unicode[STIX]{x1D6FF}\unicode[STIX]{x1D703})^{2}}\,\text{d}y, & \displaystyle\end{eqnarray}$$

where $\overline{x}=x+s/2$ ( $\overline{x}$ being the midpoint), $\unicode[STIX]{x1D6FF}u=(u^{+}-u)$ , $u^{+}=u(x+s)$ and $u=u(x)$ . The three terms defined in (2.3) to (2.5), $C(s)$ , $P(s)$ and $D(s)$ , represent convection, production and turbulent diffusion, respectively. In the limit $s\rightarrow \infty$ , equation (2.1) yields the one-point budget of $\overline{\unicode[STIX]{x1D703}^{2}}/2$ , which is expressed (neglecting molecular diffusion) as

(2.6) $$\begin{eqnarray}-U_{0}\frac{\unicode[STIX]{x2202}\overline{\unicode[STIX]{x1D703}^{2}}/2}{\unicode[STIX]{x2202}x}-\overline{u\unicode[STIX]{x1D703}}\frac{\text{d}\unicode[STIX]{x1D6E9}_{0}}{\text{d}x}-\frac{1}{2}\frac{\unicode[STIX]{x2202}\overline{u\unicode[STIX]{x1D703}^{2}}}{\unicode[STIX]{x2202}x}-\frac{\unicode[STIX]{x2202}\overline{v\unicode[STIX]{x1D703}^{2}}}{\unicode[STIX]{x2202}r}\simeq \unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}},\end{eqnarray}$$

where $\unicode[STIX]{x1D6E9}_{0}$ is the local mean temperature excess relative to the ambient for $r=0$ . The first term on the left-hand side of this expression represents convection ( $C_{\unicode[STIX]{x1D703}}$ ), the second stands for production ( $P_{\unicode[STIX]{x1D703}}$ ) and the last two terms correspond to the longitudinal ( $D_{\unicode[STIX]{x1D703}_{x}}$ ) and radial ( $D_{\unicode[STIX]{x1D703}_{r}}$ ) contributions to the turbulent diffusion ( $D_{\unicode[STIX]{x1D703}}$ ). Darisse, Lemay & Benaïssa (Reference Darisse, Lemay and Benaïssa2014) showed that $D_{\unicode[STIX]{x1D703}_{r}}$ is more than 300 times larger than $D_{\unicode[STIX]{x1D703}_{x}}$ on the jet axis. For the one-point budget, it is thus justified to neglect the longitudinal contribution from the turbulent diffusion term (see § 2.6). However, since this term is not necessarily negligible at all scales, we have included its contribution in the definition of $D(s)$ (2.5) in the sbs budget.

Self-preservation form of the equations (2.1)–(2.5) can be obtained when the different terms are expressed using the following functional forms (see Antonia et al. Reference Antonia, Smalley, Zhou, Anselmet and Danaila2004; Burattini et al. Reference Burattini, Antonia and Danaila2005b )

(2.7) $$\begin{eqnarray}\displaystyle & \displaystyle \overline{(\unicode[STIX]{x1D6FF}\unicode[STIX]{x1D703})^{2}}=\tilde{\unicode[STIX]{x1D703}}^{2}(x)h_{\unicode[STIX]{x1D703}}(s^{\ast }), & \displaystyle\end{eqnarray}$$
(2.8) $$\begin{eqnarray}\displaystyle & \displaystyle \overline{(\unicode[STIX]{x1D6FF}u)^{2}}=\tilde{u} ^{2}(x)h_{u}(s^{\ast }), & \displaystyle\end{eqnarray}$$
(2.9) $$\begin{eqnarray}\displaystyle & \displaystyle \overline{(\unicode[STIX]{x1D6FF}u)(\unicode[STIX]{x1D6FF}\unicode[STIX]{x1D703})}=\tilde{A}_{\unicode[STIX]{x1D703}}(x)a(s^{\ast }), & \displaystyle\end{eqnarray}$$
(2.10) $$\begin{eqnarray}\displaystyle & \displaystyle \overline{(\unicode[STIX]{x1D6FF}u)(\unicode[STIX]{x1D6FF}\unicode[STIX]{x1D703})^{2}}=\tilde{B}_{\unicode[STIX]{x1D703}}(x)b_{1}(s^{\ast }), & \displaystyle\end{eqnarray}$$
(2.11) $$\begin{eqnarray}\displaystyle & \displaystyle \overline{(u^{+}+u)(\unicode[STIX]{x1D6FF}\unicode[STIX]{x1D703})^{2}}=\tilde{B}_{\unicode[STIX]{x1D703}}(x)b_{2}(s^{\ast }), & \displaystyle\end{eqnarray}$$
(2.12) $$\begin{eqnarray}\displaystyle & \displaystyle \overline{v(\unicode[STIX]{x1D6FF}\unicode[STIX]{x1D703})^{2}}=\tilde{C}_{\unicode[STIX]{x1D703}}(x)c(s^{\ast }), & \displaystyle\end{eqnarray}$$

where $s^{\ast }\equiv s/l_{\unicode[STIX]{x1D703}}$ , $l_{\unicode[STIX]{x1D703}}$ being the characteristic length scale for the temperature field. Note that the expression for $\overline{(\unicode[STIX]{x1D6FF}u)^{2}}$ is not directly related to (2.1)–(2.5), but it will be used later on, for the analysis of the mixed skewness increment. The scaling functions $\tilde{\unicode[STIX]{x1D703}}^{2}(x)$ , $\tilde{u} ^{2}(x)$ , $\tilde{A}_{\unicode[STIX]{x1D703}}(x)$ , $\tilde{B}_{\unicode[STIX]{x1D703}}(x)$ , $\tilde{C}_{\unicode[STIX]{x1D703}}(x)$ characterize the streamwise evolution of the structure functions, while the dimensionless ( $s^{\ast }$ ) functions represent the shape of these structure functions. For the mixed velocity–temperature scales, as in Antonia et al. (Reference Antonia, Smalley, Zhou, Anselmet and Danaila2004), we avoid making the a priori assumption that these functions are represented by a combination of uncoupled velocity and temperature scales. For example, $\tilde{A}_{\unicode[STIX]{x1D703}}$ is a mixed scale with dimensions of velocity times temperature, but it is not a priori considered as being represented by the product of a velocity scale with a temperature scale.

Substitution of expressions (2.7)–(2.12) (in which we drop $x$ and $s^{\ast }$ for convenience) into (2.1) and dividing by $(\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}s^{\ast })$ yields, after some manipulation,

(2.13) $$\begin{eqnarray}\displaystyle & & \displaystyle -\left[\frac{\tilde{B}_{\unicode[STIX]{x1D703}}}{l_{\unicode[STIX]{x1D703}}\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}\right]\frac{b_{1}}{s^{\ast }}+\left[\frac{\unicode[STIX]{x1D6FC}\tilde{\unicode[STIX]{x1D703}}^{2}}{{l_{\unicode[STIX]{x1D703}}}^{2}\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}\right]2\frac{\text{d}h_{\unicode[STIX]{x1D703}}}{\text{d}s^{\ast }}-\left[\frac{U_{0}}{\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}\frac{\text{d}\tilde{\unicode[STIX]{x1D703}}^{2}}{\text{d}x}\right]\unicode[STIX]{x1D6E4}_{1}^{\ast }+\left[\frac{U_{0}\tilde{\unicode[STIX]{x1D703}}^{2}}{l_{\unicode[STIX]{x1D703}}\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}\frac{\text{d}l_{\unicode[STIX]{x1D703}}}{\text{d}x}\right]\unicode[STIX]{x1D6E4}_{2}^{\ast }-\left[\frac{\tilde{A}_{\unicode[STIX]{x1D703}}}{\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}\frac{\text{d}\unicode[STIX]{x1D6E9}_{0}}{\text{d}x}\right]\unicode[STIX]{x1D6E4}_{3}^{\ast }\nonumber\\ \displaystyle & & \displaystyle \qquad +\,\left[\frac{\tilde{B}_{\unicode[STIX]{x1D703}}}{l_{\unicode[STIX]{x1D703}}\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}\frac{1}{2}\frac{\text{d}l_{\unicode[STIX]{x1D703}}}{\text{d}\overline{x}}\right]\unicode[STIX]{x1D6E4}_{4}^{\ast }-\left[\frac{1}{\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}\frac{\text{d}\tilde{B}_{\unicode[STIX]{x1D703}}}{\text{d}\overline{x}}\right]\unicode[STIX]{x1D6E4}_{5}^{\ast }-\left[\frac{\tilde{C}_{\unicode[STIX]{x1D703}}}{l_{\unicode[STIX]{x1D703}}\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}\right]\unicode[STIX]{x1D6E4}_{6}^{\ast }=\left[\frac{4}{3}\right],\end{eqnarray}$$

where the functions $\unicode[STIX]{x1D6E4}_{i}^{\ast }$ , which represent the shape functions of the large scale terms ( $LST$ ), are expressed as

(2.14) $$\begin{eqnarray}\displaystyle & \displaystyle \unicode[STIX]{x1D6E4}_{1}^{\ast }=\frac{1}{s^{\ast 3}}\int _{0}^{s^{\ast }}h_{\unicode[STIX]{x1D703}}y^{\ast 2}\,\text{d}y^{\ast }, & \displaystyle\end{eqnarray}$$
(2.15) $$\begin{eqnarray}\displaystyle & \displaystyle \unicode[STIX]{x1D6E4}_{2}^{\ast }=\frac{1}{s^{\ast 3}}\int _{0}^{s^{\ast }}\frac{\text{d}h_{\unicode[STIX]{x1D703}}}{\text{d}y^{\ast }}y^{\ast 3}\,\text{d}y^{\ast }, & \displaystyle\end{eqnarray}$$
(2.16) $$\begin{eqnarray}\displaystyle & \displaystyle \unicode[STIX]{x1D6E4}_{3}^{\ast }=\frac{2}{s^{\ast 2}}\int _{0}^{s^{\ast }}ay^{\ast 2}\,\text{d}y^{\ast }, & \displaystyle\end{eqnarray}$$
(2.17) $$\begin{eqnarray}\displaystyle & \displaystyle \unicode[STIX]{x1D6E4}_{4}^{\ast }=\frac{1}{s^{\ast 3}}\int _{0}^{s^{\ast }}\frac{\text{d}b_{2}}{\text{d}y^{\ast }}y^{\ast 3}\,\text{d}y^{\ast }, & \displaystyle\end{eqnarray}$$
(2.18) $$\begin{eqnarray}\displaystyle & \displaystyle \unicode[STIX]{x1D6E4}_{5}^{\ast }=\frac{1}{2s^{\ast 3}}\int _{0}^{s^{\ast }}b_{2}y^{\ast 2}\,\text{d}y^{\ast }, & \displaystyle\end{eqnarray}$$
(2.19) $$\begin{eqnarray}\displaystyle & \displaystyle \unicode[STIX]{x1D6E4}_{6}^{\ast }=\frac{4}{s^{\ast 3}}\int _{0}^{s^{\ast }}\frac{\text{d}c}{\text{d}y^{\ast }}\frac{\unicode[STIX]{x2202}y^{\ast }}{\unicode[STIX]{x2202}r^{\ast }}y^{\ast 2}\,\text{d}y^{\ast }. & \displaystyle\end{eqnarray}$$

Self-preservation requires that all terms within brackets should evolve in $x$ in exactly the same manner. Since the term on the right-hand side of (2.13) is constant, all the terms within brackets should be constant. This yields the following self-preservation constraints (where $C_{i}$ are constants):

(2.20) $$\begin{eqnarray}\displaystyle & \displaystyle \left[\frac{\tilde{B}_{\unicode[STIX]{x1D703}}}{l_{\unicode[STIX]{x1D703}}\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}\right]=C_{1}, & \displaystyle\end{eqnarray}$$
(2.21) $$\begin{eqnarray}\displaystyle & \displaystyle \left[\frac{\unicode[STIX]{x1D6FC}\tilde{\unicode[STIX]{x1D703}}^{2}}{{l_{\unicode[STIX]{x1D703}}}^{2}\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}\right]=C_{2}, & \displaystyle\end{eqnarray}$$
(2.22) $$\begin{eqnarray}\displaystyle & \displaystyle \left[\frac{U_{0}}{\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}\frac{\text{d}\tilde{\unicode[STIX]{x1D703}}^{2}}{\text{d}x}\right]=C_{3}, & \displaystyle\end{eqnarray}$$
(2.23) $$\begin{eqnarray}\displaystyle & \displaystyle \left[\frac{U_{0}\tilde{\unicode[STIX]{x1D703}}^{2}}{l_{\unicode[STIX]{x1D703}}\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}\frac{\text{d}l_{\unicode[STIX]{x1D703}}}{\text{d}x}\right]=C_{4}, & \displaystyle\end{eqnarray}$$
(2.24) $$\begin{eqnarray}\displaystyle & \displaystyle \left[\frac{\tilde{A}_{\unicode[STIX]{x1D703}}}{\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}\frac{\text{d}\unicode[STIX]{x1D6E9}_{0}}{\text{d}x}\right]=C_{5}, & \displaystyle\end{eqnarray}$$
(2.25) $$\begin{eqnarray}\displaystyle & \displaystyle \left[\frac{\tilde{B}_{\unicode[STIX]{x1D703}}}{l_{\unicode[STIX]{x1D703}}\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}\frac{1}{2}\frac{\text{d}l_{\unicode[STIX]{x1D703}}}{\text{d}\overline{x}}\right]=\left[\frac{\tilde{B}_{\unicode[STIX]{x1D703}}}{l_{\unicode[STIX]{x1D703}}\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}\frac{\text{d}l_{\unicode[STIX]{x1D703}}}{\text{d}x}\right]=C_{6}, & \displaystyle\end{eqnarray}$$
(2.26) $$\begin{eqnarray}\displaystyle & \displaystyle \left[\frac{1}{\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}\frac{\text{d}\tilde{B}_{\unicode[STIX]{x1D703}}}{\text{d}\overline{x}}\right]=\left[\frac{2}{\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}\frac{\text{d}\tilde{B}_{\unicode[STIX]{x1D703}}}{\text{d}x}\right]=C_{7}, & \displaystyle\end{eqnarray}$$
(2.27) $$\begin{eqnarray}\displaystyle & \displaystyle \left[\frac{\tilde{C}_{\unicode[STIX]{x1D703}}}{l_{\unicode[STIX]{x1D703}}\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}\right]=C_{8}. & \displaystyle\end{eqnarray}$$

Following the definition of the midpoint $\overline{x}$ , we can write $\text{d}l_{\unicode[STIX]{x1D703}}/\text{d}\overline{x}=\text{d}l_{\unicode[STIX]{x1D703}}/\text{d}x^{+}+\text{d}l_{\unicode[STIX]{x1D703}}/\text{d}x$ (see for example Hill Reference Hill2001; Danaila et al. Reference Danaila, Krawczynski, Thiesset and Renou2012). Along the jet axis, in the far field, it is considered that $\text{d}l_{\unicode[STIX]{x1D703}}/\text{d}x=\text{d}l_{\unicode[STIX]{x1D703}}/\text{d}x^{+}$ , and the derivative in $C_{6}$ has been written as $\text{d}l_{\unicode[STIX]{x1D703}}/\text{d}\overline{x}=2\text{d}l_{\unicode[STIX]{x1D703}}/\text{d}x$ . The same reasoning applies to the derivative of $\tilde{B}_{\unicode[STIX]{x1D703}}$ involved in $C_{7}$ .

Manipulation of $C_{i}$ constants yields the following constraints:

(2.28) $$\begin{eqnarray}\displaystyle & \displaystyle \frac{C_{6}}{C_{1}}=\frac{\text{d}l_{\unicode[STIX]{x1D703}}}{\text{d}x}, & \displaystyle\end{eqnarray}$$
(2.29) $$\begin{eqnarray}\displaystyle & \displaystyle \frac{C_{1}C_{4}}{C_{2}C_{6}}=\frac{U_{0}l_{\unicode[STIX]{x1D703}}}{\unicode[STIX]{x1D6FC}}, & \displaystyle\end{eqnarray}$$
(2.30) $$\begin{eqnarray}\displaystyle & \displaystyle \frac{C_{5}C_{6}}{C_{1}C_{4}}=\frac{\tilde{A}_{\unicode[STIX]{x1D703}}l_{\unicode[STIX]{x1D703}}}{U_{0}\tilde{\unicode[STIX]{x1D703}}^{2}}\frac{\text{d}\unicode[STIX]{x1D6E9}_{0}}{\text{d}x}, & \displaystyle\end{eqnarray}$$
(2.31) $$\begin{eqnarray}\displaystyle & \displaystyle \frac{C_{6}}{C_{4}}=\frac{\tilde{B}_{\unicode[STIX]{x1D703}}}{U_{0}\tilde{\unicode[STIX]{x1D703}}^{2}}, & \displaystyle\end{eqnarray}$$
(2.32) $$\begin{eqnarray}\displaystyle & \displaystyle \frac{C_{1}}{C_{8}}=\frac{\tilde{B}_{\unicode[STIX]{x1D703}}}{\tilde{C}_{\unicode[STIX]{x1D703}}}. & \displaystyle\end{eqnarray}$$

The two first constraints allow us to write $l_{\unicode[STIX]{x1D703}}\propto \hat{x}$ and, for a given value of $\unicode[STIX]{x1D6FC}$ , $U_{0}\propto \hat{x}^{-1}$ . It is worth mentioning that the usual power law for $U_{0}$ , viz. $U_{0}/U_{j}=B_{U}\hat{x}^{-1}$ , is more formally obtained here than the one resulting from the momentum equation.

In addition, the self-preservation analysis of the mean temperature equation presented in § A.1 strictly shows that $\unicode[STIX]{x1D6E9}_{0}$ , scales with $\hat{x}^{-1}$ (yielding $\text{d}\unicode[STIX]{x1D6E9}_{0}/\text{d}x\propto \hat{x}^{-2}$ ). This self-preservation analysis also shows, from (A 11), that $H_{0}\propto \hat{x}^{-2}$ ( $H_{0}$ represents $\overline{u\unicode[STIX]{x1D703}}$ on the jet axis). It is thus required that $\tilde{A}_{\unicode[STIX]{x1D703}}$ has to scale with $\hat{x}^{-2}$ , because $\tilde{A}_{\unicode[STIX]{x1D703}}=2H_{0}$ when $s\rightarrow \infty$ .

Using the power laws just derived, viz. $l_{\unicode[STIX]{x1D703}}\propto \hat{x}$ , $U_{0}\propto \hat{x}^{-1}$ , $\tilde{A}_{\unicode[STIX]{x1D703}}\propto \hat{x}^{-2}$ and $\text{d}\unicode[STIX]{x1D6E9}_{0}/\text{d}x\propto \hat{x}^{-2}$ into constraint (2.30) yields

(2.33) $$\begin{eqnarray}\tilde{\unicode[STIX]{x1D703}}^{2}\propto \frac{\tilde{A}_{\unicode[STIX]{x1D703}}l_{\unicode[STIX]{x1D703}}}{U_{0}}\frac{\text{d}\unicode[STIX]{x1D6E9}_{0}}{\text{d}x}\propto \hat{x}^{-2}.\end{eqnarray}$$

Finally, introducing $\text{d}\unicode[STIX]{x1D6E9}_{0}/\text{d}x\propto \hat{x}^{-2}$ and $\tilde{A}_{\unicode[STIX]{x1D703}}\propto \hat{x}^{-2}$ into expression (2.24), which defines the fifth constraint $C_{5}$ , yields the $-4$ th power law for $\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}$ :

(2.34) $$\begin{eqnarray}\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}\propto \hat{x}^{-4}.\end{eqnarray}$$

This expression is similar to the $-4$ th power law previously derived by TAD and DALL for $\unicode[STIX]{x1D716}_{k}$ . The mean dissipation rate of $\overline{\unicode[STIX]{x1D703}^{2}}/2$ , normalized by the jet exit parameters, can finally be written as

(2.35) $$\begin{eqnarray}\frac{\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}D}{U_{j}\unicode[STIX]{x1D6E9}_{j}^{2}}=A_{\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}\hat{x}^{-4}.\end{eqnarray}$$

In the context of complete self-preservation, it should be noted that other velocity scales, like $\unicode[STIX]{x1D710}_{K}$ or $\sqrt{\overline{u^{2}}}$ could have been used, interchangeably, instead of $U_{0}$ (see Burattini et al. Reference Burattini, Antonia and Danaila2005b , and TAD), as far as the chosen velocity scales with $\hat{x}^{-1}$ . The same reasoning applies to the temperature scale. For example, the Kolmogorov temperature $\unicode[STIX]{x1D703}_{K}\equiv (\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}\unicode[STIX]{x1D702}/\unicode[STIX]{x1D710}_{K})^{1/2}$ , which scales with $\hat{x}^{-1}$ , could have been used, interchangeably, instead of $\unicode[STIX]{x1D6E9}_{0}$ .

Moreover, should $\tilde{A}_{\unicode[STIX]{x1D703}}$ , defined in relation (2.9), have been a priori expressed as the product of uncoupled velocity and temperature scales, the same final result would have been obtained in a simpler way, by the sole use of the sbs budget. We however prefer to avoid this assumption and use the mean temperature equation to close the problem.

2.2 Power laws for temperature variance and mixed moments

Under the assumption of local homogeneity, the temperature variance can be defined as half the temperature structure function considered for large separations, viz. $\overline{\unicode[STIX]{x1D703}^{2}}=\overline{(\unicode[STIX]{x1D6FF}\unicode[STIX]{x1D703})^{2}}/2$ (for $s\rightarrow \infty$ ). From the previous developments, it follows immediately that $\overline{\unicode[STIX]{x1D703}^{2}}\propto \unicode[STIX]{x1D6E9}_{0}^{2}\propto \hat{x}^{-2}$ . We thus introduce the parameter

(2.36) $$\begin{eqnarray}B_{I}^{2}=\frac{\overline{\unicode[STIX]{x1D703}^{2}}}{\unicode[STIX]{x1D6E9}_{0}^{2}},\end{eqnarray}$$

which must be constant along the jet axis. The term $B_{I}$ for the temperature field is comparable to the parameter $A_{I}=\overline{u^{2}}/U_{0}^{2}$ introduced by TAD for the velocity field. Darisse et al. (Reference Darisse, Lemay and Benaïssa2015) observed in their jet that $B_{I}$ was constant for $\hat{x}>20$ . For the slightly heated round jet studied by Darisse et al. (Reference Darisse, Lemay and Benaïssa2015), the longitudinal distributions of $R_{U}$ , $U_{0}$ , $\unicode[STIX]{x1D6E9}_{0}$ , $\overline{u^{2}}$ , $\overline{\unicode[STIX]{x1D703}^{2}}$ and $\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}$ , all normalized by the jet exit parameters, are presented in figure 1. All these quantities share the same virtual origin $x_{0}=1.65$ which has been obtained from least-square regressions performed over the whole dataset related to the self-preserving region. The log–log scale highlights the power-law behaviour of these quantities. The location corresponding to the self-preserving region depends on the quantity of interest, but it is shown that full self-preservation is clearly reached for $\hat{x}>30$ . Table 1 presents the power-law equation used for each quantity and the (measured) value of their corresponding power-law coefficients.

Figure 1. Streamwise evolution of $R_{U}/D$ , $U_{0}/U_{j}$ , $\unicode[STIX]{x1D6E9}_{0}/\unicode[STIX]{x1D6E9}_{j}$ , $\overline{u^{2}}/U_{j}^{2}$ , $\overline{\unicode[STIX]{x1D703}^{2}}/\unicode[STIX]{x1D6E9}_{j}^{2}$ and $\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}D/(U_{j}\unicode[STIX]{x1D6E9}_{j}^{2})$ along the jet axis (symbols) and their corresponding power laws (measurements of Darisse, Lemay & Benaïssa Reference Darisse, Lemay and Benaïssa2013a ,Reference Darisse, Lemay and Benaïssa b ; Darisse et al. Reference Darisse, Lemay and Benaïssa2015).

Table 1. Definition of the power laws shown in figure 1 and obtained from the measurements of Darisse et al. (Reference Darisse, Lemay and Benaïssa2015) ( $Re_{D}=1.4\times 10^{5}$ , $Re_{\unicode[STIX]{x1D706}}=548$ and $Pe_{\unicode[STIX]{x1D706}}=211$ ).

Recalling from expression (2.9) that $\tilde{A}_{\unicode[STIX]{x1D703}}$ has the dimensions of velocity times temperature and that $\tilde{A}_{\unicode[STIX]{x1D703}}\propto \hat{x}^{-2}$ , it immediately follows that, along the jet axis,

(2.37) $$\begin{eqnarray}\tilde{A}_{\unicode[STIX]{x1D703}}\propto U_{0}\unicode[STIX]{x1D6E9}_{0}\propto \tilde{u} \tilde{\unicode[STIX]{x1D703}}\propto \hat{x}^{-2},\end{eqnarray}$$

where $\tilde{u} \propto \hat{x}^{-1}$ is a characteristic velocity scale (see (2.8)) used by DALL in their sbs analysis of the energy equation. Relation (2.37) indicates that $\tilde{A}_{\unicode[STIX]{x1D703}}$ (a mixed velocity–temperature scale) is finally expressed as the product of uncoupled velocity and temperature scales. This also reveals that the normalized longitudinal temperature flux $\overline{u\unicode[STIX]{x1D703}}/(U_{0}\unicode[STIX]{x1D6E9}_{0})$ is constant along the jet axis.

The temperature scale is now introduced into relations (2.31) and (2.32) which leads to

(2.38) $$\begin{eqnarray}\tilde{B}_{\unicode[STIX]{x1D703}}\propto \tilde{C}_{\unicode[STIX]{x1D703}}\propto U_{0}\unicode[STIX]{x1D6E9}_{0}^{2}\propto \hat{x}^{-3}.\end{eqnarray}$$

These mixed velocity–temperature scales are both represented by the product of an uncoupled velocity scale and a temperature squared scale. The above relations also indicate that the normalized mixed moments, $\overline{u\unicode[STIX]{x1D703}^{2}}/(U_{0}\unicode[STIX]{x1D6E9}_{0}^{2})$ and $\overline{v\unicode[STIX]{x1D703}^{2}}/(U_{0}\unicode[STIX]{x1D6E9}_{0}^{2})$ , are constant along the jet axis. This also allows the mixed skewness increment to be expressed as:

(2.39) $$\begin{eqnarray}S_{\unicode[STIX]{x1D703}}=\frac{\overline{(\unicode[STIX]{x1D6FF}u)(\unicode[STIX]{x1D6FF}\unicode[STIX]{x1D703})^{2}}}{\overline{(\unicode[STIX]{x1D6FF}u)^{2}}^{1/2}\overline{(\unicode[STIX]{x1D6FF}\unicode[STIX]{x1D703})^{2}}}=\frac{\tilde{B}_{\unicode[STIX]{x1D703}}}{\tilde{u} \tilde{\unicode[STIX]{x1D703}}^{2}}\frac{b_{1}(s^{\ast })}{h_{u}(s^{\ast })h_{\unicode[STIX]{x1D703}}(s^{\ast })}=c_{S}\unicode[STIX]{x1D719}_{\unicode[STIX]{x1D703}}(s^{\ast }).\end{eqnarray}$$

The power laws for $\tilde{B}_{\unicode[STIX]{x1D703}}$ , $\tilde{u}$ and $\tilde{\unicode[STIX]{x1D703}}^{2}$ indicate that the scaling function $c_{S}$ is constant. This direct consequence of self-preservation on the jet axis was also observed for the velocity field by DALL. They have shown that $c_{q}$ , the scaling function of the mixed skewness increment of velocity and kinetic energy, was constant.

2.3 Turbulent Péclet number along the jet axis

Another outcome of this analysis pertains to the evolution of the turbulent Péclet number ( $Pe_{\unicode[STIX]{x1D706}}\equiv \sqrt{\overline{u^{2}}}\unicode[STIX]{x1D706}_{\unicode[STIX]{x1D703}}/\unicode[STIX]{x1D6FC}$ , where $\unicode[STIX]{x1D706}_{\unicode[STIX]{x1D703}}\equiv \sqrt{3\unicode[STIX]{x1D6FC}\overline{\unicode[STIX]{x1D703}^{2}}/\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}$ ). Considering that $\overline{\unicode[STIX]{x1D703}^{2}}\propto \hat{x}^{-2}$ and $\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}\propto \hat{x}^{-4}$ , it follows immediately that $\unicode[STIX]{x1D706}_{\unicode[STIX]{x1D703}}\propto \hat{x}$ . Recalling that $\sqrt{\overline{u^{2}}}\propto \hat{x}^{-1}$ , one concludes that the turbulent Péclet number must be constant along the jet axis in the self-preserving region. This behaviour was observed by Darisse et al. (Reference Darisse, Lemay and Benaïssa2015) in their jet where $Pe_{\unicode[STIX]{x1D706}}$ was observed to be constant for $\hat{x}>30$ .

2.4 Characteristic length scales

The self-preservation analyses of the sbs budget revealed that the characteristic length scale of the temperature field evolves linearly, viz. $l_{\unicode[STIX]{x1D703}}\propto \hat{x}$ , as indicated by (2.28). Thus, any length scale evolving linearly along the jet axis would be a relevant normalization length scale, for all scales, including those corresponding to the sufficiently large values of $s$ , where the sbs budget yields the one-point $\overline{\unicode[STIX]{x1D703}^{2}}/2$ budget. It has been shown by Burattini et al. (Reference Burattini, Antonia and Danaila2005b ) and more rigorously by DALL that the Taylor and Kolmogorov microscales, viz. $\unicode[STIX]{x1D706}$ and $\unicode[STIX]{x1D702}$ , can be used interchangeably, as these quantities both evolve linearly along the jet axis. Moreover, it was shown, in the previous section, that $\unicode[STIX]{x1D706}_{\unicode[STIX]{x1D703}}$ also fulfils the requirements ensuing from the present self-preservation analysis. Finally, from § A.2, it is shown that $R_{U}$ and $R_{\unicode[STIX]{x1D6E9}}$ , the half-radii of the mean velocity and temperature profiles, respectively, evolve linearly in the streamwise direction. Thus, $\unicode[STIX]{x1D702}$ , $\unicode[STIX]{x1D706}$ , $\unicode[STIX]{x1D706}_{\unicode[STIX]{x1D703}}$ , $R_{U}$ and $R_{\unicode[STIX]{x1D6E9}}$ are all relevant quantities which could be used as an appropriate characteristic length scale. Hereafter, $R_{U}$ is used (among the other relevant options) as the local normalization length scale.

Figure 2. Second-order structure function of temperature fluctuations on the jet axis, at $x/D=30$ , normalized with Kolmogorov scales $\unicode[STIX]{x1D703}_{K}$ and $\unicode[STIX]{x1D702}$ (flow conditions similar to those of Darisse et al. Reference Darisse, Lemay and Benaïssa2013a ,Reference Darisse, Lemay and Benaïssa b , Reference Darisse, Lemay and Benaïssa2015). Symbols: measurements performed using two cold-wire probes separated by a distance $s$ in the streamwise direction; open symbols: raw data; closed symbols: signals compensated for the attenuation resulting from the wire time constant (see Lemay & Benaïssa Reference Lemay and Benaïssa2001; Lemay, Benaissa & Antonia Reference Lemay, Benaissa and Antonia2003; Darisse et al. Reference Darisse, Lemay and Benaïssa2014). Line: measurements performed using one cold-wire probe (compensated signal) and the assumption of Taylor’s hypothesis ( $s=\unicode[STIX]{x1D70F}U_{0}$ , where $\unicode[STIX]{x1D70F}$ is the time increment).

2.5 Structure functions of temperature fluctuations

Further evidence of complete self-preservation can be inferred from the analysis of the temperature structure functions. Indeed, an indicator showing the fulfilment of complete self-preservation is obtained when the structure functions measured at several positions along the jet axis all collapse over the entire range of the increment $s$ or scales, regardless of which set of scaling variables (complying with self-preservation conditions) is used for the normalization. In the present case, the second- and third-order structure functions for the temperature are considered.

Before considering the collapse of these structure functions along the jet axis, we first consider some technical aspects regarding their measurement, namely the effect of the limited time constant of the cold-wire probe and the validation of Taylor’s hypothesis used to evaluate these functions. For that purpose, figure 2 shows the distribution of $\overline{(\unicode[STIX]{x1D6FF}\unicode[STIX]{x1D703})^{2}}/\unicode[STIX]{x1D703}_{K}^{2}$ (Kolmogorov scaling) measured at a streamwise position $x/D=30$ on the jet axis. At this position, the ratio between the wire cutoff frequency and the Kolmogorov frequency is $f_{c}/f_{K}=0.7$ . The attenuation of the cold-wire signal thus needs to be compensated and this is particularly important at small scales. The compensation of the cold-wire response is performed using the numerical processing technique proposed by Lemay & Benaïssa (Reference Lemay and Benaïssa2001), Lemay et al. (Reference Lemay, Benaissa and Antonia2003). The open and closed symbols in figure 2 show respectively the uncompensated and compensated distributions of $\overline{(\unicode[STIX]{x1D6FF}\unicode[STIX]{x1D703})^{2}}/\unicode[STIX]{x1D703}_{K}^{2}$ obtained from the measurement of two distinct cold-wire probes separated by a distance $s$ along the longitudinal direction. The effect of the compensation procedure is clearly seen at small scales. The full line represents the distribution of compensated $\overline{(\unicode[STIX]{x1D6FF}\unicode[STIX]{x1D703})^{2}}/\unicode[STIX]{x1D703}_{K}^{2}$ inferred from a single cold-wire probe using Taylor’s hypothesis. The time increment $\unicode[STIX]{x1D70F}$ is converted to spatial separation $s$ using a constant convection velocity, constant for all scales, defined by the mean streamwise velocity on the jet axis ( $s=\unicode[STIX]{x1D70F}U_{0}$ ). The agreement between the full line and the closed symbols in figure 2 indicates that the use of Taylor’s hypothesis is thoroughly validated.

Hereinafter, the distributions of temperature structure functions measured at several positions along the jet axis ( $x/D=30$ to 60) are those obtained using Taylor’s hypothesis and compensated signals. For the locations $x/D=30$ , 35, 40, 45, 50, 55 and 60, the ratio of the cutoff to the Kolmogorov frequencies evolves respectively from $f_{c}/f_{K}=0.7$ , 0.9, 1.2, 1.4, 1.7, 2.1 to 2.4. Thus, the use of the compensation procedure is mandatory for $x/D=30$ , but it becomes less important as one moves downstream to $x/D=60$ . Nevertheless, the compensation technique is applied to the temperature signals for all the measurement locations.

Figure 3. Temperature structure functions $\overline{(\unicode[STIX]{x1D6FF}\unicode[STIX]{x1D703})^{2}}$ and $-\overline{(\unicode[STIX]{x1D6FF}\unicode[STIX]{x1D703})^{3}}$ on the jet axis at 7 streamwise locations extending from $x/D$ = 30 to 60, in steps of 5; normalization with $\unicode[STIX]{x1D703}_{K}$ and $\unicode[STIX]{x1D702}$ (top), $\overline{\unicode[STIX]{x1D703}^{2}}$ and $\unicode[STIX]{x1D706}$ (centre), $\unicode[STIX]{x1D6E9}_{0}$ and $R_{U}$ (bottom); flow conditions similar to those of Darisse et al. (Reference Darisse, Lemay and Benaïssa2013a ,Reference Darisse, Lemay and Benaïssa b , Reference Darisse, Lemay and Benaïssa2015).

Figure 3 shows the distribution of the normalized second- and third-order temperature structure functions measured on the jet axis in the range $30\leqslant x/D\leqslant 60$ . Three different sets of scaling variables are used for normalization, viz. Kolmogorov scales ( $\unicode[STIX]{x1D703}_{K}$ and $\unicode[STIX]{x1D702}$ ), Taylor scales ( $\overline{\unicode[STIX]{x1D703}^{2}}$ and $\unicode[STIX]{x1D706}$ ) and large scales ( $\unicode[STIX]{x1D6E9}_{0}$ and $R_{U}$ ). Regardless of the normalization, both the second- and third-order temperature structure functions present a very good collapse over the entire range of scales for all the positions in the range $x/D=30$ to 60. This indicates that complete self-preservation is satisfied for these positions on the jet axis, reinforcing confidence in the analytical development and results presented in § 2.1. Note that despite the moderately large value of $Re_{\unicode[STIX]{x1D706}}$ ( $=548$ ), $\overline{(\unicode[STIX]{x1D6FF}\unicode[STIX]{x1D703})^{2}}$ does not strictly present a power-law variation of the form $r^{\unicode[STIX]{x1D701}_{n}}$ in the scaling range. However, the figure suggests that increasing $Re_{\unicode[STIX]{x1D706}}$ would eventually lead to $\unicode[STIX]{x1D701}_{2}$ approaching the value ( $\unicode[STIX]{x1D701}_{2}=2/3$ ) predicted by Yaglom (Reference Yaglom1949). Note that nothing can be said about $-\overline{(\unicode[STIX]{x1D6FF}\unicode[STIX]{x1D703})^{3}}$ which should not be confused with $-\overline{\unicode[STIX]{x1D6FF}u(\unicode[STIX]{x1D6FF}\unicode[STIX]{x1D703})^{2}}$ for which the exponent $\unicode[STIX]{x1D701}_{3}=1$ if there is an inertial range. If $-\overline{(\unicode[STIX]{x1D6FF}\unicode[STIX]{x1D703})^{3}}$ is assumed to follow a power-law variation $r^{\unicode[STIX]{x1D701}}$ in the inertial range, its power-law exponent is yet to be determined. Interestingly, the data suggest that as $Re_{\unicode[STIX]{x1D706}}$ increases $\unicode[STIX]{x1D701}$ is likely to approach 1, a result which can be also observed in the data of Antonia & Van Atta (Reference Antonia and Van Atta1978) (see their figures 1 and 2).

2.6 One-point $\overline{\unicode[STIX]{x1D703}^{2}}/2$ budget

As previously pointed out, Darisse et al. (Reference Darisse, Lemay and Benaïssa2014) have shown that the one-point $\overline{\unicode[STIX]{x1D703}^{2}}/2$ budget on the jet axis is dominated by convection, production, the radial component of turbulent diffusion and dissipation. Their budget was evaluated in the self-preserving region since they showed that both $Re_{\unicode[STIX]{x1D706}}$ and $Pe_{\unicode[STIX]{x1D706}}$ were constant along the jet axis. Retaining only these predominant terms and recalling that outer scales $U_{0}$ , $\unicode[STIX]{x1D6E9}_{0}$ and $R_{U}$ are relevant self-preservation scales, the normalized one-point budget on the jet axis is written as (see Darisse et al. Reference Darisse, Lemay and Benaïssa2015, for details):

(2.40) $$\begin{eqnarray}\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}^{\ast }=\frac{\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}R_{U}}{U_{0}\unicode[STIX]{x1D6E9}_{0}^{2}}\simeq B_{R_{U}}\frac{\overline{\unicode[STIX]{x1D703}^{2}}}{\unicode[STIX]{x1D6E9}_{0}^{2}}+B_{R_{U}}\frac{\overline{u\unicode[STIX]{x1D703}}}{U_{0}\unicode[STIX]{x1D6E9}_{0}}-\frac{\text{d}}{\text{d}\unicode[STIX]{x1D709}}\left(\frac{\overline{v\unicode[STIX]{x1D703}^{2}}}{U_{0}\unicode[STIX]{x1D6E9}_{0}^{2}}\right)\simeq C_{\unicode[STIX]{x1D703}}^{\ast }+P_{\unicode[STIX]{x1D703}}^{\ast }+D_{\unicode[STIX]{x1D703}}^{\ast },\end{eqnarray}$$

where $\unicode[STIX]{x1D709}=r/R_{U}$ is the normalized radial coordinate. The power laws for $\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}$ , $R_{U}$ , $U_{0}$ and $\unicode[STIX]{x1D6E9}_{0}$ allow the dissipation term to be written as

(2.41) $$\begin{eqnarray}\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}^{\ast }=\frac{A_{\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}B_{R_{U}}}{B_{U}B_{\unicode[STIX]{x1D6E9}}^{2}}.\end{eqnarray}$$

Introducing this expression and the definition of $B_{I}$ (2.36) into the normalized budget (2.40) and solving for $A_{\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}$ yields

(2.42) $$\begin{eqnarray}A_{\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}\simeq B_{U}B_{\unicode[STIX]{x1D6E9}}^{2}\left[B_{I}^{2}+\frac{\overline{u\unicode[STIX]{x1D703}}}{U_{0}\unicode[STIX]{x1D6E9}_{0}}-\frac{1}{B_{R_{U}}}\frac{\text{d}}{\text{d}\unicode[STIX]{x1D709}}\left(\frac{\overline{v\unicode[STIX]{x1D703}^{2}}}{U_{0}\unicode[STIX]{x1D6E9}_{0}^{2}}\right)\right].\end{eqnarray}$$

This expression for the prefactor of the $\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}$ power law (2.34) is more involved than its counterpart for $\unicode[STIX]{x1D716}_{k}$ , i.e. equation (1.2). The two mixed velocity–temperature moments accounting for the production and turbulent diffusion terms of the $\overline{\unicode[STIX]{x1D703}^{2}}/2$ budget add a level of complexity rendering this relation less attractive, from a modelling point of view, than the expression for $A_{\unicode[STIX]{x1D716}_{k}}$ . In the following section, we propose preliminary models for the longitudinal evolution of $\overline{u\unicode[STIX]{x1D703}}$ and $\overline{v\unicode[STIX]{x1D703}^{2}}$ along the jet axis. This will help us develop an ‘empirical’ expression for the prefactor $A_{\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}$ that will account for the effect of the jet boundary conditions and that can be estimated from relatively simple measurements.

3 Modelling the mixed velocity–temperature moments

3.1 Longitudinal velocity–temperature correlation model

The correlation $\overline{u\unicode[STIX]{x1D703}}$ is modelled by assuming weak anisotropy of the temperature field in the self-preserving region of the jet. This hypothesis applied to the scalar flux underpins the development of the well-known algebraic second moment closure for the scalar transport (see for example Gibson & Launder Reference Gibson and Launder1976; Hanjalić & Launder Reference Hanjalić and Launder2011). In the present case, this means that the ratio $\overline{u\unicode[STIX]{x1D703}}/\sqrt{k\overline{\unicode[STIX]{x1D703}^{2}}}$ is expected to be constant. The approach can be simplified further by considering that the correlation coefficient $\unicode[STIX]{x1D70C}_{u\unicode[STIX]{x1D703}}=\overline{u\unicode[STIX]{x1D703}}/\sqrt{\overline{u^{2}}\;\overline{\unicode[STIX]{x1D703}^{2}}}$ is constant across the jet width. The data reported by Darisse et al. (Reference Darisse, Lemay and Benaïssa2015) ( $x/D=30$ ) and Chevray & Tutu (Reference Chevray and Tutu1978) ( $x/D=15$ ) indicate that $\unicode[STIX]{x1D70C}_{u\unicode[STIX]{x1D703}}(r)$ does not vary significantly ( $0.46<\unicode[STIX]{x1D70C}_{u\unicode[STIX]{x1D703}}<0.51$ ) around the centreline. Pietri, Amielh & Anselmet (Reference Pietri, Amielh and Anselmet2000) made the same observation in a coflowing jet exiting from a pipe. They also showed that, along the centreline, in the region $15\geqslant x/D\geqslant 28$ , the value of $c_{u\unicode[STIX]{x1D703}}$ ( $\equiv \unicode[STIX]{x1D70C}_{u\unicode[STIX]{x1D703}}$ at $r=0$ ) was almost constant. More precisely, $c_{u\unicode[STIX]{x1D703}}$ decreased slightly between 0.52 at $x/D=15$ and 0.49 at $x/D=28$ . Their measurements are consistent with the value of 0.48 found here. Using a combination of laser-induced fluorescence and laser Doppler anemometry (LDA) techniques in a slightly heated water jet, Lemoine et al. (Reference Lemoine, Antoine, Wolff and Lebouche1999) reported a measured value of $c_{u\unicode[STIX]{x1D703}}=0.48$ , which also confirms the validity of the present value. Finally, additional support is obtained with the results presented by Panchapakesan & Lumley (Reference Panchapakesan and Lumley1993b ) in a helium jet. From their reported longitudinal evolution of $\sqrt{\overline{u^{2}}}$ on the centreline, we can extrapolate this quantity up to the far non-buoyant region and find that, for this flow configuration too, $c_{u\unicode[STIX]{x1D703}}=0.48$ .

Based on these reported observations, the following model is adopted on the jet axis:

(3.1) $$\begin{eqnarray}\overline{u\unicode[STIX]{x1D703}}=c_{u\unicode[STIX]{x1D703}}\sqrt{\overline{u^{2}}\overline{\unicode[STIX]{x1D703}^{2}}}\quad \text{with }c_{u\unicode[STIX]{x1D703}}=\unicode[STIX]{x1D70C}_{u\unicode[STIX]{x1D703}}(r=0)=0.48.\end{eqnarray}$$

This expression is also a direct consequence of the self-preservation analysis performed on the sbs budget in § 2.1. It is shown that, along the jet axis, $\overline{u\unicode[STIX]{x1D703}}\propto \hat{x}^{-2}$ (see also expression (A 11)), $\sqrt{\overline{u^{2}}}\propto \hat{x}^{-1}$ and $\sqrt{\overline{\unicode[STIX]{x1D703}^{2}}}\propto \hat{x}^{-1}$ ; it immediately follows that $\unicode[STIX]{x1D70C}_{u\unicode[STIX]{x1D703}}$ is constant along the jet axis. Darisse et al. (Reference Darisse, Lemay and Benaïssa2015) showed that $Re_{\unicode[STIX]{x1D706}}$ and $Pe_{\unicode[STIX]{x1D706}}$ are constant along the jet axis, and concluded that complete self-preservation for the velocity and the scalar fields was satisfied. Thus, based on their measurements, $c_{u\unicode[STIX]{x1D703}}=0.48$ is adopted in (3.1) for $\overline{u\unicode[STIX]{x1D703}}$ . Introducing the previously defined coefficients terms $A_{I}$ and $B_{I}$ , the model for the normalized mixed correlation is written as

(3.2) $$\begin{eqnarray}\frac{\overline{u\unicode[STIX]{x1D703}}}{U_{0}\unicode[STIX]{x1D6E9}_{0}}=c_{u\unicode[STIX]{x1D703}}A_{I}B_{I}\quad \text{with }c_{u\unicode[STIX]{x1D703}}=0.48.\end{eqnarray}$$

3.2 Turbulent diffusion model

Using the transport equation of $\overline{u_{i}\unicode[STIX]{x1D703}^{2}}$ in which the convective transport was neglected, Dekeyser & Launder (Reference Dekeyser, Launder, Bradbury, Durst, Launder, Schmidt and Whitelaw1983) proposed a model based on a Gaussian approximation for the fourth-order moments and a return-to-isotropy representation for pressure interactions:

(3.3) $$\begin{eqnarray}\overline{u_{i}\unicode[STIX]{x1D703}^{2}}=-c_{\unicode[STIX]{x1D703}}\frac{k}{\unicode[STIX]{x1D716}_{k}}\left(\overline{u_{i}u_{j}}\frac{\unicode[STIX]{x2202}\overline{\unicode[STIX]{x1D703}^{2}}}{\unicode[STIX]{x2202}x_{j}}+2\overline{u_{i}\unicode[STIX]{x1D703}}\frac{\unicode[STIX]{x2202}\overline{u_{j}\unicode[STIX]{x1D703}}}{\unicode[STIX]{x2202}x_{j}}\right)\quad \text{with }c_{\unicode[STIX]{x1D703}}=0.11.\end{eqnarray}$$

When the last term on the right-hand side of relation (3.3) is omitted, the truncated expression represents the simple gradient-type model widely used in numerical calculations. As reported by Dekeyser & Launder (Reference Dekeyser, Launder, Bradbury, Durst, Launder, Schmidt and Whitelaw1983), this truncation is usually made to facilitate numerical solutions. They, however, mentioned that the complete form was in better agreement with their measurements. In the present case, we are interested in the radial component of (3.3) which is explicitly expressed as

(3.4) $$\begin{eqnarray}\overline{v\unicode[STIX]{x1D703}^{2}}=-c_{\unicode[STIX]{x1D703}}\frac{k}{\unicode[STIX]{x1D716}_{k}}\left[\overline{uv}\frac{\unicode[STIX]{x2202}\overline{\unicode[STIX]{x1D703}^{2}}}{\unicode[STIX]{x2202}x}+\overline{v^{2}}\frac{\unicode[STIX]{x2202}\overline{\unicode[STIX]{x1D703}^{2}}}{\unicode[STIX]{x2202}r}+2\overline{v\unicode[STIX]{x1D703}}\left(\frac{\unicode[STIX]{x2202}\overline{u\unicode[STIX]{x1D703}}}{\unicode[STIX]{x2202}x}+\frac{\unicode[STIX]{x2202}\overline{v\unicode[STIX]{x1D703}}}{\unicode[STIX]{x2202}r}\right)\right].\end{eqnarray}$$

The terms involving longitudinal derivatives are relatively small compared to the two other terms within the brackets, particularly in the mid and far fields. The measurements reported by Darisse et al. (Reference Darisse, Lemay and Benaïssa2015) indicate that the longitudinal derivative terms represent approximately $-16\,\%$ of the radial derivative contributions. To simplify the model, the longitudinal derivative terms are omitted and, to compensate for the effect of this omission, the value of $c_{\unicode[STIX]{x1D703}}$ is reduced by 16 % to 0.095. This yields a simplified model:

(3.5) $$\begin{eqnarray}\overline{v\unicode[STIX]{x1D703}^{2}}\simeq -c_{\unicode[STIX]{x1D703}}\frac{k}{\unicode[STIX]{x1D716}_{k}}\left[\overline{v^{2}}\frac{\unicode[STIX]{x2202}\overline{\unicode[STIX]{x1D703}^{2}}}{\unicode[STIX]{x2202}r}+2\overline{v\unicode[STIX]{x1D703}}\frac{\unicode[STIX]{x2202}\overline{v\unicode[STIX]{x1D703}}}{\unicode[STIX]{x2202}r}\right].\end{eqnarray}$$

Figure 4. Radial distributions of normalized velocity–temperature moments near the jet axis; right side, correlation coefficient $\unicode[STIX]{x1D70C}_{u\unicode[STIX]{x1D703}}$ : Darisse et al. (Reference Darisse, Lemay and Benaïssa2015) (black diamonds), Chevray & Tutu (Reference Chevray and Tutu1978) (blue squares) and model defined by (3.1) (red dashed line); left scale, $\overline{v\unicode[STIX]{x1D703}^{2}}/(U_{0}\unicode[STIX]{x1D6E9}_{0}^{2})$ : Darisse et al. (Reference Darisse, Lemay and Benaïssa2015) (black circles) with best fit (black line), model of Dekeyser & Launder (Reference Dekeyser, Launder, Bradbury, Durst, Launder, Schmidt and Whitelaw1983), equation (3.4), (green line) and simplified model defined by (3.5) (red dashed line).

The database of Darisse et al. (Reference Darisse, Lemay and Benaïssa2015) is used here to test the validity of these models on the jet axis. All the radial distributions of the measured quantities reported by these authors have been fitted with least-square regressions using the mathematical expressions given in Hussein, Capp & George (Reference Hussein, Capp and George1994). Near the jet axis ( $r/R_{U}\leqslant 0.5$ ), figure 4 indicates that both, relations (3.4) and (3.5) are in good agreement with the measurements of $\overline{v\unicode[STIX]{x1D703}^{2}}/(U_{0}\unicode[STIX]{x1D6E9}_{0}^{2})$ . Moreover, very close to the centreline (around $r/R_{U}\leqslant 0.2$ ), simplified and Dekeyser–Launder models are almost perfectly equivalent. As the present development focuses on the evolution of different quantities on the centreline, the simplified model (3.5) is used. The radial derivative of $\overline{v\unicode[STIX]{x1D703}^{2}}$ , which represents the dominant part of the turbulent diffusion of $\overline{\unicode[STIX]{x1D703}^{2}}/2$ on the jet axis (where $\overline{v\unicode[STIX]{x1D703}}\equiv 0$ and $\unicode[STIX]{x2202}\overline{\unicode[STIX]{x1D703}^{2}}/\unicode[STIX]{x2202}r\equiv 0$ on the axis by virtue of axisymmetry), is written as

(3.6) $$\begin{eqnarray}\frac{\unicode[STIX]{x2202}\overline{v\unicode[STIX]{x1D703}^{2}}}{\unicode[STIX]{x2202}r}\simeq -c_{\unicode[STIX]{x1D703}}\frac{k}{\unicode[STIX]{x1D716}_{k}}\left[\overline{v^{2}}\frac{\unicode[STIX]{x2202}^{2}\overline{\unicode[STIX]{x1D703}^{2}}}{\unicode[STIX]{x2202}r^{2}}+2\left(\frac{\unicode[STIX]{x2202}\overline{v\unicode[STIX]{x1D703}}}{\unicode[STIX]{x2202}r}\right)^{2}\right].\end{eqnarray}$$

The data of Darisse et al. (Reference Darisse, Lemay and Benaïssa2015) reveal that the two terms within brackets are approximately equal. Thus, we can write (3.6) as follows

(3.7) $$\begin{eqnarray}\frac{\unicode[STIX]{x2202}\overline{v\unicode[STIX]{x1D703}^{2}}}{\unicode[STIX]{x2202}r}\simeq -4c_{\unicode[STIX]{x1D703}}\frac{k}{\unicode[STIX]{x1D716}_{k}}\left(\frac{\unicode[STIX]{x2202}\overline{v\unicode[STIX]{x1D703}}}{\unicode[STIX]{x2202}r}\right)^{2}.\end{eqnarray}$$

The radial gradient of $\overline{v\unicode[STIX]{x1D703}}$ on the jet axis can be obtained directly from the mean temperature equation, after neglecting $\unicode[STIX]{x2202}\overline{u\unicode[STIX]{x1D703}}/\unicode[STIX]{x2202}x$ and the molecular diffusion term, viz.

(3.8) $$\begin{eqnarray}\frac{\unicode[STIX]{x2202}\overline{v\unicode[STIX]{x1D703}}}{\unicode[STIX]{x2202}r}\simeq -\frac{1}{2}U_{0}\frac{\text{d}\unicode[STIX]{x1D6E9}_{0}}{\text{d}x}.\end{eqnarray}$$

The combination of (3.7) and (3.8) yields, on the jet axis, a model for the turbulent diffusion of $\overline{\unicode[STIX]{x1D703}^{2}}/2$

(3.9) $$\begin{eqnarray}\frac{\unicode[STIX]{x2202}\overline{v\unicode[STIX]{x1D703}^{2}}}{\unicode[STIX]{x2202}r}\simeq -c_{\unicode[STIX]{x1D703}}\frac{k}{\unicode[STIX]{x1D716}_{k}}U_{0}^{2}\left(\frac{\text{d}\unicode[STIX]{x1D6E9}_{0}}{\text{d}x}\right)^{2}\quad \text{with }c_{\unicode[STIX]{x1D703}}=0.095.\end{eqnarray}$$

Using relation (A 17) for the streamwise gradient of $\unicode[STIX]{x1D6E9}_{0}$ , the normalized form of this model is

(3.10) $$\begin{eqnarray}\frac{\text{d}}{\text{d}\unicode[STIX]{x1D709}}\left(\frac{\overline{v\unicode[STIX]{x1D703}^{2}}}{U_{0}\unicode[STIX]{x1D6E9}_{0}^{2}}\right)\simeq -c_{\unicode[STIX]{x1D703}}B_{R_{U}}^{2}\frac{k}{U_{0}^{2}}\frac{U_{0}^{3}}{\unicode[STIX]{x1D716}_{k}R_{U}}.\end{eqnarray}$$

Introducing the notation $k/U_{0}^{2}=A_{I}^{2}({\mathcal{R}}+1/2)$ and relation (1.3), which represents $\unicode[STIX]{x1D716}_{k}^{\ast }$ , yields the final form of the normalized turbulent diffusion model:

(3.11) $$\begin{eqnarray}\frac{\text{d}}{\text{d}\unicode[STIX]{x1D709}}\left(\frac{\overline{v\unicode[STIX]{x1D703}^{2}}}{U_{0}\unicode[STIX]{x1D6E9}_{0}^{2}}\right)\simeq -c_{\unicode[STIX]{x1D703}}B_{R_{U}}\frac{({\mathcal{R}}+1/2)}{({\mathcal{R}}+2)}\quad \text{with }c_{\unicode[STIX]{x1D703}}=0.095.\end{eqnarray}$$

4 Outcomes related to the dissipation

4.1 Mean dissipation rate of $\overline{\unicode[STIX]{x1D703}^{2}}/2$

The models for the mixed correlation (3.2) and the turbulent diffusion (3.11) introduced into relation (2.42) allow an empirical expression for the prefactor $A_{\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}$ to be obtained:

(4.1) $$\begin{eqnarray}A_{\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}\simeq B_{U}B_{\unicode[STIX]{x1D703}}^{2}\left[B_{I}^{2}+c_{u\unicode[STIX]{x1D703}}A_{I}B_{I}+c_{\unicode[STIX]{x1D703}}\left(\frac{{\mathcal{R}}+1/2}{{\mathcal{R}}+2}\right)\right],\end{eqnarray}$$

with $c_{u\unicode[STIX]{x1D703}}=0.48$ , $c_{\unicode[STIX]{x1D703}}=0.095$ . Using the data of Darisse et al. (Reference Darisse, Lemay and Benaïssa2015) expression (4.1) yields $A_{\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}=15.02$ . This is in very good agreement with the value $A_{\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}=14.58$ obtained from the measurements presented in figure 1. It is important to note that the coefficient $A_{\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}=14.58$ was obtained from the power law of $\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}$ shown in figure 1. The values of $\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}$ were estimated from the integral of the dissipation spectra of temperature fluctuations, under the assumption of local isotropy and Taylor’s hypothesis. It was shown by Darisse et al. (Reference Darisse, Lemay and Benaïssa2015) that, at $\hat{x}=30$ , the isotropic estimate of $\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}$ on the jet centreline was in very good agreement with the value inferred from the balance of the temperature budget. Therefore, this value of 14.58 was evaluated from a dataset different than the one used to evaluate $A_{\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}=15.02$ with expression (4.1). This guarantees that the validation of the empirical expression is made using an independent test.

Table 2. Estimates obtained from the simple empirical expressions compared with the results presented by Darisse et al. (Reference Darisse, Lemay and Benaïssa2015) (DLB, measurements in a heated jet), Antonia & Mi (Reference Antonia and Mi1993) (AM, measurements in a heated jet), Panchapakesan & Lumley (Reference Panchapakesan and Lumley1993b ) (PL, measurements in the non-buoyant region of a helium jet) and Ruffin et al. (Reference Ruffin, Schiestel, Anselmet, Amielh and Fulachier1994) (RAL, simulations using second-order turbulence models with a passive scalar).

After inserting (4.1) into (2.41), $\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}$ normalized with local outer scales, $R_{U}$ , $U_{0}$ and $\unicode[STIX]{x1D6E9}_{0}$ , can be expressed as:

(4.2) $$\begin{eqnarray}\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}^{\ast }\simeq B_{R_{U}}\left[B_{I}^{2}+c_{u\unicode[STIX]{x1D703}}A_{I}B_{I}+c_{\unicode[STIX]{x1D703}}\left(\frac{{\mathcal{R}}+1/2}{{\mathcal{R}}+2}\right)\right].\end{eqnarray}$$

By virtue of (2.40), expression (4.2) also defines a simple empirical expression for the dissipation of $\overline{\unicode[STIX]{x1D703}^{2}}/2$ on the jet axis. The validity of this empirical expression for $\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}^{\ast }$ has been tested using four datasets (see table 2). We have to point out that very few datasets in the literature provide measurements for $\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}^{\ast }$ in a well-established self-similar region of a slightly heated turbulent round jet. As already mentioned, Darisse et al. (Reference Darisse, Lemay and Benaïssa2015) reported such measurements (inferred from the $\overline{\unicode[STIX]{x1D703}^{2}}/2$ budget and locally isotropic estimates). In their case, the outcome of the empirical expression (4.2) differs by only $-0.3\,\%$ from their measurement. Antonia & Mi (Reference Antonia and Mi1993) presented direct measurements and locally isotropic estimates for $\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}^{\ast }$ (which were in agreement on the jet centreline) in a slightly heated round jet in the region $\hat{x}=15$ . The outcome of relation (4.2) differs by $+5\,\%$ from their measurement. This is in rather good agreement considering that, in their case, self-similarity is probably not completely achieved at $\hat{x}=15$ . The third dataset was provided by Panchapakesan & Lumley (Reference Panchapakesan and Lumley1993b ) which reported measurements (inferred from the budget of the variance of concentration fluctuations) in the non-buoyant region of a helium round jet. Considering that this flow field shows some analogies with the present case, it is interesting to observe that the outcome of relation (4.2) differs by only $+5.1\,\%$ from their measurement. Finally, the Reynolds-averaged Navier–Stokes (RANS) simulations reported by Ruffin et al. (Reference Ruffin, Schiestel, Anselmet, Amielh and Fulachier1994) (air jet with a passive contaminant) were used to provide a computational fluid dynamics (CFD) test case. The value of  $\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}^{\ast }$ calculated with expression (4.2) differs by $-12.5\,\%$ from their simulation on the jet centreline. Considering that RANS simulations of turbulent round jets are usually not in perfect concordance with the measurements (as reported by these authors), the present level of agreement is rather satisfactory.

4.2 Turbulent Péclet number

It was shown in § 2.3 that the turbulent Péclet number $Pe_{\unicode[STIX]{x1D706}}$ , which can be expressed as

(4.3) $$\begin{eqnarray}Pe_{\unicode[STIX]{x1D706}}=\sqrt{\frac{3\overline{u^{2}}\overline{\unicode[STIX]{x1D703}^{2}}}{\unicode[STIX]{x1D6FC}\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}},\end{eqnarray}$$

is constant along the jet axis. The power laws for $\overline{u^{2}}$ , $\overline{\unicode[STIX]{x1D703}^{2}}$ and $\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}$ are now introduced in order to express $Pe_{\unicode[STIX]{x1D706}}$ as follows

(4.4) $$\begin{eqnarray}Pe_{\unicode[STIX]{x1D706}}=A_{I}B_{I}B_{U}B_{\unicode[STIX]{x1D6E9}}\sqrt{\frac{3Re_{D}Pr}{A_{\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}}}},\end{eqnarray}$$

where $Pr\equiv \unicode[STIX]{x1D6FC}/\unicode[STIX]{x1D708}$ , the Prandtl number, is considered to be equal to 0.71. Using (4.1), the final expression for $Pe_{\unicode[STIX]{x1D706}}$ is

(4.5) $$\begin{eqnarray}Pe_{\unicode[STIX]{x1D706}}=A_{Pe}\sqrt{Re_{D}Pr}=\left[A_{I}B_{I}\sqrt{\frac{3B_{U}}{B_{I}^{2}+c_{u\unicode[STIX]{x1D703}}A_{I}B_{I}+c_{\unicode[STIX]{x1D703}}{\displaystyle \frac{({\mathcal{R}}+1/2)}{({\mathcal{R}}+2)}}}}\right]\sqrt{Re_{D}Pr}.\end{eqnarray}$$

Since the prefactor $A_{Pe}$ (term in brackets), depends on $B_{U}$ , $A_{I}$ , $B_{I}$ and ${\mathcal{R}}$ , a set of parameters influenced by the jet boundary conditions, its value is likely to depend on the initial/boundary conditions. However, equation (4.5) gives $A_{Pe}=0.66$ and 0.71 for the Darisse et al. (Reference Darisse, Lemay and Benaïssa2015) and Antonia & Mi (Reference Antonia and Mi1993) data respectively. It appears that the boundary conditions do not severely impact this parameter. Thus, it seems reasonable to use an average – and constant – value of $A_{Pe}=0.68$ . With this assumption, relation (4.5) yields $Pe_{\unicode[STIX]{x1D706}}=211$ for Darisse et al. (Reference Darisse, Lemay and Benaïssa2015) and 79 for Antonia & Mi (Reference Antonia and Mi1993), which results are in good agreement with the measurements.

4.3 Thermal-to-mechanical time scale ratio

The thermal-to-mechanical time scale ratio is defined as

(4.6) $$\begin{eqnarray}R=\frac{\overline{\unicode[STIX]{x1D703}^{2}}/(2\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}})}{k/\unicode[STIX]{x1D716}_{k}}.\end{eqnarray}$$

Using the different power laws derived here for the physical quantities involved in this expression allows the following simple model on the jet axis to be written as

(4.7) $$\begin{eqnarray}R\simeq \frac{B_{I}^{2}({\mathcal{R}}+2)}{\left(B_{I}^{2}+c_{u\unicode[STIX]{x1D703}}A_{I}B_{I}+c_{\unicode[STIX]{x1D703}}{\displaystyle \frac{({\mathcal{R}}+1/2)}{({\mathcal{R}}+2)}}\right)(2{\mathcal{R}}+1)}.\end{eqnarray}$$

The value reported by Darisse et al. (Reference Darisse, Lemay and Benaïssa2015), which is inferred from the budgets of $k$ and $\overline{\unicode[STIX]{x1D703}^{2}}/2$ measured in the same flow conditions, was $R=0.41$ on the jet axis. Introducing the parameters listed in table 2 for Darisse et al. (Reference Darisse, Lemay and Benaïssa2015), the simple model (4.7) gives $R\simeq 0.40$ , which is in good agreement with the measurements.

5 Conclusions

The paper reports self-preservation analyses of the transport equations for the mean temperature and the second-order temperature structure function with a view to deriving a simple analytical expression for $\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}$ on the axis of a slightly heated round jet. Comparison between the theoretical results and experimental data shows that complete self-preservation (i.e. self-preservation at all scales of motion) is observed on the axis of the jet. In particular, the condition for complete self-preservation, which indicates that all the characteristic length scales must evolve linearly with $x$ and are proportional to each other, viz. $\unicode[STIX]{x1D702}\propto \unicode[STIX]{x1D706}\propto \unicode[STIX]{x1D706}_{\unicode[STIX]{x1D703}}\propto R_{U}\propto R_{\unicode[STIX]{x1D6E9}}\propto \hat{x}$ , is very well verified by the measurements, providing confidence in the theoretical analysis. Also confirmed by the measurements, the analysis shows that both $Re_{\unicode[STIX]{x1D706}}$ and $Pe_{\unicode[STIX]{x1D706}}$ , the turbulent Reynolds and Péclet numbers, must be constant under complete self-preservation. This is of practical interest since then they can be formally used as parameters to confirm the validity of self-preservation along the jet axis. Finally, the mean temperature excess and the temperature variance must evolve like $\unicode[STIX]{x1D6E9}_{0}\propto \hat{x}^{-1}$ and $\overline{\unicode[STIX]{x1D703}^{2}}\propto \hat{x}^{-2}$ while the mixed velocity–temperature moments vary like $\overline{u\unicode[STIX]{x1D703}}\propto \hat{x}^{-2}$ and $\overline{v\unicode[STIX]{x1D703}^{2}}\propto \hat{x}^{-3}$ .

The self-preservation analysis formally yields $\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}\propto \hat{x}^{-4}$ , thus providing a solid analytical foundation for the derivation of the $-4$ th power law which characterizes the evolution along the jet axis of the mean dissipation rate of $\overline{\unicode[STIX]{x1D703}^{2}}/2$ . The present results support and confirm the experimental observation of the $-4$ th power law. The analysis also leads to the following two important outcomes:

  1. (i) A new analytical expression for $\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}$ on the jet axis was developed and shown to be in a good agreement with experimental data.

  2. (ii) New simple models are also developed for the turbulent Péclet number and the thermal-to-mechanical time scale ratio. Although the models are in reasonably good agreement with measurements, it should be kept in mind that they involve global parameters which are influenced by the boundary conditions and are therefore flow dependent.

Acknowledgements

J.L. and A.B. gratefully acknowledge the financial support of the Natural Sciences and Engineering Research Council of Canada. L.D. gratefully acknowledges the financial support of the Australian Research Council.

Appendix A. Derivation of some additional power laws

A.1 Evolution of the mean temperature excess along the jet axis

The streamwise evolution of $\unicode[STIX]{x1D6E9}_{0}$ is determined from the equilibrium self-preservation analysis of the mean temperature equation. Using the continuity equation, mean axisymmetry properties and the condition of high Reynolds and Péclet numbers (negligible contribution of molecular diffusion), the mean temperature equation for the slightly heated round jet can be written as

(A 1) $$\begin{eqnarray}\frac{\unicode[STIX]{x2202}\overline{U}\overline{\unicode[STIX]{x1D6E9}}}{\unicode[STIX]{x2202}x}+\frac{1}{r}\frac{\unicode[STIX]{x2202}r\overline{V}\overline{\unicode[STIX]{x1D6E9}}}{\unicode[STIX]{x2202}r}+\frac{\unicode[STIX]{x2202}\overline{u\unicode[STIX]{x1D703}}}{\unicode[STIX]{x2202}x}+\frac{1}{r}\frac{\unicode[STIX]{x2202}r\overline{v\unicode[STIX]{x1D703}}}{\unicode[STIX]{x2202}r}\simeq 0.\end{eqnarray}$$

The data of Darisse et al. (Reference Darisse, Lemay and Benaïssa2014, Reference Darisse, Lemay and Benaïssa2015) ( $Re_{D}=1.4\times 10^{5}$ and $Re_{\unicode[STIX]{x1D706}}=548$ ) show that the molecular diffusion terms are negligible – by several orders of magnitude – when compared to the dominant terms. Integration of this expression in the radial direction yields, after some manipulations,

(A 2) $$\begin{eqnarray}\frac{\unicode[STIX]{x2202}}{\unicode[STIX]{x2202}x}\int _{0}^{\infty }\left(\overline{U}\overline{\unicode[STIX]{x1D6E9}}+\overline{u\unicode[STIX]{x1D703}}\right)r\,\text{d}r+\left.\left(r\overline{V}\overline{\unicode[STIX]{x1D6E9}}\right)\right|_{0}^{\infty }+\left.\left(r\overline{v\unicode[STIX]{x1D703}}\right)\right|_{0}^{\infty }\simeq 0.\end{eqnarray}$$

Since the last two terms on the left-hand side are zero, equation (A 2) reduces to

(A 3) $$\begin{eqnarray}\int _{0}^{\infty }\left(\overline{U}\overline{\unicode[STIX]{x1D6E9}}+\overline{u\unicode[STIX]{x1D703}}\right)r\,\text{d}r\simeq C_{9},\end{eqnarray}$$

where $C_{9}$ is constant. This expression represents the integral form of the energy conservation applied to the slightly heated round jet.

In a similar fashion to that reported for the velocity field by Hussein et al. (Reference Hussein, Capp and George1994) and Ewing et al. (Reference Ewing, Frohnapfel, George, Pedersen and Westerweel2007), we assume that the mean temperature equation admits equilibrium self-preservation solution for $\overline{U}$ , $\overline{\unicode[STIX]{x1D6E9}}$ and $\overline{u\unicode[STIX]{x1D703}}$ of the form

(A 4) $$\begin{eqnarray}\displaystyle & \displaystyle \overline{U}=U_{0}(x)f(r/l_{\unicode[STIX]{x1D703}}), & \displaystyle\end{eqnarray}$$
(A 5) $$\begin{eqnarray}\displaystyle & \displaystyle \overline{\unicode[STIX]{x1D6E9}}=\unicode[STIX]{x1D6E9}_{0}(x)g(r/l_{\unicode[STIX]{x1D703}}), & \displaystyle\end{eqnarray}$$
(A 6) $$\begin{eqnarray}\displaystyle & \displaystyle \overline{u\unicode[STIX]{x1D703}}=H_{0}(x)h(r/l_{\unicode[STIX]{x1D703}}), & \displaystyle\end{eqnarray}$$

where $l_{\unicode[STIX]{x1D703}}$ is a characteristic length scale introduced in § 2.1. For the velocity field, Thiesset et al. (Reference Thiesset, Antonia and Djenidi2014) showed that the product $U_{0}l_{u}$ is constant, $l_{u}$ being the characteristic length scale. For the passive temperature field, it is shown in § 2.1 that the product $U_{0}l_{\unicode[STIX]{x1D703}}$ is constant. Thus, as the observation made for the velocity field still applied to the passive temperature field, it immediately follows that $l_{\unicode[STIX]{x1D703}}\propto l_{u}$ . Consequently, the shape function $f(r/l_{\unicode[STIX]{x1D703}})$ in (A 4) can also be considered as a function of $r/l_{u}$ , as it is usually done for the mean streamwise velocity profile. The terms in capital letters on the right-hand sides of expressions (A 4)–(A 6) represent scales characterizing the streamwise evolution of the given quantity on the jet axis, while the lower-case functions represent the shape of the radial profile of these quantities. Introducing these forms (in which we omit to note the $x$ and $r/l_{\unicode[STIX]{x1D703}}$ dependency for scale and shape functions, respectively) into (A 3) yields

(A 7) $$\begin{eqnarray}(U_{0}\unicode[STIX]{x1D6E9}_{0}l_{\unicode[STIX]{x1D703}}^{2}+H_{0}l_{\unicode[STIX]{x1D703}}^{2})\int _{0}^{\infty }(fg+h)(r/l_{\unicode[STIX]{x1D703}})\,\text{d}(r/l_{\unicode[STIX]{x1D703}})\simeq C_{9}.\end{eqnarray}$$

The term in front of the integral involves functions of $x$ only while the integral involves functions of $r/l_{\unicode[STIX]{x1D703}}$ only. The product of the two terms being constant, each of them must be constant, yielding the two following constraints:

(A 8) $$\begin{eqnarray}\displaystyle & \displaystyle U_{0}\unicode[STIX]{x1D6E9}_{0}l_{\unicode[STIX]{x1D703}}^{2}\simeq C_{10}, & \displaystyle\end{eqnarray}$$
(A 9) $$\begin{eqnarray}\displaystyle & \displaystyle H_{0}l_{\unicode[STIX]{x1D703}}^{2}\simeq C_{11}, & \displaystyle\end{eqnarray}$$

where $C_{10}$ and $C_{11}$ are both constant. Finally, as is shown in § 2.1 that $U_{0}\propto \hat{x}^{-1}$ and $l_{\unicode[STIX]{x1D703}}\propto \hat{x}$ , the power laws for $\unicode[STIX]{x1D6E9}_{0}$ and $H_{0}$ immediately follow:

(A 10) $$\begin{eqnarray}\displaystyle & \displaystyle \unicode[STIX]{x1D6E9}_{0}\propto \hat{x}^{-1}, & \displaystyle\end{eqnarray}$$
(A 11) $$\begin{eqnarray}\displaystyle & \displaystyle H_{0}\propto \hat{x}^{-2}. & \displaystyle\end{eqnarray}$$

The power law for $\unicode[STIX]{x1D6E9}_{0}$ is usually written as

(A 12) $$\begin{eqnarray}\frac{\unicode[STIX]{x1D6E9}_{0}}{\unicode[STIX]{x1D6E9}_{j}}=B_{\unicode[STIX]{x1D6E9}}\hat{x}^{-1}.\end{eqnarray}$$

The streamwise derivative of this power law gives

(A 13) $$\begin{eqnarray}\frac{\text{d}\unicode[STIX]{x1D6E9}_{0}}{\text{d}\hat{x}}=-\unicode[STIX]{x1D6E9}_{0}\hat{x}^{-1}.\end{eqnarray}$$

A.2 Longitudinal evolution of the jet half-radius

Admitting a self-preserving solution for the mean temperature excess as described by (A 5), the normalized radial profile of $\overline{\unicode[STIX]{x1D6E9}}$ is written as

(A 14) $$\begin{eqnarray}\frac{\overline{\unicode[STIX]{x1D6E9}}}{\unicode[STIX]{x1D6E9}_{0}}=g\left(\frac{r}{l_{\unicode[STIX]{x1D703}}}\right).\end{eqnarray}$$

Let $R_{\unicode[STIX]{x1D6E9}}$ be defined as the radial location where $\overline{\unicode[STIX]{x1D6E9}}=\unicode[STIX]{x1D6E9}_{0}/2$ , yielding

(A 15) $$\begin{eqnarray}g\left(\frac{R_{\unicode[STIX]{x1D6E9}}}{l_{\unicode[STIX]{x1D703}}}\right)=\frac{1}{2}.\end{eqnarray}$$

The shape function $g$ being not a function of $x$ and $g(R_{\unicode[STIX]{x1D6E9}}/l_{\unicode[STIX]{x1D703}})$ being constant ( $\equiv 1/2$ ), it immediately follows that $R_{\unicode[STIX]{x1D6E9}}/l_{\unicode[STIX]{x1D703}}$ must be constant. Since relation (2.28) $l_{\unicode[STIX]{x1D703}}\propto \hat{x}$ , we obtain $R_{\unicode[STIX]{x1D703}}\propto \hat{x}$ . This confirms that $R_{\unicode[STIX]{x1D6E9}}$ , the half-radius based on the mean temperature profile, evolves linearly. The same reasoning applied to the velocity field, with $R_{U}$ defining the radial location where $\overline{U}=U_{0}/2$ , allows us to obtain $R_{U}\propto \hat{x}$ . The usual normalized relation defining the longitudinal evolution of the half-radius based on the mean velocity profile is finally written as:

(A 16) $$\begin{eqnarray}\frac{R_{U}}{D}=B_{R_{U}}\hat{x},\end{eqnarray}$$

where $B_{R_{U}}$ is the spreading rate of the jet.

Moreover, using outer scaling $R_{u}$ and $\unicode[STIX]{x1D6E9}_{0}$ in expression (A 13), the normalized form of the streamwise derivative of $\unicode[STIX]{x1D6E9}_{0}$ is written as

(A 17) $$\begin{eqnarray}\frac{R_{U}}{D}\frac{1}{\unicode[STIX]{x1D6E9}_{0}}\frac{\text{d}\unicode[STIX]{x1D6E9}_{0}}{\text{d}\hat{x}}=-B_{R_{U}}.\end{eqnarray}$$

References

Antonia, R. A. & Mi, J. 1993 Temperature dissipation in a turbulent round jet. J. Fluid Mech. 250, 531551.Google Scholar
Antonia, R. A., Ould-Rouis, M., Anselmet, F. & Zhu, Y. 1997 Analogy between predictions of Kolmogorov and Yaglom. J. Fluid Mech. 332, 395409.Google Scholar
Antonia, R. A., Satyaprakash, R. J. & Hussain, A. K. M. F. 1980 Measurements of dissipation rate and some other characteristics of turbulent plane and circular jets. Phys. Fluids 23 (4), 695700.Google Scholar
Antonia, R. A., Smalley, R. J., Zhou, T., Anselmet, F. & Danaila, L. 2004 Similarity solution of temperature structure functions in decaying homogeneous isotropic turbulence. Phys. Rev. E 69, 016305.Google Scholar
Antonia, R. A. & Van Atta, C. W. 1978 Structure functions of temperature fluctuations in turbulent shear flows. J. Fluid Mech. 84, 561580.Google Scholar
Burattini, P., Antonia, R. A. & Danaila, L. 2005a Scale-by-scale energy budget on the axis of a turbulent round jet. J. Turbul. 6, 111.Google Scholar
Burattini, P., Antonia, R. A. & Danaila, L. 2005b Similarity in the far field of a turbulent round jet. Phys. Fluids 17, 114.Google Scholar
Chevray, R. & Tutu, N. K. 1978 Intermittency and preferential transport of heat in a round jet. J. Fluid Mech. 88 (01), 133160.Google Scholar
Danaila, L., Anselmet, F., Zhou, T. & Antonia, R. A. 1999 A generalization of Yaglom’s equation which accounts for the large-scale forcing in heated decaying turbulence. J. Fluid Mech. 391, 359372.Google Scholar
Danaila, L., Krawczynski, J. F., Thiesset, F. & Renou, B. 2012 Yaglom-like equation in axisymmetric anisotropic turbulence. Physica  D 241 (3), 216223.Google Scholar
Darisse, A., Lemay, J & Benaïssa, A. 2013a Investigation of passive scalar mixing in a turbulent free jet using simultaneous LDV and cold wire measurements. Intl J. Heat Fluid Flow 44, 284292.Google Scholar
Darisse, A., Lemay, J. & Benaïssa, A. 2013b LDV measurements of well converged third order moments in the far field of a free turbulent round jet. Exp. Therm. Fluid Sci. 44, 825833.Google Scholar
Darisse, A., Lemay, J. & Benaïssa, A. 2014 Extensive study of temperature dissipation measurements on the centerline of a turbulent round jet based on the 𝜃2 /2 budget. Exp. Fluids 55 (1), 115.Google Scholar
Darisse, A., Lemay, J. & Benaïssa, A. 2015 Budgets of turbulent kinetic energy, Reynolds stresses, variance of temperature fluctuations and turbulent heat fluxes in a round jet. J. Fluid Mech. 774, 95142.Google Scholar
Dekeyser, I. & Launder, B. E. 1983 A comparison of triple-moment temperature-velocity correlations in the asymmetric heated jet with alternative closure models. In Turbulent Shear Flows 4 (ed. Bradbury, L. J. S., Durst, F., Launder, B. E., Schmidt, F. W. & Whitelaw, J. H.), pp. 102117. Springer.Google Scholar
Djenidi, L., Antonia, R. A., Lefeuvre, N. & Lemay, J. 2016 Complete self-preservation on the axis of a turbulent round jet. J. Fluid Mech. 790, 5770.Google Scholar
Ewing, D., Frohnapfel, B., George, W. K., Pedersen, J. M. & Westerweel, J. 2007 Two-point similarity in the round jet. J. Fluid Mech. 577, 309330.Google Scholar
Friehe, C. A., Van Atta, C. W. & Gibson, C. H. 1971 Jet turbulence: dissipation rate measurements and correlations. Turbulent Shear Flows, AGARD CP-93, pp. 18.1–18.7.Google Scholar
Gibson, M. M. & Launder, B. E. 1976 On the calculation of horizontal, turbulent, free shear flows under gravitational influence. ASME J. Heat Transfer 98 (1), 8187.Google Scholar
Hanjalić, K. & Launder, B. 2011 Modelling Turbulence in Engineering and the Environment: Second-Moments Routes to Closure. Cambridge University Press.Google Scholar
Hill, R. J. 1997 Applicability of Kolmogorov’s and Monin’s equations of turbulence. J. Fluid Mech. 353, 6781.Google Scholar
Hill, R. J. 2001 Equations relating structure functions of all orders. J. Fluid Mech. 434, 379388.Google Scholar
Hussein, H. J., Capp, S. P. & George, W. K. 1994 Velocity measurements in a high-Reynolds-number, momentum-conserving, axisymmetric, turbulent jet. J. Fluid Mech. 258, 3175.Google Scholar
Lemay, J., Benaissa, A. & Antonia, R. A. 2003 Correction of cold-wire response for mean temperature dissipation rate measurements. Exp. Therm. Fluid Sci. 27 (2), 133143; 6th International Thermal Anemometry Symposium.Google Scholar
Lemay, J. & Benaïssa, A. 2001 Improvement of cold-wire response for measurement of temperature dissipation. Exp. Fluids 31, 347356.Google Scholar
Lemoine, F., Antoine, Y., Wolff, M. & Lebouche, M. 1999 Simultaneous temperature and 2D velocity measurements in a turbulent heated jet using combined laser-induces fluorescence and lda. Exp. Fluids 26, 315323.Google Scholar
Mi, J., Xu, M. & Zhou, T. 2013 Reynolds number influence on statistical behaviors of turbulence in a circular free jet. Phys. Fluids 25, 075101.Google Scholar
Panchapakesan, N. R. & Lumley, J. L. 1993a Turbulence measurements in axisymmetric jets of air and helium. Part 1. Air jet. J. Fluid Mech. 246, 197223.Google Scholar
Panchapakesan, N. R. & Lumley, J. L. 1993b Turbulence measurements in axisymmetric jets of air and helium. Part 2. Helium jet. J. Fluid Mech. 246, 225247.Google Scholar
Pietri, L., Amielh, M. & Anselmet, F. 2000 Simultaneous measurements of temperature and velocity fluctuations in a slightly heated jet combining a cold wire and Laser Doppler Anemometry. Intl J. Heat Fluid Flow 21 (1), 2236.Google Scholar
Pope, S. B. 2000 Turbulent Flows. Cambridge University Press.Google Scholar
Ruffin, E., Schiestel, R., Anselmet, F., Amielh, M. & Fulachier, L. 1994 Investigation of characteristic scales in variable density turbulent jets using a second-order model. Phys. Fluids 6 (8), 27852799.Google Scholar
Shivamoggi, B. K. & Antonia, R. A. 2000 Isotropic and axisymmetric turbulence of passive scalars. Fluid Dyn. Res. 26, 95104.Google Scholar
Tennekes, H. & Lumley, J. L. 1972 A First Course in Turbulence. MIT Press.Google Scholar
Thiesset, F., Antonia, R. A. & Djenidi, L. 2014 Consequences of self-preservation on the axis of a turbulent round jet. J. Fluid Mech. 748, R2.Google Scholar
Yaglom, A. M. 1949 On the local structure of the temperature field in a turbulent flow. Dokl. Akad. Nauk SSSR 69, 743746.Google Scholar
Figure 0

Figure 1. Streamwise evolution of $R_{U}/D$, $U_{0}/U_{j}$, $\unicode[STIX]{x1D6E9}_{0}/\unicode[STIX]{x1D6E9}_{j}$, $\overline{u^{2}}/U_{j}^{2}$, $\overline{\unicode[STIX]{x1D703}^{2}}/\unicode[STIX]{x1D6E9}_{j}^{2}$ and $\unicode[STIX]{x1D716}_{\unicode[STIX]{x1D703}}D/(U_{j}\unicode[STIX]{x1D6E9}_{j}^{2})$ along the jet axis (symbols) and their corresponding power laws (measurements of Darisse, Lemay & Benaïssa 2013a,b; Darisse et al.2015).

Figure 1

Table 1. Definition of the power laws shown in figure 1 and obtained from the measurements of Darisse et al. (2015) ($Re_{D}=1.4\times 10^{5}$, $Re_{\unicode[STIX]{x1D706}}=548$ and $Pe_{\unicode[STIX]{x1D706}}=211$).

Figure 2

Figure 2. Second-order structure function of temperature fluctuations on the jet axis, at $x/D=30$, normalized with Kolmogorov scales $\unicode[STIX]{x1D703}_{K}$ and $\unicode[STIX]{x1D702}$ (flow conditions similar to those of Darisse et al.2013a,b, 2015). Symbols: measurements performed using two cold-wire probes separated by a distance $s$ in the streamwise direction; open symbols: raw data; closed symbols: signals compensated for the attenuation resulting from the wire time constant (see Lemay & Benaïssa 2001; Lemay, Benaissa & Antonia 2003; Darisse et al.2014). Line: measurements performed using one cold-wire probe (compensated signal) and the assumption of Taylor’s hypothesis ($s=\unicode[STIX]{x1D70F}U_{0}$, where $\unicode[STIX]{x1D70F}$ is the time increment).

Figure 3

Figure 3. Temperature structure functions $\overline{(\unicode[STIX]{x1D6FF}\unicode[STIX]{x1D703})^{2}}$ and $-\overline{(\unicode[STIX]{x1D6FF}\unicode[STIX]{x1D703})^{3}}$ on the jet axis at 7 streamwise locations extending from $x/D$= 30 to 60, in steps of 5; normalization with $\unicode[STIX]{x1D703}_{K}$ and $\unicode[STIX]{x1D702}$ (top), $\overline{\unicode[STIX]{x1D703}^{2}}$ and $\unicode[STIX]{x1D706}$ (centre), $\unicode[STIX]{x1D6E9}_{0}$ and $R_{U}$ (bottom); flow conditions similar to those of Darisse et al. (2013a,b, 2015).

Figure 4

Figure 4. Radial distributions of normalized velocity–temperature moments near the jet axis; right side, correlation coefficient $\unicode[STIX]{x1D70C}_{u\unicode[STIX]{x1D703}}$: Darisse et al. (2015) (black diamonds), Chevray & Tutu (1978) (blue squares) and model defined by (3.1) (red dashed line); left scale, $\overline{v\unicode[STIX]{x1D703}^{2}}/(U_{0}\unicode[STIX]{x1D6E9}_{0}^{2})$: Darisse et al. (2015) (black circles) with best fit (black line), model of Dekeyser & Launder (1983), equation (3.4), (green line) and simplified model defined by (3.5) (red dashed line).

Figure 5

Table 2. Estimates obtained from the simple empirical expressions compared with the results presented by Darisse et al. (2015) (DLB, measurements in a heated jet), Antonia & Mi (1993) (AM, measurements in a heated jet), Panchapakesan & Lumley (1993b) (PL, measurements in the non-buoyant region of a helium jet) and Ruffin et al. (1994) (RAL, simulations using second-order turbulence models with a passive scalar).