This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Lagrangian acceleration in fully developed turbulence and its Eulerian decompositions

Dhawal Buaria dhawal.buaria@nyu.edu Tandon School of Engineering, New York University, New York, NY 11201, USA Max Planck Institute for Dynamics and Self-Organization, 37077 Göttingen, Germany    Katepalli R. Sreenivasan Tandon School of Engineering, New York University, New York, NY 11201, USA
Abstract

We study the properties of various Eulerian contributions to fluid particle acceleration by using well-resolved direct numerical simulations of isotropic turbulence, with the grid resolution as high as 12288312288^{3} and the Taylor-scale Reynolds number Rλ{R_{\lambda}} in the range between 140 and 1300. The variance of convective acceleration, when normalized by Kolmogorov scales, increases linearly with RλR_{\lambda}, consistent with simple theoretical arguments, but very strongly differing from phenomenological predictions of Kolmogorov’s hypothesis as well as Eulerian multifractal models. The scaling of the local acceleration is also linear RλR_{\lambda} to the leading order, but more complex in detail. The strong cancellation between the local and convective acceleration – faithful to the random sweeping hypothesis – results in the variance of the Lagrangian acceleration increasing only as Rλ0.25R_{\lambda}^{0.25}, as recently shown by Buaria & Sreenivasan [Phys. Rev. Lett. 128, 234502 (2022)]. The acceleration variance is dominated by irrotational pressure gradient contributions, whose variance also follows an Rλ0.25R_{\lambda}^{0.25} scaling; the solenoidal viscous contributions are relatively small and follow a Rλ0.13R_{\lambda}^{0.13}, consistent with Eulerian multifractal predictions.

I Introduction

In classical mechanics, the dynamics of particle motion is characterized by the acceleration 𝐚{\mathbf{a}}, defined by the rate of change of the particle velocity 𝐮{\mathbf{u}} in a Lagrangian frame. Given its fundamental role, the statistics of acceleration are of substantial interest in the study of turbulent flows La Porta et al. (2001); Toschi and Bodenschatz (2009); Stelzenmuller et al. (2017); Buaria et al. (2020a); Bec et al. (2006) and also for stochastic modeling of transport phenomena Sawford (1991); Wyngaard (1992); Pope (1994); Wilson and Sawford (1996). Of particular interest is the scaling of acceleration variance |𝐚|2\langle|{\mathbf{a}}|^{2}\rangle which, according to Kolmogorov’s 1941 mean-field phenomenology Kolmogorov (1941), can be written as |𝐚|2=a0ϵ3/2ν1/2\langle|{\mathbf{a}}|^{2}\rangle=a_{0}\langle\epsilon\rangle^{3/2}\nu^{-1/2} Monin and Yaglom (1975), where ϵ\langle\epsilon\rangle is the mean dissipation rate, ν\nu is the kinematic viscosity and and a0a_{0} is a universal constant. However, it is widely known that due to small scale intermittency, a0a_{0} is instead a variable that depends on the Reynolds number. Following a few decades of investigations Yeung and Pope (1989); Vedula and Yeung (1999); Gotoh and Rogallo (1999); Voth et al. (2002); Sawford et al. (2003); Mordant et al. (2004); Gylfason et al. (2004); Yeung et al. (2006); Ishihara et al. (2007), recent data from direct numerical simulations (DNS) of isotropic turbulence at high Reynolds numbers have established that a0Rλ0.25a_{0}\sim R_{\lambda}^{0.25} Buaria and Sreenivasan (2022a), where Rλ{R_{\lambda}} is the Taylor-scale Reynolds number. This result is at odds with predictions from both Kolmogorov’s theory and multifractal models Borgas (1993); Biferale et al. (2004), and hence requires better understanding. In this Letter, our aim is to further analyze this result, especially in terms of various underlying contributions to acceleration.

We first note that, in fluid flows, an Eulerian viewpoint is often more convenient, whereby the Lagrangian or material derivative is defined as

𝐚=D𝐮/Dt=𝐮/t+𝐮𝐮,\displaystyle{\mathbf{a}}=D{\mathbf{u}}/Dt=\partial{\mathbf{u}}/\partial t+{\mathbf{u}}\cdot\nabla{\mathbf{u}}\ , (1)

where 𝐚L𝐮/t{\mathbf{a}}_{L}\equiv\partial{\mathbf{u}}/\partial t is the local component, capturing the unsteady rate of change at a fixed spatial position, and 𝐚C𝐮𝐮{\mathbf{a}}_{C}\equiv{\mathbf{u}}\cdot\nabla{\mathbf{u}} is the convective component, capturing the rate of change due to spatial variations. That is, 𝐚=𝐚L+𝐚C{\mathbf{a}}={\mathbf{a}}_{L}+{\mathbf{a}}_{C}. In addition, the dynamics of fluid motion in incompressible turbulent flows is governed by the Navier-Stokes equations

𝐚=P+ν2𝐮,\displaystyle{\mathbf{a}}=-\nabla P+\nu\nabla^{2}{\mathbf{u}}\ , (2)

where PP is the kinematic pressure. Since 𝐮=0\nabla\cdot{\mathbf{u}}=0 from incompressibility, the viscous term is solenoidal as well, whereas the pressure gradient term is irrotational, i.e., its curl is zero. Thus, the acceleration can also be written as 𝐚=𝐚I+𝐚S{\mathbf{a}}={\mathbf{a}}_{I}+{\mathbf{a}}_{S} (with suffixes II and SS for irrotational and solenoidal, respectively), with 𝐚IP{\mathbf{a}}_{I}\equiv-\nabla P and 𝐚Sν2𝐮{\mathbf{a}}_{S}\equiv\nu\nabla^{2}{\mathbf{u}}.

In this Letter, we shall consider both methods of decomposition and study how they relate to observed scaling of acceleration variance. Utilizing data from the state-of-the-art DNS of isotropic turbulence, we show that the variance of convective acceleration varies as RλR_{\lambda}, which follows from very simple theoretical arguments, but differs from multifractal predictions. The variance of local component also varies as RλR_{\lambda} to the leading order (with weaker second order dependencies), while always remaining slightly smaller than 𝐚C2{\mathbf{a}}_{C}^{2}. The Lagrangian acceleration results from strong cancellation between these two large quantities, varying as Rλ0.25R_{\lambda}^{0.25}. We additionally explore how the properties 𝐚I{\mathbf{a}}_{I} and 𝐚S{\mathbf{a}}_{S} relate to local and convective accelerations.

II Numerical approach and database

The DNS data analyzed here are obtained by solving the incompressible Navier-Stokes equations, corresponding to the canonical setup of forced stationary isotropic turbulence in a periodic domain Ishihara et al. (2009); Buaria et al. (2019). Highly accurate Fourier pseudo-spectral methods are utilized for spatial calculations, with aliasing errors controlled using a combination of grid-shifting and truncation Rogallo (1981). An explicit second-order Runge-Kutta scheme is used for time integration. The database for the present work is the same as that of our recent study on acceleration Buaria and Sreenivasan (2022a) and several other recent works Buaria and Sreenivasan (2020); Buaria et al. (2020b); Buaria and Pumir (2021, 2022); Buaria et al. (2022); Buaria and Sreenivasan (2022b). The grid resolution is as high as 12288312288^{3} and the Taylor-scale Reynolds number Rλ{R_{\lambda}} lies in the range 1401300140-1300. Convergence with respect to small-scale resolution and statistical sampling has been assessed in these previous studies.

As in Buaria and Sreenivasan (2022a), we have also calculated the relevant statistics using Lagrangian fluid particle trajectories in the same range of Rλ{R_{\lambda}}, albeit with lower small-scale resolution Buaria et al. (2015, 2016); Buaria and Yeung (2017). At the level of second order moments reported in this work, the statistics are essentially identical from both Eulerian and Lagrangian data. However, the Lagrangian particle data are not suitable for studying higher order moments, due both to the lack of resolution and accumulated numerical errors resulting from interpolation of particle velocities Yeung and Pope (1989).

III Results

III.1 Theoretical analysis

Before analyzing the DNS data, we present a brief theoretical analysis to obtain simplified relations between various Eulerian components of acceleration. For instance, it is straightforward to prove that in homogeneous turbulence, the correlation between an irrotational and a solenoidal vector is always zero (see Appendix). From this property, it follows that

𝐚I𝐚S=0\displaystyle\langle{\mathbf{a}}_{I}\cdot{\mathbf{a}}_{S}\rangle=0 (3)
𝐚I𝐚L=0.\displaystyle\langle{\mathbf{a}}_{I}\cdot{\mathbf{a}}_{L}\rangle=0. (4)

Additionally, using statistical stationarity, we can show (see Appendix) that

𝐚S𝐚L=0.\displaystyle\langle{\mathbf{a}}_{S}\cdot{\mathbf{a}}_{L}\rangle=0. (5)

Since 𝐚=𝐚I+𝐚S{\mathbf{a}}={\mathbf{a}}_{I}+{\mathbf{a}}_{S}, it also follows from Eqs. (4) and (5) that

𝐚𝐚L=0,\displaystyle\langle{\mathbf{a}}\cdot{\mathbf{a}}_{L}\rangle=0\ , (6)

i.e., the Lagrangian acceleration D𝐮/DtD{\mathbf{u}}/Dt is uncorrelated to the Eulerian acceleration u/t\partial u/\partial t. This property directly yields the following result:

𝐚L𝐚C=|𝐚L|2.\displaystyle\langle{\mathbf{a}}_{L}\cdot{\mathbf{a}}_{C}\rangle=-\langle|{\mathbf{a}}_{L}|^{2}\rangle. (7)

These relations allow us to write the acceleration variance as

|𝐚|2\displaystyle\langle|{\mathbf{a}}|^{2}\rangle =|𝐚I|2+|𝐚S|2\displaystyle=\langle|{\mathbf{a}}_{I}|^{2}\rangle+\langle|{\mathbf{a}}_{S}|^{2}\rangle (8)
|𝐚|2\displaystyle\langle|{\mathbf{a}}|^{2}\rangle =|𝐚C|2|𝐚L|2.\displaystyle=\langle|{\mathbf{a}}_{C}|^{2}\rangle-\langle|{\mathbf{a}}_{L}|^{2}\rangle. (9)

Thus, while the acceleration variance is given by the sum of variances of the pressure gradient and viscous terms, it is also obtained via a direct cancellation of convective and local components. We will now explore how the scaling of all these Eulerian contributions affect the scaling of acceleration variance itself.

III.2 Properties of 𝐚I{\mathbf{a}}_{I} and 𝐚S{\mathbf{a}}_{S}

It is well known that acceleration variance is dominated by the irrotational pressure gradient contribution and the corresponding viscous contribution is negligible, i.e., |𝐚I||𝐚S||{\mathbf{a}}_{I}|\gg|{\mathbf{a}}_{S}| Vedula and Yeung (1999); Tsinober et al. (2001). We first reaffirm this result in Fig. 1a which shows the fractional contributions of 𝐚I{\mathbf{a}}_{I} and 𝐚S{\mathbf{a}}_{S}, and also the correlation 𝐚I𝐚S\langle{\mathbf{a}}_{I}\cdot{\mathbf{a}}_{S}\rangle, which is zero as expected. It is evident that |𝐚|2|𝐚I|2\langle|{\mathbf{a}}|^{2}\rangle\approx\langle|{\mathbf{a}}_{I}|^{2}\rangle. We can readily show that

𝐚𝐚I=𝐚C𝐚I=|𝐚I|2,𝐚𝐚S=𝐚C𝐚S=|𝐚S|2,\displaystyle\langle{\mathbf{a}}\cdot{\mathbf{a}}_{I}\rangle=\langle{\mathbf{a}}_{C}\cdot{\mathbf{a}}_{I}\rangle=\langle|{\mathbf{a}}_{I}|^{2}\rangle\ ,\ \ \ \ \langle{\mathbf{a}}\cdot{\mathbf{a}}_{S}\rangle=\langle{\mathbf{a}}_{C}\cdot{\mathbf{a}}_{S}\rangle=\langle|{\mathbf{a}}_{S}|^{2}\rangle\ , (10)

which demonstrate that both the pressure gradient and viscous contributions predominantly arise from the convective component. This is not surprising since we saw earlier that 𝐚I𝐚L\langle{\mathbf{a}}_{I}\cdot{\mathbf{a}}_{L}\rangle and 𝐚S𝐚L\langle{\mathbf{a}}_{S}\cdot{\mathbf{a}}_{L}\rangle are both zero. We shall further elaborate this point in the next subsection.

Refer to caption
Refer to caption
Figure 1: (a) Fractional contributions of the irrotational pressure gradient and solenoidal viscous terms to the acceleration variance, as well as their mutual correlation, as functions of Rλ{R_{\lambda}}. (b) Variance of the solenoidal viscous acceleration normalized by the Kolmogorov scales, as a function of Rλ{R_{\lambda}}.

It is worth noting that, while the contribution from the viscous term 𝐚S{\mathbf{a}}_{S} is negligible in comparison to 𝐚I{\mathbf{a}}_{I}, it is nevertheless finite and intimately connected to the fundamental dynamics of turbulence. In particular, its variance can be written as Monin and Yaglom (1975); Vedula and Yeung (1999)

|𝐚S|2/aK2=352𝒮(15)3/2,\displaystyle\langle|{\mathbf{a}}_{S}|^{2}\rangle/a_{K}^{2}=-\frac{35}{2}\frac{\mathcal{S}}{(15)^{3/2}}\ , (11)

where aK=ϵ3/4ν1/4a_{K}=\langle\epsilon\rangle^{3/4}\nu^{-1/4} and 𝒮\mathcal{S} is the skewness of longitudinal velocity gradients, which is always negative in turbulence, characterizing the energy cascade from large to small scales Batchelor (1967); Frisch (1995). The skewness can also be related to vortex stretching Betchov (1956) and is known to weakly increase in magnitude as Rλ0.13R_{\lambda}^{0.13} Buaria et al. (2020c). This scaling matches the prediction from extending Eulerian multifractals to Lagrangian variables Sreenivasan and Meneveau (1988); Frisch (1995); Borgas (1993). Figure 1b shows a satisfactory agreement with this result (except at the lowest Reynolds number).

However, our recent work Buaria and Sreenivasan (2022a) computed the acceleration variance |𝐚I|2\langle|{\mathbf{a}}_{I}|^{2}\rangle and found it to vary as Rλ0.25R_{\lambda}^{0.25}. We expressed the acceleration analytically in terms of the fourth order velocity structure functions Hill and Wilczak (1995); Hill (2002), and showed that

|𝐚|2/aK2|𝐚I|2/aK2Rλ0.25.\displaystyle\langle|{\mathbf{a}}|^{2}\rangle/a_{K}^{2}\approx\langle|{\mathbf{a}}_{I}|^{2}\rangle/a_{K}^{2}\sim R_{\lambda}^{0.25}\ . (12)

The data from various sources, including our own DNS, show excellent agreement with this prediction (see Buaria and Sreenivasan (2022a); we also reaffirm it below). It then follows that an extension of Eulerian multifractals to explain intermittency of Lagrangian quantities is fraught with major uncertainties.

III.3 Properties of 𝐚L{\mathbf{a}}_{L} and 𝐚C{\mathbf{a}}_{C}

From Eq. (9), acceleration variance results from direct cancellation between the variances of 𝐚C{\mathbf{a}}_{C} and 𝐚L{\mathbf{a}}_{L}. This cancellation is consistent with the random sweeping hypothesis proposed by Kraichnan Kraichnan (1964) – see also Tennekes Tennekes (1975) – which states that the small scales of turbulence are swept past an Eulerian observer on a much shorter time scale than the time scale governing their dynamical evolution. The nominal validity of this hypothesis is also implicitly reflected in the fact that 𝐚=D𝐮/Dt{\mathbf{a}}=D{\mathbf{u}}/Dt and 𝐚L=𝐮/t{\mathbf{a}}_{L}=\partial{\mathbf{u}}/\partial t are uncorrelated (see Eq. (6)).

The convective acceleration 𝐚C=𝐮𝐮{\mathbf{a}}_{C}={\mathbf{u}}\cdot\nabla{\mathbf{u}} essentially represents a correlation between the velocity and its gradients. Given the general understanding that the former characterizes the large scales and the latter the small scales, we can assume that the two are essentially uncorrelated (provided RλR_{\lambda} is sufficiently high). Thus, simple scaling arguments suggest that |𝐚C|u/τK|{\mathbf{a}}_{C}|\sim u^{\prime}/\tau_{K} where uu^{\prime} is the root-mean-square (rms) velocity and τK\tau_{K} is the Kolmogorov time scale, characterizing the rms of velocity gradients. We then have

|𝐚C|2/aK2=cRλ,\displaystyle\langle|{\mathbf{a}}_{C}|^{2}\rangle/a_{K}^{2}=c\ R_{\lambda}\ , (13)

where cc is some proportionality constant and we have utilized the classical estimate u/uKRλ1/2u^{\prime}/u_{K}\sim R_{\lambda}^{1/2} Frisch (1995) (uKu_{K} being the Kolmogorov velocity scale). To the first order, it can be also expected that |𝐚L|2\langle|{\mathbf{a}}_{L}|^{2}\rangle also follows a similar scaling. This can be inferred indirectly by noting that

|𝐚L|2\displaystyle\langle|{\mathbf{a}}_{L}|^{2}\rangle =|𝐚C|2|𝐚|2\displaystyle=\langle|{\mathbf{a}}_{C}|^{2}\rangle-\langle|{\mathbf{a}}|^{2}\rangle (14)
=|𝐚C|2|𝐚I|2|𝐚S|2.\displaystyle=\langle|{\mathbf{a}}_{C}|^{2}\rangle-\langle|{\mathbf{a}}_{I}|^{2}\rangle-\langle|{\mathbf{a}}_{S}|^{2}\rangle. (15)

where observations show that the scaling of |𝐚C|2|{\mathbf{a}}_{C}|^{2} dominates over other two components.

Refer to caption
Refer to caption
Refer to caption
Figure 2: (a) Variances of local and convective acceleration, normalized by Kolmogorov scales, as a function of Rλ{R_{\lambda}}. (b) Ratio of variance of local and convective acceleration; inset shows deficit of the ratio from unity (c) Difference between the convective and local acceleration as a function of Rλ{R_{\lambda}}, compared directly with acceleration variance.

Figure 2a shows the variances of local and convective acceleration. It can be immediately seen that |𝐚C|2/aK2|{\mathbf{a}}_{C}|^{2}/a_{K}^{2} follows a simple linear scaling in RλR_{\lambda}, as anticipated in Eq. (13). On the other hand, |𝐚L|2/aK2|{\mathbf{a}}_{L}|^{2}/a_{K}^{2} approaches this scaling as Rλ{R_{\lambda}} increases, but noticeably deviates at low Rλ{R_{\lambda}}. Using the results in Eqs. (11)-(15), the precise scaling of |𝐚L|2\langle|{\mathbf{a}}_{L}|^{2}\rangle can be quantified as

|𝐚L|2/aK2=cRλc1Rλ0.25c2Rλ0.13,\displaystyle\langle|{\mathbf{a}}_{L}|^{2}\rangle/a_{K}^{2}=cR_{\lambda}-c_{1}R_{\lambda}^{0.25}-c_{2}R_{\lambda}^{0.13}\ , (16)

where c1c_{1} and c2c_{2} are proportionality constants. Evidently, the deviations at lower Rλ{R_{\lambda}} can be understood in terms of these additional terms. It also follows from here that the ratio |𝐚L|2/|𝐚C|2\langle|{\mathbf{a}}_{L}|^{2}\rangle/\langle|{\mathbf{a}}_{C}|^{2}\rangle has the form 1(c1/c)Rλ0.75(c2/c)Rλ0.871-(c_{1}/c)R_{\lambda}^{-0.75}-(c_{2}/c)R_{\lambda}^{-0.87}. Asymptotically, when normalized by Kolmogorov variable, |𝐚|2=|𝐚C|2|𝐚L|2c1Rλ0.25\langle|{\mathbf{a}}|^{2}\rangle=\langle|{\mathbf{a}}_{C}|^{2}\rangle-\langle|{\mathbf{a}}_{L}|^{2}\rangle\approx c_{1}R_{\lambda}^{-0.25}.

To verify the behaviors of 𝐚L{\mathbf{a}}_{L} and 𝐚C{\mathbf{a}}_{C} we plot in Fig. 2b the ratio y=|𝐚L|2/|𝐚C|2y=\langle|{\mathbf{a}}_{L}|^{2}\rangle/\langle|{\mathbf{a}}_{C}|^{2}\rangle as a function of Rλ{R_{\lambda}}. The inset shows 1y1-y, which is in excellent agreement with a power law Rλ0.75R_{\lambda}^{-0.75}. The ratio steadily approaches unity, which demonstrates that, asymptotically, only the Rλ0.25R_{\lambda}^{0.25} term contributes to Lagrangian acceleration. This is also confirmed by Fig. 2c, which shows |𝐚C|2|𝐚L|2\langle|{\mathbf{a}}_{C}|^{2}\rangle-\langle|{\mathbf{a}}_{L}|^{2}\rangle normalized by Kolmogorov scales with the acceleration variance data from Buaria and Sreenivasan (2022a) – both sets of points are indistinguishable and in excellent agreement with Rλ0.25R_{\lambda}^{0.25} scaling.

III.4 Further analyzing the role of 𝐚C{\mathbf{a}}_{C}

The near cancellation between 𝐚L{\mathbf{a}}_{L} and 𝐚C{\mathbf{a}}_{C} can be further analyzed by noticing that 𝐚L{\mathbf{a}}_{L} is solenoidal, whereas 𝐚C{\mathbf{a}}_{C} is not; thus, they can never completely cancel each other. Further, since 𝐚I{\mathbf{a}}_{I} is irrotational and 𝐚S{\mathbf{a}}_{S} is solenoidal, we can write Tsinober et al. (2001)

𝐚CI\displaystyle{\mathbf{a}}_{C_{I}} =𝐚I,\displaystyle={\mathbf{a}}_{I}\ , (17)
𝐚L+𝐚CS\displaystyle{\mathbf{a}}_{L}+{\mathbf{a}}_{C_{S}} =𝐚S,\displaystyle={\mathbf{a}}_{S}\ , (18)

where we have decomposed 𝐚C{\mathbf{a}}_{C} into irrotational and solenoidal components, i.e., 𝐚C=𝐚CI+𝐚CS{\mathbf{a}}_{C}={\mathbf{a}}_{C_{I}}+{\mathbf{a}}_{C_{S}}. Such a decomposition can readily be implemented in Fourier space using the Helmholtz decomposition, i.e., for a vector 𝐕\mathbf{V} with Fourier coefficient 𝐕^\hat{\mathbf{V}}, the Fourier coefficients of irrotational and solenoidal parts are, respectively, given as

𝐕^I(𝐤)=(𝐤𝐕^)𝐤/k2,𝐕^S(𝐤)=𝐕^𝐕^I,\displaystyle\hat{{\mathbf{V}}}_{I}({\mathbf{k}})=({\mathbf{k}}\cdot\hat{{\mathbf{V}}}){\mathbf{k}}/k^{2}\ ,\ \ \ \ \hat{{\mathbf{V}}}_{S}({\mathbf{k}})=\hat{{\mathbf{V}}}-\hat{{\mathbf{V}}}_{I}\ , (19)

where 𝐤{\mathbf{k}} is the wave-vector and k=|𝐤|k=|{\mathbf{k}}|. Note that irrotationality is imposed in Fourier space by the condition 𝐤×𝐕^I=𝟎{\mathbf{k}}\times\hat{{\mathbf{V}}}_{I}=\mathbf{0}, and solenoidality by 𝐤𝐕S=0{\mathbf{k}}\cdot{\mathbf{V}}_{S}=0 (both of which can be easily verified).

Refer to caption
Refer to caption
Figure 3: (a) Ratio of variance of local acceleration to that of solenoidal component of convective acceleration, as a function of Rλ{R_{\lambda}}. (b) Ratio of variance of solenoidal viscous acceleration to that of local acceleration.

From the above decomposition, it trivially follows that |𝐚CI|2=|𝐚I|2|{\mathbf{a}}_{C_{I}}|^{2}=|{\mathbf{a}}_{I}|^{2} and |𝐚C|2=|𝐚CI|2+|𝐚CS|2\langle|{\mathbf{a}}_{C}|^{2}\rangle=\langle|{\mathbf{a}}_{C_{I}}|^{2}\rangle+\langle|{\mathbf{a}}_{C_{S}}|^{2}\rangle and we can also show that

𝐚L𝐚CS\displaystyle\langle{\mathbf{a}}_{L}\cdot{\mathbf{a}}_{C_{S}}\rangle =|𝐚L|2,\displaystyle=-\langle|{\mathbf{a}}_{L}|^{2}\rangle\ , (20)
|𝐚S|2\displaystyle\langle|{\mathbf{a}}_{S}|^{2}\rangle =|𝐚CS|2|𝐚L|2.\displaystyle=\langle|{\mathbf{a}}_{C_{S}}|^{2}\rangle-\langle|{\mathbf{a}}_{L}|^{2}\rangle\ . (21)

Thus, the very small solenoidal component 𝐚S{\mathbf{a}}_{S} results from near perfect cancellation between 𝐚CS{\mathbf{a}}_{C_{S}} and 𝐚L{\mathbf{a}}_{L}. We quantify this in Fig. 3. Panel a shows that the ratio |aL|2/|aCS|2\langle|a_{L}|^{2}\rangle/\langle|a_{C_{S}}|^{2}\rangle steadily approaches unity as Rλ{R_{\lambda}} increases (though it cannot be strictly unity since |𝐚S||{\mathbf{a}}_{S}| is always finite). In panel b, the ratio |aS|2/|aL|2\langle|a_{S}|^{2}\rangle/\langle|a_{L}|^{2}\rangle is shown, which steadily decreases with Rλ{R_{\lambda}} as expected.

In summary, based on the present analysis, we can essentially write |𝐚C||𝐚CS||𝐚L||𝐚||𝐚I|=|𝐚CI||𝐚S||{\mathbf{a}}_{C}|\gtrsim|{\mathbf{a}}_{C_{S}}|\gtrsim|{\mathbf{a}}_{L}|\gg|{\mathbf{a}}|\approx|{\mathbf{a}}_{I}|=|{\mathbf{a}}_{C_{I}}|\gg|{\mathbf{a}}_{S}|. Perhaps surprisingly, |𝐚|2|𝐚I|2Rλ0.25\langle|{\mathbf{a}}|^{2}\rangle\approx\langle|{\mathbf{a}}_{I}|^{2}\rangle\sim R_{\lambda}^{0.25}, while variances of the components 𝐚C{\mathbf{a}}_{C} and 𝐚L{\mathbf{a}}_{L} have far stronger dependencies on RλR_{\lambda}; for example, |𝐚C|2\langle|{\mathbf{a}}_{C}|^{2}\rangle essentially scales as RλR_{\lambda}.

IV conclusions

The most interesting result of the analysis is that the local and convection terms of acceleration are anti-correlated and both of them depart from Kolmogorov’s paradigm very strongly (also from the multifractal formalism). In particular, the variance of both essentially scale linearly with RλR_{\lambda}. The two terms are, however, strongly anti-correlated. Thus, the difference between the two, which specifies the Lagrangian acceleration, increases as Rλ0.25R_{\lambda}^{0.25} Buaria and Sreenivasan (2022a). This result, which is an indication that the two terms are separately much more intermittent than their algebraic (vector) sum, is at odds with Kolmogorov’s and multifractal formalisms, but not nearly as much as their sum. The scaling |𝐚C|2Rλ\langle|{\mathbf{a}}_{C}|^{2}\rangle\sim R_{\lambda} comes from the assumption that 𝐮\bf u and 𝐮\nabla\bf u are uncorrelated, so essentially |𝐚C|2\langle|{\mathbf{a}}_{C}|^{2}\rangle follows the same scaling as |𝐮|2\langle|{\mathbf{u}}|^{2}\rangle. The interpretation is that the small scales are simply swept by the large scale velocity without getting affected, which is consistent with the random sweeping hypothesis.

Appendix A Vanishing correlations between various Eulerian contributions to acceleration

Let us assume that vector 𝐀\mathbf{A} is irrotational and vector 𝐁\mathbf{B} is solenoidal; this implies ×A=𝟎\nabla\times A=\mathbf{0} and B=0\nabla\cdot B=0 (or Bi/xi=0\partial B_{i}/\partial x_{i}=0). For the former, we can write Ai=ϕ/xiA_{i}=\partial\phi/\partial x_{i}, where ϕ\phi is some scalar quantity. Thus, the correlation can be simplified as:

𝐀𝐁\displaystyle\langle\mathbf{A}\cdot\mathbf{B}\rangle =ϕxiBi\displaystyle=\left\langle\frac{\partial\phi}{\partial x_{i}}B_{i}\right\rangle (22)
=(ϕBi)xiϕBixi\displaystyle=\left\langle\frac{\partial(\phi B_{i})}{\partial x_{i}}-\phi\frac{\partial B_{i}}{\partial x_{i}}\right\rangle (23)
=ϕBixiϕBixi\displaystyle=\frac{\partial\langle\phi B_{i}\rangle}{\partial x_{i}}-\left\langle\phi\frac{\partial B_{i}}{\partial x_{i}}\right\rangle (24)
=0,\displaystyle=0, (25)

where the first term is zero from statistical homogeneity and the second term is zero since Bi/xi\partial B_{i}/\partial x_{i}.

Thus, for the components of acceleration, we can write

𝐚I𝐚S=0\displaystyle\langle{\mathbf{a}}_{I}\cdot{\mathbf{a}}_{S}\rangle=0 (26)
𝐚I𝐚L=0.\displaystyle\langle{\mathbf{a}}_{I}\cdot{\mathbf{a}}_{L}\rangle=0. (27)

For the correlation between 𝐚S{\mathbf{a}}_{S} and 𝐚L{\mathbf{a}}_{L}, the following steps have to be considered:

𝐚S𝐚L\displaystyle\langle{\mathbf{a}}_{S}\cdot{\mathbf{a}}_{L}\rangle =ν2uixkxkuit\displaystyle=\left\langle\nu\frac{\partial^{2}u_{i}}{\partial x_{k}\partial x_{k}}\cdot\frac{\partial u_{i}}{\partial t}\right\rangle (28)
=νxk(uixkuit)uixkt(uixk)\displaystyle=\nu\left\langle\frac{\partial}{\partial x_{k}}\left(\frac{\partial u_{i}}{\partial x_{k}}\frac{\partial u_{i}}{\partial t}\right)-\frac{\partial u_{i}}{\partial x_{k}}\cdot\frac{\partial}{\partial t}\left(\frac{\partial u_{i}}{\partial x_{k}}\right)\right\rangle (29)
=νxk(uixkuit)ν2t(uixk)2\displaystyle=\nu\left\langle\frac{\partial}{\partial x_{k}}\left(\frac{\partial u_{i}}{\partial x_{k}}\frac{\partial u_{i}}{\partial t}\right)\right\rangle-\frac{\nu}{2}\left\langle\frac{\partial}{\partial t}\left(\frac{\partial u_{i}}{\partial x_{k}}\right)^{2}\right\rangle (30)
=0.\displaystyle=0. (31)

The last step follows from the fact that the first term is zero from statistical homogeneity, whereas the second term is zero from statistical stationarity. Since 𝐚=𝐚I+𝐚S{\mathbf{a}}={\mathbf{a}}_{I}+{\mathbf{a}}_{S}, it also follows that

𝐚𝐚L=0.\displaystyle\langle{\mathbf{a}}\cdot{\mathbf{a}}_{L}\rangle=0. (32)

References

  • La Porta et al. (2001) A. La Porta, G. A. Voth, A. M. Crawford, J. Alexander,  and E. Bodenschatz, “Fluid particle accelerations in fully developed turbulence,” Nature 409, 1017–1019 (2001).
  • Toschi and Bodenschatz (2009) F. Toschi and E. Bodenschatz, “Lagrangian properties of particles in turbulence,” Annu. Rev. Fluid Mech. 41, 375–404 (2009).
  • Stelzenmuller et al. (2017) N. Stelzenmuller, J. I. Polanco, L. Vignal, I. Vinkovic,  and N. Mordant, “Lagrangian acceleration statistics in a turbulent channel flow,” Phys. Rev. Fluids 2, 054602 (2017).
  • Buaria et al. (2020a) D. Buaria, A. Pumir, F. Feraco, R. Marino, A. Pouquet, D. Rosenberg,  and L. Primavera, “Single-particle Lagrangian statistics from direct numerical simulations of rotating-stratified turbulence,” Phys. Rev. Fluids 5, 064801 (2020a).
  • Bec et al. (2006) J. Bec, L. Biferale, G. Boffetta, A. Celani, M. Cencini, A. Lanotte, S Musacchio,  and F. Toschi, “Acceleration statistics of heavy particles in turbulence,” J. Fluid Mech. 550, 349–358 (2006).
  • Sawford (1991) B. L. Sawford, “Reynolds number effects in Lagrangian stochastic models of turbulent dispersion,” Phys. Fluids A 3, 1577–1586 (1991).
  • Wyngaard (1992) J. C. Wyngaard, “Atmospheric turbulence,” Annu. Rev. Fluid Mech. 24, 205–234 (1992).
  • Pope (1994) S. B. Pope, “Lagrangian PDF methods for turbulent flows,” Annu. Rev. Fluid Mech. 26, 23–63 (1994).
  • Wilson and Sawford (1996) J. D. Wilson and B. L. Sawford, “Review of Lagrangian stochastic models for trajectories in the turbulent atmosphere,”  (1996).
  • Kolmogorov (1941) A. N. Kolmogorov, “Local structure of turbulence in an incompressible fluid for very large Reynolds numbers,” Dokl. Akad. Nauk. SSSR 30, 299–303 (1941).
  • Monin and Yaglom (1975) A. S. Monin and A. M. Yaglom, Statistical Fluid Mechanics, Vol. 2 (MIT Press, 1975).
  • Yeung and Pope (1989) P. K. Yeung and S. B. Pope, “Lagrangian statistics from direct numerical simulations of isotropic turbulence,” J. Fluid Mech. 207, 531–586 (1989).
  • Vedula and Yeung (1999) P. Vedula and P. K. Yeung, “Similarity scaling of acceleration and pressure statistics in numerical simulations of isotropic turbulence,” Phys Fluids 11, 1208–1220 (1999).
  • Gotoh and Rogallo (1999) T. Gotoh and R. S. Rogallo, “Intermittency and scaling of pressure at small scales in forced isotropic turbulence,” J. Fluid Mech. 396, 257–285 (1999).
  • Voth et al. (2002) G. A. Voth, A. La Porta, A. M. Crawford, J. Alexander,  and E. Bodenschatz, “Measurement of particle accelerations in fully developed turbulence,” J. Fluid Mech. 469, 121–160 (2002).
  • Sawford et al. (2003) B. L. Sawford, P. K. Yeung, M. S. Borgas, P. Vedula, A. La Porta, A. M. Crawford,  and E. Bodenschatz, “Conditional and unconditional acceleration statistics in turbulence,” Phys. Fluids 15, 3478–3489 (2003).
  • Mordant et al. (2004) N. Mordant, E. Lévêque,  and J.-F. Pinton, “Experimental and numerical study of the Lagrangian dynamics of high Reynolds turbulence,” New J. Phys. 6, 116 (2004).
  • Gylfason et al. (2004) A. Gylfason, S. Ayyalasomayajula,  and Z. Warhaft, “Intermittency, pressure and acceleration statistics from hot-wire measurements in wind-tunnel turbulence,” J. Fluid Mech. 501, 213–229 (2004).
  • Yeung et al. (2006) P. K. Yeung, S. B. Pope, A. G. Lamorgese,  and D. A. Donzis, “Acceleration and dissipation statistics of numerically simulated isotropic turbulence,” Physics of fluids 18, 065103 (2006).
  • Ishihara et al. (2007) T. Ishihara, Y. Kaneda, M. Yokokawa, K. Itakura,  and A. Uno, “Small-scale statistics in high resolution of numerically isotropic turbulence,” J. Fluid Mech. 592, 335–366 (2007).
  • Buaria and Sreenivasan (2022a) D. Buaria and K. R. Sreenivasan, “Scaling of acceleration statistics in high Reynolds number turbulence,” Phys. Rev. Lett. 128, 234502 (2022a).
  • Borgas (1993) M. S. Borgas, “The multifractal Lagrangian nature of turbulence,” Philos. Trans. R. Soc. A 342, 379–411 (1993).
  • Biferale et al. (2004) L. Biferale, G. Boffetta, A. Celani, B. J. Devenish, A. Lanotte,  and F. Toschi, “Multifractal statistics of Lagrangian velocity and acceleration in turbulence,” Phys. Rev. Lett. 93, 064502 (2004).
  • Ishihara et al. (2009) T. Ishihara, T. Gotoh,  and Y. Kaneda, “Study of high-Reynolds number isotropic turbulence by direct numerical simulations,” Ann. Rev. Fluid Mech. 41, 165–80 (2009).
  • Buaria et al. (2019) D. Buaria, A. Pumir, E. Bodenschatz,  and P. K. Yeung, “Extreme velocity gradients in turbulent flows,” New J. Phys. 21, 043004 (2019).
  • Rogallo (1981) R. S. Rogallo, “Numerical experiments in homogeneous turbulence,” NASA Technical Memo  (1981).
  • Buaria and Sreenivasan (2020) D. Buaria and K. R. Sreenivasan, “Dissipation range of the energy spectrum in high Reynolds number turbulence,” Phys. Rev. Fluids 5, 092601(R) (2020).
  • Buaria et al. (2020b) D. Buaria, A. Pumir,  and E. Bodenschatz, “Self-attenuation of extreme events in Navier-Stokes turbulence,” Nat. Commun. 11, 5852 (2020b).
  • Buaria and Pumir (2021) D. Buaria and A. Pumir, “Nonlocal amplification of intense vorticity in turbulent flows,” Phys. Rev. Research 3, 042020 (2021).
  • Buaria and Pumir (2022) D. Buaria and A. Pumir, “Vorticity-strain rate dynamics and the smallest scales of turbulence,” Phys. Rev. Lett. 128, 094501 (2022).
  • Buaria et al. (2022) D. Buaria, A. Pumir,  and E. Bodenschatz, “Generation of intense dissipation in high Reynolds number turbulence,” Philos. Trans. R. Soc. A 380, 20210088 (2022).
  • Buaria and Sreenivasan (2022b) D. Buaria and K. R. Sreenivasan, “Intermittency of turbulent velocity and scalar fields using three-dimensional local averaging,” Phys. Rev. Fluids 7, L072601 (2022b).
  • Buaria et al. (2015) D. Buaria, B. L. Sawford,  and P. K. Yeung, “Characteristics of backward and forward two-particle relative dispersion in turbulence at different Reynolds numbers,” Phys. Fluids 27, 105101 (2015).
  • Buaria et al. (2016) D. Buaria, P. K. Yeung,  and B. L. Sawford, “A Lagrangian study of turbulent mixing: forward and backward dispersion of molecular trajectories in isotropic turbulence,” J. Fluid Mech. 799, 352–382 (2016).
  • Buaria and Yeung (2017) D. Buaria and P. K. Yeung, “A highly scalable particle tracking algorithm using partitioned global address space (PGAS) programming for extreme-scale turbulence simulations,” Comput. Phys. Commun. 221, 246 – 258 (2017).
  • Tsinober et al. (2001) A. Tsinober, P. Vedula,  and P. K. Yeung, “Random Taylor hypothesis and the behavior of local and convective accelerations in isotropic turbulence,” Phys. Fluids 13, 1974–1984 (2001).
  • Batchelor (1967) G. K. Batchelor, An Introduction to Fluid Dynamics (Cambridge University Press, Cambridge, 1967).
  • Frisch (1995) U. Frisch, Turbulence: The Legacy of Kolmogorov (Cambridge University Press, Cambridge, 1995).
  • Betchov (1956) R. Betchov, “An inequality concerning the production of vorticity in isotropic turbulence,” J. Fluid Mech. 1, 497–504 (1956).
  • Buaria et al. (2020c) D. Buaria, E. Bodenschatz,  and A. Pumir, “Vortex stretching and enstrophy production in high Reynolds number turbulence,” Phys. Rev. Fluids 5, 104602 (2020c).
  • Sreenivasan and Meneveau (1988) K. R. Sreenivasan and C. Meneveau, “Singularities of the equations of fluid motion,” Phys. Rev. A 38, 6287–6295 (1988).
  • Hill and Wilczak (1995) R. J. Hill and J. M. Wilczak, “Pressure structure functions and spectra for locally isotropic turbulence,” J. Fluid Mech. 296, 247–269 (1995).
  • Hill (2002) R. J. Hill, “Scaling of acceleration in locally isotropic turbulence,” J. Fluid Mech. 452, 361–370 (2002).
  • Kraichnan (1964) R. H. Kraichnan, “Kolmogorov’s hypotheses and Eulerian turbulence theory,” Phys. Fluids 7, 1723–1734 (1964).
  • Tennekes (1975) H. Tennekes, “Eulerian and Lagrangian time microscales in isotropic turbulence,” J. Fluid Mech. 67, 561–567 (1975).