License: arXiv.org perpetual non-exclusive license
arXiv:2307.04846v2 [physics.flu-dyn] 02 Mar 2024

Saturation and multifractality of Lagrangian and Eulerian
scaling exponents in 3D isotropic turbulence

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 Department of Physics and the Courant Institute of Mathematical Sciences, New York University, New York, NY 10012, USA
(March 2, 2024)
Abstract

Inertial range scaling exponents for both Lagrangian and Eulerian structure functions are obtained from direct numerical simulations of isotropic turbulence in triply periodic domains at Taylor-scale Reynolds number up to 1300. We reaffirm that transverse Eulerian scaling exponents saturate at 2.1absent2.1\approx 2.1≈ 2.1 for moment orders p10𝑝10p\geq 10italic_p ≥ 10, significantly differing from the longitudinal exponents (which are predicted to saturate at 7.3absent7.3\approx 7.3≈ 7.3 for p30𝑝30p\geq 30italic_p ≥ 30 from a recent theory). The Lagrangian scaling exponents likewise saturate at 2absent2\approx 2≈ 2 for p8𝑝8p\geq 8italic_p ≥ 8. The saturation of Lagrangian exponents and transverse Eulerian exponents is related by the same multifractal spectrum by utilizing the well known frozen hypothesis to relate spatial and temporal scales. Furthermore, this spectrum is different from the known spectra for Eulerian longitudinal exponents, suggesting that that Lagrangian intermittency is characterized solely by transverse Eulerian intermittency. We discuss possible implication of this outlook when extending multifractal predictions to the dissipation range, especially for Lagrangian acceleration.

Turbulent flows in nature and engineering comprise a hierarchy of eddies, with smaller eddies coexisting within larger ones and extracting energy from them. To understand the deformation and rotation of smaller eddies, the key mechanisms driving energy transfers, it is essential to examine the velocity increments across a smaller eddy of size rLmuch-less-than𝑟𝐿r\ll Litalic_r ≪ italic_L (say), where L𝐿Litalic_L is the large-eddy size [1, 2, 3]. The longitudinal velocity increment δur=u(x+r)u(x)𝛿subscript𝑢𝑟𝑢𝑥𝑟𝑢𝑥\delta u_{r}=u(x+r)-u(x)italic_δ italic_u start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = italic_u ( italic_x + italic_r ) - italic_u ( italic_x ) corresponds to the case when the velocity component u(x)𝑢𝑥u(x)italic_u ( italic_x ) is in the direction of separation r𝑟ritalic_r. For velocity v(x)𝑣𝑥v(x)italic_v ( italic_x ) taken orthogonal to r𝑟ritalic_r, transverse velocity increment δvr=v(x+r)v(x)𝛿subscript𝑣𝑟𝑣𝑥𝑟𝑣𝑥\delta v_{r}=v(x+r)-v(x)italic_δ italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = italic_v ( italic_x + italic_r ) - italic_v ( italic_x ) is obtained.

The motivation to study the small eddies (and hence velocity increments) stems from their purported universality, postulated by Kolmogorov (1941) [1]—K41 henceforth—which has since become the backbone of turbulence theory and modeling [3, 4]. Building upon K41, one surmises that moments of increments (δur)pdelimited-⟨⟩superscript𝛿subscript𝑢𝑟𝑝\langle(\delta u_{r})^{p}\rangle⟨ ( italic_δ italic_u start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT ⟩, called structure functions, follow a universal power-law scaling in the so-called inertial-range:

Sp(r)(δur)prζp,ηrL,formulae-sequencesubscript𝑆𝑝𝑟delimited-⟨⟩superscript𝛿subscript𝑢𝑟𝑝similar-tosuperscript𝑟subscript𝜁𝑝much-less-than𝜂𝑟much-less-than𝐿\displaystyle S_{p}(r)\equiv\langle(\delta u_{r})^{p}\rangle\sim r^{\zeta_{p}}% \ ,\ \ \ \ \eta\ll r\ll L\ ,italic_S start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( italic_r ) ≡ ⟨ ( italic_δ italic_u start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT ⟩ ∼ italic_r start_POSTSUPERSCRIPT italic_ζ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_POSTSUPERSCRIPT , italic_η ≪ italic_r ≪ italic_L , (1)

where η𝜂\etaitalic_η is the viscous cutoff scale. Establishing such a simple scaling enables dramatic simplification in studying a wide range of turbulent flows, and thus, structure functions have been of persistent interest and a cornerstone of turbulence theory [5, 2, 3, 6]. K41 originally postulated ζp=p/3subscript𝜁𝑝𝑝3\zeta_{p}=p/3italic_ζ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = italic_p / 3; this result is known to be exact for p=3𝑝3p=3italic_p = 3, i.e., ζ3=1subscript𝜁31\zeta_{3}=1italic_ζ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = 1, but extensive studies from [7] to [8] (and others in between) have clearly established nonlinear deviations of ζpsubscript𝜁𝑝\zeta_{p}italic_ζ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT from p/3𝑝3p/3italic_p / 3 for p3𝑝3p\neq 3italic_p ≠ 3. This so-called anomalous scaling is attributed to the intermittency of interscale energy transfer processes (see, e.g., [5, 2, 3, 6]).

Since turbulence can also be fundamentally explored from a Lagrangian viewpoint [2, 9, 10, 11, 12], forceful arguments can be similarly made for Lagrangian velocity increments δuτ=u(t+τ)u(t)𝛿subscript𝑢𝜏𝑢𝑡𝜏𝑢𝑡\delta u_{\tau}=u(t+\tau)-u(t)italic_δ italic_u start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT = italic_u ( italic_t + italic_τ ) - italic_u ( italic_t ) over time lag τ𝜏\tauitalic_τ, measured along fluid-particle trajectories, and Lagrangian structure functions |δuτ|pdelimited-⟨⟩superscript𝛿subscript𝑢𝜏𝑝\langle|\delta u_{\tau}|^{p}\rangle⟨ | italic_δ italic_u start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT ⟩ defined therefrom 111Absolute value is taken for Lagrangian increments since their odd moments are zero. Extension of K41 phenomenology to Lagrangian increments gives:

SpL(τ)|δuτ|pτζpL,τητTLformulae-sequencesubscriptsuperscript𝑆𝐿𝑝𝜏delimited-⟨⟩superscript𝛿subscript𝑢𝜏𝑝similar-tosuperscript𝜏superscriptsubscript𝜁𝑝𝐿much-less-thansubscript𝜏𝜂𝜏much-less-thansubscript𝑇𝐿\displaystyle S^{L}_{p}(\tau)\equiv\langle|\delta u_{\tau}|^{p}\rangle\sim\tau% ^{\zeta_{p}^{L}}\ ,\ \ \ \ \tau_{\eta}\ll\tau\ll T_{L}italic_S start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( italic_τ ) ≡ ⟨ | italic_δ italic_u start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT ⟩ ∼ italic_τ start_POSTSUPERSCRIPT italic_ζ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT , italic_τ start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT ≪ italic_τ ≪ italic_T start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT (2)

where the temporal inertial-range is defined using TLsubscript𝑇𝐿T_{L}italic_T start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT, the Lagrangian integral time and τηsubscript𝜏𝜂\tau_{\eta}italic_τ start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT, the time-scale of viscous dissipation [2]. Since Lagrangian trajectories trace the underlying Eulerian field, it is natural to expect that a relation between Lagrangian and Eulerian exponents can be obtained.

Using K41, one obtains ζpL=p/2superscriptsubscript𝜁𝑝𝐿𝑝2\zeta_{p}^{L}=p/2italic_ζ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT = italic_p / 2 [2]; but, experimental and numerical studies again show nonlinear deviations from this prediction [14, 15, 16, 17]. Several attempts have been made [18, 19, 20] to quantify these deviations in terms of Eulerian intermittency, but they remain deficient for at least two reasons. First, the temporal scaling range in turbulence is substantially more restrictive than spatial scaling range [2, 3], making it difficult to robustly extract the Lagrangian scaling exponents. Second, past attempts have overwhelmingly focused on characterizing Lagrangian intermittency from longitudinal Eulerian intermittency, assuming that longitudinal and transverse exponents are identical, despite counter-evidence [21, 22, 23, 24, 25, 26].

In this Letter, presenting new data from direct numerical simulations (DNS) of isotropic turbulence at higher Reynolds numbers, we address both these challenges. We extract both Lagrangian and Eulerian scaling exponents. Our Eulerian results reaffirm recent results [8]. We then demonstrate an excellent correspondence between Lagrangian exponents and transverse Eulerian exponents, using as basis the same multifractal spectrum; this is different from the multifractal spectrum for longitudinal exponents, whose use in the past has failed to explain Lagrangian intermittency [27, 14, 15, 16, 17]).

Direct Numerical Simulations:

The description of DNS is necessarily brief here because they have been described in many recent works [28, 29, 30, 31, 32, 33]. The simulations correspond to the canonical setup of forced stationary isotropic turbulence in a triply periodic domain and are carried out using the highly accurate Fourier pseudo-spectral methods in space and second-order Runge-Kutta integration in time; the large scales are numerically forced to achieve statistical stationarity [34, 35]. A key feature of the present data is that we have achieved a wide range of Taylor-scale Reynolds number Rλsubscript𝑅𝜆R_{\lambda}italic_R start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT, going from 14013001401300140-1300140 - 1300 (on grids of up to 122883superscript12288312288^{3}12288 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT points) while maintaining excellent small-scale resolution [36, 29]. For Lagrangian statistics, a large population of fluid particles is tracked together with the Eulerian field. For Rλ650subscript𝑅𝜆650R_{\lambda}\leq 650italic_R start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ≤ 650, up to 64M particles are tracked for each case, whereas for Rλ=1300subscript𝑅𝜆1300R_{\lambda}=1300italic_R start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT = 1300, 256M particles are tracked (with M=10242absentsuperscript10242=1024^{2}= 1024 start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT) [37, 38, 39], providing ample statistics for convergence.

Refer to caption
Figure 1: Inertial-range scaling exponents for longitudinal and transverse Eulerian structure functions, the former from [8, 40] and the latter from the present data (consistent with [8]). Various theoretical predictions [1, 41, 42, 40] are also shown. The transverse exponents depart from all predictions and saturate.

Saturation of transverse exponents:

Anomalous scaling confers upon each moment order a separate and independent significance, instead of a mutual dependence (such as ζp=p/3subscript𝜁𝑝𝑝3\zeta_{p}=p/3italic_ζ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = italic_p / 3 based on K41). Multifractals have enjoyed considerable success in describing this behavior [3, 6], but lack any direct connection to Navier-Stokes equations. Further, recent DNS at high Rλsubscript𝑅𝜆R_{\lambda}italic_R start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT have shown noticeable departures of ζpsubscript𝜁𝑝\zeta_{p}italic_ζ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT from multifractal predictions for high orders [8]. Instead, starting from Navier-Stokes equations, a recent theory [40] was able to provide an improved prediction for ζpsubscript𝜁𝑝\zeta_{p}italic_ζ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT. Additionally, this theory also predicts that longitudinal exponents saturate with the moment-order, i.e., limpζpconstantsubscript𝑝subscript𝜁𝑝constant\lim\limits_{p\to\infty}\zeta_{p}\to\text{constant}roman_lim start_POSTSUBSCRIPT italic_p → ∞ end_POSTSUBSCRIPT italic_ζ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT → constant.

Recall that the transverse exponents are defined by the relation Sptrrζptrsimilar-tosuperscriptsubscript𝑆𝑝𝑡𝑟superscript𝑟superscriptsubscript𝜁𝑝𝑡𝑟S_{p}^{tr}\sim r^{\zeta_{p}^{tr}}italic_S start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t italic_r end_POSTSUPERSCRIPT ∼ italic_r start_POSTSUPERSCRIPT italic_ζ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t italic_r end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT, where Sptr(r)|δvr|psuperscriptsubscript𝑆𝑝𝑡𝑟𝑟delimited-⟨⟩superscript𝛿subscript𝑣𝑟𝑝S_{p}^{tr}(r)\equiv\langle|\delta v_{r}|^{p}\rangleitalic_S start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t italic_r end_POSTSUPERSCRIPT ( italic_r ) ≡ ⟨ | italic_δ italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT ⟩. (Absolute values are taken as the odd-moments are zero from symmetry.) Multifractal models based on phenomenological considerations do not differentiate between longitudinal and transverse exponents, i.e. ζ2ptr=ζ2psuperscriptsubscript𝜁2𝑝𝑡𝑟subscript𝜁2𝑝\zeta_{2p}^{tr}=\zeta_{2p}italic_ζ start_POSTSUBSCRIPT 2 italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t italic_r end_POSTSUPERSCRIPT = italic_ζ start_POSTSUBSCRIPT 2 italic_p end_POSTSUBSCRIPT, and general arguments have also been advanced to the same end [43, 44]. However, several studies have persistently pointed out that the two sets of exponents are different [21, 22, 23, 24, 25, 26]; recent work at high Rλsubscript𝑅𝜆R_{\lambda}italic_R start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT [8] has confirmed the differences, also showing that transverse exponents saturate: ζtr2.1superscriptsubscript𝜁𝑡𝑟2.1\zeta_{\infty}^{tr}\approx 2.1italic_ζ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t italic_r end_POSTSUPERSCRIPT ≈ 2.1 for p10𝑝10p\geq 10italic_p ≥ 10. Incidentally, this saturation is very different from ζ7.3subscript𝜁7.3\zeta_{\infty}\approx 7.3italic_ζ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT ≈ 7.3 (for p30𝑝30p\geq 30italic_p ≥ 30) predicted for longitudinal exponents in [40].

Refer to caption
Figure 2: Local slopes for (a) second and (b) fourth-order Lagrangian structure functions at various Rλsubscript𝑅𝜆R_{\lambda}italic_R start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT.

These findings are summarized in Fig. 1, showing the longitudinal and transverse exponents. Also included are K41 prediction, multifractal results [41, 42] and the result from [40]. Important considerations go into establishing the reliability of high-order exponents with respect to statistical convergence, adequacy of grid resolution, and Rλsubscript𝑅𝜆R_{\lambda}italic_R start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT-dependence. This discussion can be found in [8] and will not be repeated here. Instead, we focus on ζptrsuperscriptsubscript𝜁𝑝𝑡𝑟\zeta_{p}^{tr}italic_ζ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t italic_r end_POSTSUPERSCRIPT, which clearly depart from ζpsubscript𝜁𝑝\zeta_{p}italic_ζ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT and saturate for p10𝑝10p\geq 10italic_p ≥ 10. The implication of different longitudinal and transverse exponents for small-scale universality is discussed later; we first demonstrate how ζptrsuperscriptsubscript𝜁𝑝𝑡𝑟\zeta_{p}^{tr}italic_ζ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t italic_r end_POSTSUPERSCRIPT is directly related to the Lagrangian exponents.

Lagrangian exponents from DNS:

Robust extraction of scaling exponents requires sufficient scale separation to allow a proper inertial-range to exist. The Eulerian spatial scale separation for the highest Rλ=1300subscript𝑅𝜆1300R_{\lambda}=1300italic_R start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT = 1300 is L/η2500𝐿𝜂2500L/\eta\approx 2500italic_L / italic_η ≈ 2500 [8], while the temporal range is TL/τK105subscript𝑇𝐿subscript𝜏𝐾105T_{L}/\tau_{K}\approx 105italic_T start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT / italic_τ start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT ≈ 105 [45], thus making it inherently difficult to obtain a proper Lagrangian inertial-range [46, 47]. This difficulty is highlighted in Fig. 2, which shows the log local slope of SpL(τ)superscriptsubscript𝑆𝑝𝐿𝜏S_{p}^{L}(\tau)italic_S start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT ( italic_τ ) at various Rλsubscript𝑅𝜆R_{\lambda}italic_R start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT, for p=2𝑝2p=2italic_p = 2 and 4444 in panels (a) and (b), respectively; although there is a suggestion of a plateau for the fourth-order, the local slopes of the curves are still changing with Rλsubscript𝑅𝜆R_{\lambda}italic_R start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT. This is in contrast to the corresponding Eulerian result for p=2𝑝2p=2italic_p = 2, shown in Fig. 3, where a clear inertial-range emerges with Rλsubscript𝑅𝜆R_{\lambda}italic_R start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT.

Refer to caption
Figure 3: Local slopes for the Eulerian second-order structure functions at different Rλsubscript𝑅𝜆R_{\lambda}italic_R start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT. In contrast to Lagrangian data in Fig. 2, a clear inertial-range emerges with Reynolds number.

Because of this difficulty, Lagrangian exponents cannot be directly extracted even at the highest Rλsubscript𝑅𝜆R_{\lambda}italic_R start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT available. However, by using extended self-similarity [48], we can obtain the exponents with respect to the second-order [17]. Fig. 4 shows the ratio of local slope of SpL(τ)subscriptsuperscript𝑆𝐿𝑝𝜏S^{L}_{p}(\tau)italic_S start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( italic_τ ) to that of S2L(τ)subscriptsuperscript𝑆𝐿2𝜏S^{L}_{2}(\tau)italic_S start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_τ ). Evidently, a conspicuous plateau emerges for different orders in the same scaling range, seemingly independent of Rλsubscript𝑅𝜆R_{\lambda}italic_R start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT. Thus, we can extract the ratios ζpL/ζ2Lsubscriptsuperscript𝜁𝐿𝑝subscriptsuperscript𝜁𝐿2\zeta^{L}_{p}/\zeta^{L}_{2}italic_ζ start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT / italic_ζ start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, which also was the practice in earlier works [15, 16, 17]. The justification for using ζ2Lsuperscriptsubscript𝜁2𝐿\zeta_{2}^{L}italic_ζ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT as the reference comes from the expectation S2Lϵτsimilar-tosuperscriptsubscript𝑆2𝐿delimited-⟨⟩italic-ϵ𝜏S_{2}^{L}\sim\langle\epsilon\rangle\tauitalic_S start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT ∼ ⟨ italic_ϵ ⟩ italic_τ [2]; since the mean dissipation appears linearly, the result ζ2L=1superscriptsubscript𝜁2𝐿1\zeta_{2}^{L}=1italic_ζ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT = 1 is free of intermittency (akin to ζ3=1subscript𝜁31\zeta_{3}=1italic_ζ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = 1 for Eulerian exponents [49]).

Refer to caption
Figure 4: Ratio of local slope for p𝑝pitalic_p-th order Lagrangian structure function to second-order, for p=35𝑝35p=3-5italic_p = 3 - 5, at Rλ=1300subscript𝑅𝜆1300R_{\lambda}=1300italic_R start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT = 1300 (solid lines) and Rλ=650subscript𝑅𝜆650R_{\lambda}=650italic_R start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT = 650 (dashed lines).

Extending the procedure in Fig. 4, the ratios ζpL/ζ2Lsubscriptsuperscript𝜁𝐿𝑝subscriptsuperscript𝜁𝐿2\zeta^{L}_{p}/\zeta^{L}_{2}italic_ζ start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT / italic_ζ start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT are extracted for upto p=10𝑝10p=10italic_p = 10 and shown in Fig. 5. We also include earlier results from both experiments and DNS [20, 15, 16, 17], obtained at comparatively lower Rλsubscript𝑅𝜆R_{\lambda}italic_R start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT. Overall, the current results at higher Rλsubscript𝑅𝜆R_{\lambda}italic_R start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT are in excellent agreement with prior results (which had larger error bars). A remarkable result, endemic to all cases, is that the Lagrangian exponents saturate for p8greater-than-or-equivalent-to𝑝8p\gtrsim 8italic_p ≳ 8, similar to the transverse Eulerian exponents in Fig. 1. The data in Fig. 5 are also compared with various predictions, which we discuss next.

Refer to caption
Figure 5: Lagrangian scaling exponents and comparison with prior results and various multifractal models. The prediction from the transverse exponents is shown by the green curve that saturates for large p𝑝pitalic_p.

The multifractal framework:

Evidently, the data in Fig. 5 strongly deviate from K41. Following [3, 19], we will consider the well known multifractal model for relating Eulerian and Lagrangian exponents. The key concept in multifractals is that the (Eulerian) velocity increment δur𝛿subscript𝑢𝑟\delta u_{r}italic_δ italic_u start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT over a scale r𝑟ritalic_r is Hölder continuous, i.e., δurrhsimilar-to𝛿subscript𝑢𝑟superscript𝑟\delta u_{r}\sim r^{h}italic_δ italic_u start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ∼ italic_r start_POSTSUPERSCRIPT italic_h end_POSTSUPERSCRIPT, where hhitalic_h is the local Hölder exponent with the multifracal spectrum D(h)𝐷D(h)italic_D ( italic_h ) [50, 3]. From this local scaling, Eulerian structure functions are readily derived by integrating over all possible hhitalic_h, as (δur)phrph+3D(h)𝑑hsimilar-todelimited-⟨⟩superscript𝛿subscript𝑢𝑟𝑝subscriptsuperscript𝑟𝑝3𝐷differential-d\langle(\delta u_{r})^{p}\rangle\sim\int_{h}r^{ph+3-D(h)}dh⟨ ( italic_δ italic_u start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT ⟩ ∼ ∫ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT italic_r start_POSTSUPERSCRIPT italic_p italic_h + 3 - italic_D ( italic_h ) end_POSTSUPERSCRIPT italic_d italic_h. Using steepest-descent argument for rLmuch-less-than𝑟𝐿r\ll Litalic_r ≪ italic_L gives

ζp=infh[ph+3D(h)].subscript𝜁𝑝subscriptinfimumdelimited-[]𝑝3𝐷\displaystyle\zeta_{p}=\inf_{h}\left[ph+3-D(h)\right]\ .italic_ζ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = roman_inf start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT [ italic_p italic_h + 3 - italic_D ( italic_h ) ] . (3)

The Lagrangian extension of multifractals relies on the phenomenological assumption that spatial and temporal separations are interchangeable: rτδursimilar-to𝑟𝜏𝛿subscript𝑢𝑟r\sim\tau\delta u_{r}italic_r ∼ italic_τ italic_δ italic_u start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT, akin to frozen flow hypothesis, with δurδuτsimilar-to𝛿subscript𝑢𝑟𝛿subscript𝑢𝜏\delta u_{r}\sim\delta u_{\tau}italic_δ italic_u start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ∼ italic_δ italic_u start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT [19]. This stipulation gives δuττh/(1h)similar-to𝛿subscript𝑢𝜏superscript𝜏1\delta u_{\tau}\sim\tau^{h/(1-h)}italic_δ italic_u start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT ∼ italic_τ start_POSTSUPERSCRIPT italic_h / ( 1 - italic_h ) end_POSTSUPERSCRIPT, resulting in the Lagrangian exponents

ζpL=infh[ph+3D(h)1h].superscriptsubscript𝜁𝑝𝐿subscriptinfimumdelimited-[]𝑝3𝐷1\displaystyle\zeta_{p}^{L}=\inf_{h}\left[\frac{ph+3-D(h)}{1-h}\right]\ .italic_ζ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT = roman_inf start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT [ divide start_ARG italic_p italic_h + 3 - italic_D ( italic_h ) end_ARG start_ARG 1 - italic_h end_ARG ] . (4)

Thus, Lagrangian exponents can be directly predicted using the Eulerian multifractal spectrum D(h)𝐷D(h)italic_D ( italic_h ). Since past works have predominantly focused on Eulerian longitudinal exponents, with the implicit assumption that transverse exponents are same, the D(h)𝐷D(h)italic_D ( italic_h ) of the longitudinal exponents has been used to infer Lagrangian exponents. However, such predictions do not work as we see next.

The Lagrangian exponents can be computed from Eq. (4) by using Eulerian multifractal spectrum D(h)𝐷D(h)italic_D ( italic_h ) from Eq. (3). The D(h)𝐷D(h)italic_D ( italic_h ) corresponding to the Eulerian multifractal models shown in Fig. 1 are plotted in Fig. 6. They are obtained from ζpsubscript𝜁𝑝\zeta_{p}italic_ζ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT by taking a Legendre transform to invert the relations [3], giving

D(h)=infp[ph+3ζp].𝐷subscriptinfimum𝑝delimited-[]𝑝3subscript𝜁𝑝\displaystyle D(h)=\inf_{p}\left[ph+3-\zeta_{p}\right].italic_D ( italic_h ) = roman_inf start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT [ italic_p italic_h + 3 - italic_ζ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ] . (5)

For reference, the D(h)𝐷D(h)italic_D ( italic_h ) for She-Leveque model is [42]

D(h)=1+c1(hh*)c2(hh*)log(hh*)𝐷1subscript𝑐1superscriptsubscript𝑐2superscriptsuperscript\displaystyle D(h)=1+c_{1}(h-h^{*})-c_{2}(h-h^{*})\log(h-h^{*})italic_D ( italic_h ) = 1 + italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_h - italic_h start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ) - italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_h - italic_h start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ) roman_log ( italic_h - italic_h start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ) (6)

where h*=1/9superscript19h^{*}=1/9italic_h start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT = 1 / 9, c1=c2(1+loglogγlogγ)subscript𝑐1subscript𝑐21𝛾𝛾c_{1}=c_{2}(1+\log\log\gamma-\log\gamma)italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( 1 + roman_log roman_log italic_γ - roman_log italic_γ ) and c2=3/logγsubscript𝑐23𝛾c_{2}=3/\log\gammaitalic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 3 / roman_log italic_γ, with γ=3/2𝛾32\gamma=3/2italic_γ = 3 / 2. That for the Sreenivasan-Yakhot result of ζp=ζp/(p+β)subscript𝜁𝑝subscript𝜁𝑝𝑝𝛽\zeta_{p}=\zeta_{\infty}p/(p+\beta)italic_ζ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = italic_ζ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT italic_p / ( italic_p + italic_β ) [40] is

D(h)=3ζβh+2ζβh𝐷3subscript𝜁𝛽2subscript𝜁𝛽\displaystyle D(h)=3-\zeta_{\infty}-\beta h+2\sqrt{\zeta_{\infty}\beta h}italic_D ( italic_h ) = 3 - italic_ζ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT - italic_β italic_h + 2 square-root start_ARG italic_ζ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT italic_β italic_h end_ARG (7)

where ζ7.3subscript𝜁7.3\zeta_{\infty}\approx 7.3italic_ζ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT ≈ 7.3 and β=3ζ3𝛽3subscript𝜁3\beta=3\zeta_{\infty}-3italic_β = 3 italic_ζ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT - 3. The result for p-model can be found in [41].

In Fig. 6, in addition to the D(h)𝐷D(h)italic_D ( italic_h ) from these known Eulerian cases, we also utilize Eq. (5) to numerically obtain the D(h)𝐷D(h)italic_D ( italic_h ) for transverse exponents (with ζptr2.1superscriptsubscript𝜁𝑝𝑡𝑟2.1\zeta_{p}^{tr}\approx 2.1italic_ζ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t italic_r end_POSTSUPERSCRIPT ≈ 2.1 for p10𝑝10p\geq 10italic_p ≥ 10, as shown in Fig. 1). Note, since the D(h)𝐷D(h)italic_D ( italic_h ) for ζptrsuperscriptsubscript𝜁𝑝𝑡𝑟\zeta_{p}^{tr}italic_ζ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t italic_r end_POSTSUPERSCRIPT is obtained numerically, the inversion formula in Eq. 5 can only provide the concave hull [3]—which is what we plot in Fig. 6. The saturation value of exponents is reflected in the corresponding D(h)𝐷D(h)italic_D ( italic_h ) curve for h=00h=0italic_h = 0, as D(0)=3ζ𝐷03subscript𝜁D(0)=3-\zeta_{\infty}italic_D ( 0 ) = 3 - italic_ζ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT (0.9absent0.9\approx 0.9≈ 0.9 for ζtr2.1superscriptsubscript𝜁𝑡𝑟2.1\zeta_{\infty}^{tr}\approx 2.1italic_ζ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t italic_r end_POSTSUPERSCRIPT ≈ 2.1). Note, h<00h<0italic_h < 0 is not allowed in the multifractal framework [3]; the p-model and She-Leveque results respectively correspond to hmin=13log2(0.7)0.172subscriptmin13subscript20.70.172h_{\rm min}=\frac{1}{3}\log_{2}(0.7)\approx 0.172italic_h start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 3 end_ARG roman_log start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( 0.7 ) ≈ 0.172 [41] and hmin=h*=19subscriptminsuperscript19h_{\rm min}=h^{*}=\frac{1}{9}italic_h start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT = italic_h start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT = divide start_ARG 1 end_ARG start_ARG 9 end_ARG [42], which preclude saturation. The Sreenivasan-Yakhot result [40] predicts saturation for longitudinal exponents (at ζ7.3subscript𝜁7.3\zeta_{\infty}\approx 7.3italic_ζ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT ≈ 7.3, giving D(0)=37.3=4.3𝐷037.34.3D(0)=3-7.3=-4.3italic_D ( 0 ) = 3 - 7.3 = - 4.3 (not shown in in Fig. 6).

Refer to caption
Figure 6: The multifractal spectra for various models. The vertical dashed lines at 13log2(0.7)(0.17)annotated13subscript20.7absent0.17\frac{1}{3}\log_{2}(0.7)(\approx 0.17)divide start_ARG 1 end_ARG start_ARG 3 end_ARG roman_log start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( 0.7 ) ( ≈ 0.17 ) and 1919\frac{1}{9}divide start_ARG 1 end_ARG start_ARG 9 end_ARG mark the minimum hhitalic_h allowed for p-model [41] She-Leveque [42], respectively, which preclude saturation; whereas D(h=0)32.1=0.9𝐷032.10.9D(h=0)\approx 3-2.1=0.9italic_D ( italic_h = 0 ) ≈ 3 - 2.1 = 0.9 marks saturation for transverse exponents at ζtr2.1superscriptsubscript𝜁𝑡𝑟2.1\zeta_{\infty}^{tr}\approx 2.1italic_ζ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t italic_r end_POSTSUPERSCRIPT ≈ 2.1.

Lagrangian exponents from the transverse multifractal spectrum:

As we saw, none of the multifractal predictions for Lagrangian exponents using Eulerian longitudinal exponents agree with the data. In contrast, the prediction corresponding to transverse Eulerian exponent (green dot-dashed line in Fig. 5) closely follows the measured results, particularly capturing the saturation at high orders. Note, the predicted saturation value ζL2.1superscriptsubscript𝜁𝐿2.1\zeta_{\infty}^{L}\approx 2.1italic_ζ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT ≈ 2.1, is the same for both transverse Eulerian and Lagrangian exponents, The actual Lagrangian data saturate at a very slightly smaller value. We believe this minor difference (of only 5%) stems from the fact that even at Rλ=1300subscript𝑅𝜆1300R_{\lambda}=1300italic_R start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT = 1300, the temporal inertial-range is underdeveloped, and the intermittency-free result of ζ2L=1superscriptsubscript𝜁2𝐿1\zeta_{2}^{L}=1italic_ζ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT = 1 is not unambiguously realized. Since Lagrangian exponents shown in Fig. 5 are extracted as ratios ζpL/ζ2Lsuperscriptsubscript𝜁𝑝𝐿superscriptsubscript𝜁2𝐿\zeta_{p}^{L}/\zeta_{2}^{L}italic_ζ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT / italic_ζ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT, this minor discrepancy in the saturation values could be explained by small departures from the expectation of ζ2L=1superscriptsubscript𝜁2𝐿1\zeta_{2}^{L}=1italic_ζ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT = 1. Given this and also possible statistical uncertainties (at highest orders), the close correspondence between the transverse Eulerian exponents and Lagrangian exponents is quite remarkable.

It is worth noting that Lagrangian exponents saturate for slightly smaller p𝑝pitalic_p than for transverse Eulerian exponents. This readily follows from Eqs. (3)-(4) as a kinematic effect. For Eulerian exponents, ζ3=1subscript𝜁31\zeta_{3}=1italic_ζ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = 1 is exact, corresponding to h1313h\approx\frac{1}{3}italic_h ≈ divide start_ARG 1 end_ARG start_ARG 3 end_ARG, D(h)3𝐷3D(h)\approx 3italic_D ( italic_h ) ≈ 3, which conforms to the intermittency-free K41 result [3]. This gives ζ2L=1superscriptsubscript𝜁2𝐿1\zeta_{2}^{L}=1italic_ζ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT = 1 as the corresponding Lagrangian result for h1313h\approx\frac{1}{3}italic_h ≈ divide start_ARG 1 end_ARG start_ARG 3 end_ARG, D(h)3𝐷3D(h)\approx 3italic_D ( italic_h ) ≈ 3. This argument can be extended to higher orders to show that Lagrangian exponents at order p𝑝pitalic_p correspond to transverse exponents at order 3p/23𝑝23p/23 italic_p / 2. Thus, it follows that Lagrangian exponents saturate at smaller p𝑝pitalic_p. A similar correspondence can also be provided for other Lagrangian statistics, for instance, the second-moment of acceleration (the temporal velocity-gradient) corresponds to the third-moment of spatial velocity-gradients [2, 27].

Discussion:

Two significant results emerge from our work: (a) scaling exponents saturate for both transverse Eulerian and Lagrangian structure functions; and (b) the saturation of Lagrangian exponents is characterized solely by the transverse Eulerian exponents (and not the longitudinal, as previously believed). Given that the transverse exponents are smaller for large p𝑝pitalic_p, this seems reasonable from the steepest-descent argument [3].

The saturation of scaling exponents is extreme form of anomalous behavior, but is not uncommon; it holds for Burgers equation [51], passive scalar turbulence [52, 53, 54]. However, its prevalence in velocity field has become apparent only recently [8, 40]. The theory of [40] predicts that Eulerian longitudinal exponents saturate as well, although at very high moment orders that cannot be yet validated. In contrast, both transverse Eulerian exponents and Lagrangian exponents saturate and at the same value of 2absent2\approx 2≈ 2. Further, using a simple physical correspondence based on frozen flow hypothesis, they are related through the same multifractal spectrum (which differs from known spectrum for longitudinal Eulerian exponents). Interestingly, the saturation exponent of 2 implies a fractal co-dimension of 1 [3, 6], suggesting that the saturation likely comes from localized (very) thin vortex filaments, which are known to be prevalent at the smallest scales [55, 34, 36].

Our results also bring forth some important questions. First is the extension of the multifractals from inertial- to dissipative-range, i.e., describing the scaling of velocity-gradients. Such an extension relies on the phenomenological criterion that the local Reynolds number, describing the dissipative cutoff, is unity, i.e., δurr/ν=1𝛿subscript𝑢𝑟𝑟𝜈1\delta u_{r}r/\nu=1italic_δ italic_u start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT italic_r / italic_ν = 1 [56, 3, 40]. As highlighted in recent works [36, 57], this is valid for longitudinal increments, but not for transverse increments, essentially because of how vorticity and strain-rate interact in turbulence. It can thus be expected that the extension of multifractals to dissipation-range works for longitudinal velocity-gradients, but not for transverse velocity-gradients. Since the current results suggest that Lagrangian intermittency is linked to transverse Eulerian intermittency, it follows that the extension to acceleration statistics would be an issue, as confirmed by our recent studies [27, 58]. In addition, acceleration components are strongly correlated in turbulence [59, 58], which is a feature of Navier-Stokes dynamics that is not accounted for by multifractals.

A second question concerns the meaning of universality given the longitudinal and transverse exponents behave differently. One strategy could be to consider a joint multifractal spectrum for longitudinal and transverse increments. It might be possible to set appropriate conditions on both to enable the inertial-range universality and the transition from the inertial- to dissipation-range. Essentially, addressing the discrepancy between longitudinal and transverse intermittency presents a critical and pressing problem in turbulence theory.

Acknowledgements.

Acknowledgments:

We gratefully acknowledge discussions with Victor Yakhot and sustained collaboration with P.K. Yeung. We also gratefully acknowledge the Gauss Centre for Supercomputing e.V. (www.gauss-centre.eu) for providing computing time on the supercomputers JUQUEEN and JUWELS at Jülich Supercomputing Centre (JSC), where the simulations reported in this paper were primarily performed. Computations were also supported partially by the supercomputing resources under the Blue Water project at the National Center for Supercomputing Applications at the University of Illinois (Urbana-Champaign).

References

  • Kolmogorov [1941a] A. N. Kolmogorov, The local structure of turbulence in an incompressible fluid for very large Reynolds numbers, Dokl. Akad. Nauk. SSSR 30, 299 (1941a).
  • Monin and Yaglom [1975] A. S. Monin and A. M. Yaglom, Statistical Fluid Mechanics, Vol. 2 (MIT Press, 1975).
  • Frisch [1995] U. Frisch, Turbulence: the legacy of Kolmogorov (Cambridge University Press, Cambridge, 1995).
  • Pope [2000] S. B. Pope, Turbulent Flows (Cambridge University Press, 2000).
  • Kolmogorov [1962] A. N. Kolmogorov, A refinement of previous hypotheses concerning the local structure of turbulence in a viscous incompressible fluid at high Reynolds number, J. Fluid Mech. 13, 82 (1962).
  • Sreenivasan and Antonia [1997] K. R. Sreenivasan and R. A. Antonia, The phenomenology of small-scale turbulence, Annu. Rev. Fluid Mech. 29, 435 (1997).
  • Van Atta and Park [2005] C. W. Van Atta and J. Park, Statistical self-similarity and inertial subrange turbulence, in Statistical Models and Turbulence: Proceedings of a Symposium held at the University of California, San Diego (La Jolla) July 15–21, 1971 (Springer, 2005) pp. 402–426.
  • Iyer et al. [2020] K. P. Iyer, K. R. Sreenivasan, and P. K. Yeung, Scaling exponents saturate in three-dimensional isotropic turbulence, Phys. Rev. Fluids 5, 054605 (2020).
  • Wyngaard [1992] J. C. Wyngaard, Atmospheric turbulence, Annu. Rev. Fluid Mech. 24, 205 (1992).
  • Sawford [2001] B. L. Sawford, Turbulent relative dispersion, Annu. Rev. Fluid Mech. 33, 289 (2001).
  • Falkovich et al. [2001] G. Falkovich, K. Gawȩdzki, and M. Vergassola, Particles and fields in fluid turbulence, Rev. Mod. Phys. 73, 913 (2001).
  • Toschi and Bodenschatz [2009] F. Toschi and E. Bodenschatz, Lagrangian properties of particles in turbulence, Annu. Rev. Fluid Mech. 41, 375 (2009).
  • Note [1] Absolute value is taken for Lagrangian increments since the odd moments are otherwise zero.
  • Sawford et al. [2003] B. L. Sawford, P. K. Yeung, M. S. Borgas, P. Vedula, A. L. Porta, A. M. Crawford, and E. Bodenschatz, Conditional and unconditional acceleration statistics in turbulence, Phys. Fluids 15, 3478 (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).
  • Xu et al. [2006] H. Xu, M. Bourgoin, N. T. Ouellette, and E. Bodenschatz (International Collaboration for Turbulence Research), High order lagrangian velocity statistics in turbulence, Phys. Rev. Lett. 96, 024503 (2006).
  • Sawford and Yeung [2015] B. L. Sawford and P. K. Yeung, Direct numerical simulation studies of Lagrangian intermittency in turbulence, Phys. Fluids 27, 065109 (2015).
  • Borgas [1993] M. S. Borgas, The multifractal Lagrangian nature of turbulence, Philos. Trans. R. Soc. A 342, 379 (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).
  • Arnéodo et al. [2008] A. Arnéodo et al., Universal intermittent properties of particle trajectories in highly turbulent flows, Phys. Rev. Lett. 100, 254504 (2008).
  • Dhruva et al. [1997] B. Dhruva, Y. Tsuji, and K. R. Sreenivasan, Transverse structure functions in high-reynolds-number turbulence, Phys. Rev. E 56, R4928 (1997).
  • Chen et al. [1997] S. Chen, K. R. Sreenivasan, M. Nelkin, and N. Cao, Refined similarity hypothesis for transverse structure functions in fluid turbulence, Phys. Rev. Lett. 79, 2253 (1997).
  • Grossmann et al. [1997] S. Grossmann, D. Lohse, and A. Reeh, Different intermittency for longitudinal and transversal turbulent fluctuations, Physics of Fluids 9, 3817 (1997).
  • Shen and Warhaft [2002] X. Shen and Z. Warhaft, Longitudinal and transverse structure functions in sheared and unsheared wind-tunnel turbulence, Phys. Fluids 14, 370 (2002).
  • Gotoh et al. [2002] T. Gotoh, D. Fukayama, and T. Nakano, Velocity field statistics in homogeneous steady turbulence obtained using a high-resolution direct numerical simulation, Phys. Fluids 14, 1065 (2002).
  • Grauer et al. [2012] R. Grauer, H. Homann, and J.-F. Pinton, Longitudinal and transverse structure functions in high-Reynolds-number turbulence, New J. Phys. 14, 063016 (2012).
  • 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).
  • 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. [2020] D. Buaria, E. Bodenschatz, and A. Pumir, Vortex stretching and enstrophy production in high Reynolds number turbulence, Phys. Rev. Fluids 5, 104602 (2020).
  • Buaria and Pumir [2021] D. Buaria and A. Pumir, Nonlocal amplification of intense vorticity in turbulent flows, Phys. Rev. Research 3, 042020 (2021).
  • 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 and Sreenivasan [2023a] D. Buaria and K. R. Sreenivasan, Forecasting small-scale dynamics of fluid turbulence using deep neural networks, Proc. Nat. Acad. Sci. 120, e2305765120 (2023a).
  • Ishihara et al. [2009] T. Ishihara, T. Gotoh, and Y. Kaneda, Study of high-Reynolds number isotropic turbulence by direct numerical simulations, Annu. Rev. Fluid Mech. 41, 165 (2009).
  • Rogallo [1981] R. S. Rogallo, Numerical experiments in homogeneous turbulence, NASA Technical Memo  (1981).
  • 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).
  • 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 (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 (2017).
  • Sreenivasan and Yakhot [2021] K. R. Sreenivasan and V. Yakhot, Dynamics of three-dimensional turbulence from Navier-Stokes equations, Phys. Rev. Fluids 6, 104604 (2021).
  • Meneveau and Sreenivasan [1987] C. Meneveau and K. R. Sreenivasan, Simple multifractal cascade model for fully developed turbulence, Phys. Rev. Lett. 59, 1424 (1987).
  • She and Leveque [1994] Z.-S. She and E. Leveque, Universal scaling laws in fully developed turbulence, Phys. Rev. Lett. 72, 336 (1994).
  • L’vov et al. [1997] V. S. L’vov, E. Podivilov, and I. Procaccia, Invariants for correlations of velocity differences in turbulent fields, Phys. Rev. Lett. 79, 2050 (1997).
  • Nelkin [1990] M. Nelkin, Multifractal scaling of velocity derivatives in turbulence, Phys. Rev. A 42, 7226 (1990).
  • Buaria [2016] D. Buaria, Lagrangian investigations of turbulent dispersion and mixing using Petascale computing, Ph.D. thesis, Georgia Institute of Technology (2016).
  • Sawford and Yeung [2011] B. L. Sawford and P. K. Yeung, Kolmogorov similarity scaling for one-particle lagrangian statistics, Phys. Fluids 23 (2011).
  • Buaria [2023] D. Buaria, Comment on “Universality and Intermittency of Pair Dispersion in Turbulence”, Phys. Rev. Lett. 130, 029401 (2023).
  • Benzi et al. [1993] R. Benzi, S. Ciliberto, R. Tripiccione, C. Baudet, F. Massaioli, and S. Succi, Extended self-similarity in turbulent flows, Phys. Rev. E 48, R29 (1993).
  • Kolmogorov [1941b] A. N. Kolmogorov, Dissipation of energy in locally isotropic turbulence, Dokl. Akad. Nauk. SSSR 434, 16 (1941b).
  • Benzi et al. [1984] R. Benzi, G. Paladin, G. Parisi, and A. Vulpiani, On the multifractal nature of fully developed turbulence and chaotic systems, J. Phys. A 17, 3521 (1984).
  • Bec and Khanin [2007] J. Bec and K. Khanin, Burgers turbulence, Phys. Rev. 447, 1 (2007).
  • Kraichnan [1994] R. H. Kraichnan, Anomalous scaling of a randomly advected passive scalar, Phys. Rev. Lett. 72, 1016 (1994).
  • Iyer et al. [2018] K. P. Iyer, J. Schumacher, K. R. Sreenivasan, and P. K. Yeung, Steep cliffs and saturated exponents in three-dimensional scalar turbulence, Phys. Rev. Lett. 121, 264501 (2018).
  • Buaria et al. [2021] D. Buaria, M. P. Clay, K. R. Sreenivasan, and P. K. Yeung, Turbulence is an ineffective mixer when Schmidt numbers are large, Phys. Rev. Lett. 126, 074501 (2021).
  • Jiménez et al. [1993] J. Jiménez, A. A. Wray, P. G. Saffman, and R. S. Rogallo, The structure of intense vorticity in isotropic turbulence, J. Fluid Mech. 255 (1993).
  • Paladin and Vulpiani [1987] G. Paladin and A. Vulpiani, Degrees of freedom of turbulence, Phys. Rev. A 35, 1971 (1987).
  • 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 and Sreenivasan [2023b] D. Buaria and K. R. Sreenivasan, Lagrangian acceleration and its Eulerian decompositions in fully developed turbulence, Phys. Rev. Fluids 8, L032601 (2023b).
  • 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 (2001).