A unified view of elastic and elasto-inertial turbulence in channel flows

Giulio Foggi Rota ID 1    Christian Amor ID 1    Soledad Le Clainche ID 2    Marco Edoardo Rosti ID 1 marco.rosti@oist.jp 1 Complex Fluids and Flows Unit, Okinawa Institute of Science and Technology Graduate University, 1919-1 Tancha, Onna-son, Okinawa 904-0495, Japan.
2 School of Aerospace Engineering, Universidad Politécnica de Madrid, E-28040 Madrid, Spain.
Abstract

The chaotic flow of elastic fluids at low Reynolds number (Re𝑅𝑒Reitalic_R italic_e) is typically distinguished into elasto-inertial and elastic turbulence (EIT/ET). However, the clear separation among these two turbulent regimes in parallel flows with a gradual Re𝑅𝑒Reitalic_R italic_e decrease remains elusive. Here, spanning Re𝑅𝑒Reitalic_R italic_e over four orders of magnitude, we investigate the statistical and structural nature of the flows computed by fully resolved 3D numerical simulations, revealing that as inertia becomes sub-dominant all flows share similar dynamical properties. Supported by the results, we thus propose a unified view of fully developed ET and EIT in parallel flows.

preprint: APS/123-QED

Turbulence is an open problem in fluid transport applications, as it significantly enhances the drag and energy expenditures. Nevertheless, it is beneficial in mixing problems since it introduces an additional diffusion mechanism on top of the classical viscous one. Indeed, viscous diffusion is often too slow for practical applications, but it is the only mechanism operating in low Reynolds number flows, which are typically laminar. The Reynolds number, Re𝑅𝑒Reitalic_R italic_e, is the non dimensional parameter conventionally adopted to describe fluid flow, representing the relative importance of inertia compared to viscosity. Micro-scale mixing in low Reynolds number flows is a widely investigated research topic Stroock et al. (2002); Ober et al. (2015); Han et al. (2022) for its high biological relevance, e.g., in drug delivery applications Vargason et al. (2021) and in the analysis of specimens for disease diagnosis Yager et al. (2006). Among the various mixing techniques, one is to exploit elastic turbulence Groisman and Steinberg (2000), a peculiar chaotic state unique of viscoelastic fluids which are characterised by the coexistence of viscous and elastic effects.

The description of viscoelastic fluid flows necessitates of an additional parameter accounting for the ratio of the elastic and flow time scales: the Deborah number, De𝐷𝑒Deitalic_D italic_e. Comparing the Deborah to the Reynolds number allows to identify distinct regimes of motion: when inertial effects are dominant turbulent drag reduction Serafini et al. (2022) (TDR) occurs, while elastic turbulence Groisman and Steinberg (2000) (ET) is observed for dominant elasticity. In between these two extrema, an intermediate state dubbed elasto-inertial turbulence Dubief et al. (2023) (EIT) has been usually identified, where both elastic and inertial effects are relevant. There is no universal consensus on the definitions of EIT and ET; yet, for the purposes of our investigation, we follow Groisman and Steinberg (2000) and denote with ET the chaotic flow observed in polymer solutions at small values of Re𝑅𝑒Reitalic_R italic_e and large values of De𝐷𝑒Deitalic_D italic_e (ReDeless-than-or-similar-to𝑅𝑒𝐷𝑒Re\lesssim Deitalic_R italic_e ≲ italic_D italic_e). Furthermore, according to Samanta et al. (2013), we dub EIT the chaotic motion that sets in at lower Reynolds number than Newtonian turbulence due to the action of fluid elasticity, in a parameter range where inertia cannot be neglected: when Re𝑅𝑒Reitalic_R italic_e and De𝐷𝑒Deitalic_D italic_e are comparable (ReDe𝑅𝑒𝐷𝑒Re\approx Deitalic_R italic_e ≈ italic_D italic_e).

Refer to caption
Figure 1: Side views of the channel, with the mean flow going from left to right. We show the spanwise velocity fluctuations normalized with the mean turbulent kinetic energy, w/Ksuperscript𝑤𝐾w^{\prime}/\sqrt{K}italic_w start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT / square-root start_ARG italic_K end_ARG, ranging from -0.3 (blue) to 0.3 (green); superimposed are the local isosurfaces of the wall-normal fluctuations, v/Ksuperscript𝑣𝐾v^{\prime}/\sqrt{K}italic_v start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT / square-root start_ARG italic_K end_ARG, at 0.050.05-0.05- 0.05 (orange) and +0.050.05+0.05+ 0.05 (violet).

EIT occurs at significantly lower Reynolds numbers than TDR, where the laminar flow of a Newtonian fluid would be linearly stable, and it is characterised by different dynamical properties Samanta et al. (2013). When inertial effects vanish, the flow can still become turbulent (ET) due to the onset of different elastic instabilities. It is in fact well established that different instability mechanisms are responsible for the transition to a chaotic state at moderate Garg et al. (2018) and vanishing Berti and Boffetta (2010) values of Re𝑅𝑒Reitalic_R italic_e, thus motivating the distinction between EIT and ET in the stability framework Dubief et al. (2023). In curvilinear wall bounded flows, ET is triggered and sustained by a force directed as the flow curvature, called hoop stress Bird (1987); Groisman and Steinberg (2001). Recent experiments Jha and Steinberg (2021); Li and Steinberg (2023) and numerical simulations Song et al. (2022); Lellep et al. (2024) have highlighted that ET can also be attained in parallel shear flows, but a comprehensive characterisation of its mechanisms is still elusive. After the discovery of a linear elastic instability in the parallel Kolmogorov flow Boffetta et al. (2005), the weakly non-linear regime following the instability has been studied Bistagnino et al. (2007) and observed to culminate in ET Berti et al. (2008). Arrowhead-shaped traveling elastic waves were first observed in this scenario Berti and Boffetta (2010) and later found also in the bulk of channel flows in the EIT Sid et al. (2018); Garg et al. (2018); Dubief et al. (2022); Buza et al. (2022a) and ET regimes Morozov (2022), suggesting a possible continuous path between the two Dubief et al. (2023) and hinting towards the existence of a unique sustainment mechanism common to EIT and ET Khalid et al. (2021a). Channel flows at moderate to vanishing inertia have been recently observed to share also a different instability Beneitez et al. (2023), localised close to the walls. Nevertheless, due to a lack of measurements, it is still unclear if the fully developed three-dimensional EIT and ET chaotic states are the same, connected or completely decoupled.

In this Letter we show that fully developed EIT and ET in planar channel flows share a similar structure, notwithstanding the instabilities leading to them. Addressing the fundamental question 111“Is EIT simply an inertial version of ET or does inertia play a more fundamental role?” posed by Dubief et al. (2023), we thus clarify the relation between EIT and ET within this setup. Our fully-resolved three-dimensional numerical simulations, spanning Reynolds numbers over four orders of magnitude (as in FIG. 1), allow us to bridge the different regimes (TDR, EIT and ET) and clarify their relation. For every Reynolds number considered in this study we observe the system to converge towards a self-sustained chaotic state even when the flow of a Newtonian fluid would instead be laminar, and we identify only two different regimes, one at large Reynolds number (TDR) and one at smaller ones (ET).

To show this, we simulate the flow of a viscoelastic incompressible fluid between two indefinite planar walls at a distance of 2h22h2 italic_h from each other, described in a Cartesian coordinate frame with the x𝑥xitalic_x, y𝑦yitalic_y and z𝑧zitalic_z axes aligned with the streamwise, wall-normal and spanwise directions, respectively. The system is governed by the conservation of mass and momentum,

𝐮=0,𝐮0\mathbf{\nabla}\cdot\mathbf{u}=0,∇ ⋅ bold_u = 0 , (1)
ρ(t𝐮+𝐮𝐮)=p+μs(𝐮+(𝐮)T)+ρ𝐟+𝐓,𝜌subscript𝑡𝐮𝐮𝐮𝑝subscript𝜇𝑠𝐮superscript𝐮𝑇𝜌𝐟𝐓\rho\left(\partial_{t}\mathbf{u}+\mathbf{\nabla}\cdot\mathbf{u}\mathbf{u}% \right)=-\mathbf{\nabla}p+\mu_{s}\mathbf{\nabla}\cdot(\mathbf{\nabla u}+(% \mathbf{\nabla u})^{T})+\rho\mathbf{f}+\mathbf{\nabla}\cdot\mathbf{T},italic_ρ ( ∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT bold_u + ∇ ⋅ bold_uu ) = - ∇ italic_p + italic_μ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ∇ ⋅ ( ∇ bold_u + ( ∇ bold_u ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT ) + italic_ρ bold_f + ∇ ⋅ bold_T , (2)

where 𝐮𝐮\mathbf{u}bold_u is the velocity, ρ𝜌\rhoitalic_ρ the density, p𝑝pitalic_p the pressure and μssubscript𝜇𝑠\mu_{s}italic_μ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT is the dynamic viscosity of the Newtonian solvent in which the polymer molecules are dissolved. 𝐟𝐟\mathbf{f}bold_f represents the forcing term needed to drive the flow at a constant flow rate, while 𝐓𝐓\mathbf{T}bold_T is the tensor accounting for the non-Newtonian behaviour. In our study, 𝐓𝐓\mathbf{T}bold_T derives from a finitely extensible nonlinear elastic model with Peterlin’s closure (FENE-P Peterlin (1966), equivalent to a bead-spring-type model Vincenzi et al. (2007)) and relates to the polymer conformation tensor, 𝐂𝐂\mathbf{C}bold_C, as 𝐓(𝐂)=(μp/λ)(f(tr(𝐂))𝐂(Lmax2𝐈)/(Lmax23))𝐓𝐂subscript𝜇𝑝𝜆𝑓𝑡𝑟𝐂𝐂superscriptsubscript𝐿𝑚𝑎𝑥2𝐈superscriptsubscript𝐿𝑚𝑎𝑥23\mathbf{T}(\mathbf{C})=(\mu_{p}/\lambda)(f(tr(\mathbf{C}))\mathbf{C}-(L_{max}^% {2}\mathbf{I})/(L_{max}^{2}-3))bold_T ( bold_C ) = ( italic_μ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT / italic_λ ) ( italic_f ( italic_t italic_r ( bold_C ) ) bold_C - ( italic_L start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_I ) / ( italic_L start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 3 ) ), where Lmaxsubscript𝐿𝑚𝑎𝑥L_{max}italic_L start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT denotes the maximum extensibility of the polymer chains, 𝐈𝐈\mathbf{I}bold_I is the tensorial identity, μpsubscript𝜇𝑝\mu_{p}italic_μ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT is the dynamic viscosity of the polymer, and λ𝜆\lambdaitalic_λ is its relaxation time. In particular, f(s)=Lmax2/(Lmax2s)𝑓𝑠superscriptsubscript𝐿𝑚𝑎𝑥2superscriptsubscript𝐿𝑚𝑎𝑥2𝑠f(s)=L_{max}^{2}/(L_{max}^{2}-s)italic_f ( italic_s ) = italic_L start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( italic_L start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_s ). Finally, 𝐂𝐂\mathbf{C}bold_C evolves in time according to its own transport equation

t𝐂+(𝐮)𝐂𝐂𝐮(𝐮)T𝐂=1λ(f(tr(𝐂))𝐂𝐈).subscript𝑡𝐂𝐮𝐂𝐂𝐮superscript𝐮𝑇𝐂1𝜆𝑓𝑡𝑟𝐂𝐂𝐈\partial_{t}\mathbf{C}+(\mathbf{u}\cdot\mathbf{\nabla})\mathbf{C}-\mathbf{C}% \cdot\mathbf{\nabla u}-(\mathbf{\nabla u})^{T}\cdot\mathbf{C}=-\displaystyle% \frac{1}{\lambda}(f(tr(\mathbf{C}))\mathbf{C}-\mathbf{I}).∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT bold_C + ( bold_u ⋅ ∇ ) bold_C - bold_C ⋅ ∇ bold_u - ( ∇ bold_u ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT ⋅ bold_C = - divide start_ARG 1 end_ARG start_ARG italic_λ end_ARG ( italic_f ( italic_t italic_r ( bold_C ) ) bold_C - bold_I ) . (3)

The regime of motion of the viscoelastic fluid is identified by the specification of the Reynolds and the Deborah numbers, along with the solvent-to-total viscosity ratio, β𝛽\betaitalic_β, and Lmaxsubscript𝐿𝑚𝑎𝑥L_{max}italic_L start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT. In this work we adopt the definition of “bulk” Reynolds number, Re=hU/ν𝑅𝑒𝑈𝜈Re=hU/\nuitalic_R italic_e = italic_h italic_U / italic_ν, based on the mean velocity within the channel, U𝑈Uitalic_U, and the total kinematic viscosity of the polymeric solution, ν=(μs+μp)/ρ𝜈subscript𝜇𝑠subscript𝜇𝑝𝜌\nu=(\mu_{s}+\mu_{p})/\rhoitalic_ν = ( italic_μ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT + italic_μ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ) / italic_ρ; we therefore decrease Re𝑅𝑒Reitalic_R italic_e in {2800,1500,500,50,5,0.5}280015005005050.5\{2800,1500,500,50,5,0.5\}{ 2800 , 1500 , 500 , 50 , 5 , 0.5 }. Furthermore, we complement our results with those available in the literature by Lellep et al. (2024) at even smaller Reynolds number, Re=0.01𝑅𝑒0.01Re=0.01italic_R italic_e = 0.01. The Deborah number is defined as De=λU/h𝐷𝑒𝜆𝑈De=\lambda U/hitalic_D italic_e = italic_λ italic_U / italic_h and imposed equal to 50505050 throughout the whole study, while Lmaxsubscript𝐿𝑚𝑎𝑥L_{max}italic_L start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT and β=μs/(μs+μp)𝛽subscript𝜇𝑠subscript𝜇𝑠subscript𝜇𝑝\beta=\mu_{s}/(\mu_{s}+\mu_{p})italic_β = italic_μ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / ( italic_μ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT + italic_μ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ) are set to 500500500500 and 0.90.90.90.9, respectively. We start each simulation from a fully developed field of the closest case at a higher Reynolds number, and advance it in time according to the methods detailed in the Supplemental Material 222See Supplemental Material at [URL will be inserted by publisher] for technical details on the numerical methods. There, we also provide further insight on the flow statistics, elucidate the calibration of the modal decomposition algorithm and investigate the dependence of the results from the polymer concentration. Relevant references are contained therein Dubief et al. (2023); Abdelgawad et al. (2023); Kim and Moin (1985); Fattal and Kupferman (2005); De Vita et al. (2018); Sugiyama et al. (2011); Beneitez et al. (2024); De Angelis et al. (2002); Lellep et al. (2024); Pope (2000); Serafini et al. (2022); Vega and Le Clainche (2020); Schmid (2010); Dubief et al. (2022); Couchman et al. (2024); Buza et al. (2022b); Beneitez et al. (2023).. Our parameters are chosen consistently with previous stability analysis Buza et al. (2022a) and yield values of the elasticity number El=De/Re𝐸𝑙𝐷𝑒𝑅𝑒El=De/Reitalic_E italic_l = italic_D italic_e / italic_R italic_e in {0.03,0.1,1,10,100}0.030.1110100\{0.03,0.1,1,10,100\}{ 0.03 , 0.1 , 1 , 10 , 100 }, denoting how the investigated range equally encompasses cases dominated by inertia (El1much-less-than𝐸𝑙1El\ll 1italic_E italic_l ≪ 1) and elasticity (El1much-greater-than𝐸𝑙1El\gg 1italic_E italic_l ≫ 1), as well as the intermediate range. The results presented in the following have been replicated also with the Oldroyd-B liquid, thus proving their robustness against the polymer model adopted; furthermore, we have tested the effect of different viscosity ratios β𝛽\betaitalic_β, as detailed in the Supplemental Material (Note, 2).

Refer to caption
Figure 2: Flow statistics. In (𝐚)𝐚(\mathbf{a})( bold_a ) we report the viscous (dashed), turbulent (dashed-dotted) and polymer (dotted) shears, summing up to the total shear stress τ𝜏\tauitalic_τ (solid line). At Re1500𝑅𝑒1500Re\leq 1500italic_R italic_e ≤ 1500 viscosity and elasticity dominate throughout the whole half-channel height hhitalic_h. All the contributions integrated along the wall-normal direction, in (𝐛𝐛\mathbf{b}bold_b), are thus dominated by the viscous and polymer shears, but for the case at Re=2800𝑅𝑒2800Re=2800italic_R italic_e = 2800, dominated by the turbulent shear. Consistently, the polymer stretching is enhanced, as shown in (𝐜)𝐜(\mathbf{c})( bold_c ) by the PDFs of the trace of 𝐂𝐂\mathbf{C}bold_C at the channel centerline: their barycenters are far from tr(𝐂)=3𝑡𝑟𝐂3tr(\mathbf{C})=3italic_t italic_r ( bold_C ) = 3. (𝐝)𝐝(\mathbf{d})( bold_d ) shows the J-PDFs of the normalized streamwise and wall-normal velocity fluctuations at y=h/2𝑦2y=h/2italic_y = italic_h / 2, at {0.1,0.3,0.5,0.7}0.10.30.50.7\{0.1,0.3,0.5,0.7\}{ 0.1 , 0.3 , 0.5 , 0.7 } of their maximum. The typical turbulent shape is recovered at Re=2800𝑅𝑒2800Re=2800italic_R italic_e = 2800, while in all the other cases the fluctuations decorrelate. The TKE spectra (𝐞)𝐞(\mathbf{e})( bold_e ) show a good collapse for all the low-Reynolds number cases. We show additional data with Oldroyd-B at Re=5𝑅𝑒5Re=5italic_R italic_e = 5 (black circles) and from Lellep et al. (2024) at Re=0.01𝑅𝑒0.01Re=0.01italic_R italic_e = 0.01 and De=80𝐷𝑒80De=80italic_D italic_e = 80/150150150150 as dashed/dashed-dotted lines, respectively. The 44-4- 4 slope contrasts sharply with the typical 5/353-5/3- 5 / 3 decay, recovered in TDR, pointing towards a unified view of EIT and ET.

At all the Reynolds numbers but the highest one, qualitatively similar coherent fluctuations of the velocity are superimposed to a mean flow which only slightly departs from the laminar one (more details on this in the Supplemental Material Note, 2). As visible in FIG. 1, large structures span the center of the domain, while smaller ones characterize the region near the wall. All of them exhibit a quasi-periodic spatial arrangement superimposed to random perturbations. A more quantitative characterization of the flow fields is attained looking at the balance of the total shear stress, τ𝜏\tauitalic_τ, along the half-channel height, hhitalic_h. The conservation of linear momentum requires the wall-normal derivative of τ𝜏\tauitalic_τ to balance the driving pressure gradient, dyτ=dxPsubscript𝑑𝑦𝜏subscript𝑑𝑥𝑃d_{y}\tau=d_{x}Pitalic_d start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT italic_τ = italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_P, imposing a linear trend of τ𝜏\tauitalic_τ with y𝑦yitalic_y. Three different contributions sum together in the total shear stress: a polymeric one due to the shearing component of the viscoelastic stresses 𝐓𝐓\mathbf{T}bold_T (Txydelimited-⟨⟩subscript𝑇𝑥𝑦\langle T_{xy}\rangle⟨ italic_T start_POSTSUBSCRIPT italic_x italic_y end_POSTSUBSCRIPT ⟩, where angle brackets denote averaging in time and along the homogeneous directions), a purely viscous one coming from the mean velocity profile udelimited-⟨⟩𝑢\langle u\rangle⟨ italic_u ⟩ (ρνyu𝜌𝜈subscript𝑦delimited-⟨⟩𝑢\rho\nu\partial_{y}\langle u\rangleitalic_ρ italic_ν ∂ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ⟨ italic_u ⟩), and a turbulent one caused by the correlated fluctuations of the streamwise and wall-normal velocities (ρuv𝜌delimited-⟨⟩superscript𝑢superscript𝑣-\rho\langle u^{\prime}v^{\prime}\rangle- italic_ρ ⟨ italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_v start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩, where the primes indicate fluctuations). These contributions are reported in the plots of FIG. 2a𝑎aitalic_a.

As expected for the TDR regime at Re=2800𝑅𝑒2800Re=2800italic_R italic_e = 2800, the viscoelastic contribution remains small almost everywhere and viscosity is mostly relevant close to the wall, while the turbulent contribution dominates the center of the channel. For Re1500𝑅𝑒1500Re\leq 1500italic_R italic_e ≤ 1500, instead, the shear stress balance is completely different, as the polymer and viscous shears become relevant throughout the channel and the turbulent one is depleted. To better weight the relative importance of the different contributions of the shear stress balance, we integrate the momentum balance along the wall-normal direction ((Txy+ρνyuρuv)𝑑y=h2dxPdelimited-⟨⟩subscript𝑇𝑥𝑦𝜌𝜈subscript𝑦delimited-⟨⟩𝑢𝜌delimited-⟨⟩superscript𝑢superscript𝑣differential-d𝑦superscript2subscript𝑑𝑥𝑃\int{(\langle T_{xy}\rangle+\rho\nu\partial_{y}\langle u\rangle-\rho\langle u^% {\prime}v^{\prime}\rangle)dy}=h^{2}\enspace d_{x}P∫ ( ⟨ italic_T start_POSTSUBSCRIPT italic_x italic_y end_POSTSUBSCRIPT ⟩ + italic_ρ italic_ν ∂ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ⟨ italic_u ⟩ - italic_ρ ⟨ italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_v start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ ) italic_d italic_y = italic_h start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_P), as represented graphically in FIG. 2b𝑏bitalic_b. The turbulent shear is the most relevant component at Re=2800𝑅𝑒2800Re=2800italic_R italic_e = 2800, while it is practically zero in all the other cases; those are dominated by the viscous shear, along with a non-negligible polymeric contribution (individually shown in the Supplemental Material Note, 2). Significantly, the relative amplitude of the different contributions remains unchanged for all the Reynolds numbers (ET) but the highest one (TDR).

The previous results warrant two important clarifications. First, the relative small contribution of the polymers does not mean that the polymers are not stretched. Indeed, the polymer molecules are significantly stretched regardless of the case considered, as shown by the probability density functions (PDF) of the trace of 𝐂𝐂\mathbf{C}bold_C reported in FIG. 2c𝑐citalic_c. Also here, a clear separation can be appreciated between the case at Re=2800𝑅𝑒2800Re=2800italic_R italic_e = 2800 and all the others. Second, the vanishing magnitude of the turbulent shear when Re1500𝑅𝑒1500Re\leq 1500italic_R italic_e ≤ 1500 does not mean that the turbulent fluctuations disappear, but rather suggests that the streamwise and wall-normal velocity fluctuations decorrelate. We therefore compute the joint probability density function (J-PDF) of the streamwise and wall-normal velocity fluctuations at a wall distance of h/22h/2italic_h / 2, reported in FIG. 2d𝑑ditalic_d, where velocities are made dimensionless with the mean turbulent kinetic energy in the channel, K𝐾Kitalic_K, for a better comparison of the different Reynolds numbers. In the high Reynolds case the isolines exhibit an oval shape with the major axis inclined downward, denoting a negative correlation among the two velocity components, and ejections (u<0superscript𝑢0u^{\prime}<0italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT < 0, v>0superscript𝑣0v^{\prime}>0italic_v start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT > 0) dominate over sweeps (u>0superscript𝑢0u^{\prime}>0italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT > 0, v<0superscript𝑣0v^{\prime}<0italic_v start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT < 0), analogously to the Newtonian case Tardu (2014). At this wall distance, the J-PDF peaks in the fourth quadrant. This picture completely changes when decreasing the Reynolds number, as the streamwise velocity decorrelates from the wall-normal one and its fluctuations heavily dominate in intensity; the J-PDF consequently appears flattened along the x-axis, explaining the apparent disappearance of the turbulent shear stress from FIG. 2.

The features of the low Reynolds flow fields described up to this point would appear compatible with an incomplete relaminarisation of the flow; nevertheless, its turbulent nature is confirmed by the full multi-scale Fourier’s spectra E𝐸Eitalic_E of the turbulent kinetic energy (TKE) at the centerline of the channel. Those are characterised by a power-law decay of E(kx)kx4similar-to𝐸subscript𝑘𝑥superscriptsubscript𝑘𝑥4E(k_{x})\sim k_{x}^{-4}italic_E ( italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ) ∼ italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT over roughly a decade of streamwise wave numbers kxsubscript𝑘𝑥k_{x}italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT (FIG. 2e𝑒eitalic_e), in good agreement with recent observations in a similar setup at lower Reynolds number Lellep et al. (2024) and in homogeneous isotropic elastic turbulence Singh et al. (2024). Contrarily, in the TDR regime at Re=2800𝑅𝑒2800Re=2800italic_R italic_e = 2800, our data exhibit a tendency towards the conventional E(kx)kx5/3similar-to𝐸subscript𝑘𝑥superscriptsubscript𝑘𝑥53E(k_{x})\sim k_{x}^{-5/3}italic_E ( italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ) ∼ italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 5 / 3 end_POSTSUPERSCRIPT trend expected for Newtonian turbulence at larger values of the Reynolds number. Once more, no significant difference is observed among all the low Reynolds number cases; the robustness of this statement to significant variations of the polymer concentration is detailed in the Supplemental Material Note (2). Our results therefore point towards a unified view of EIT and ET in three-dimensional fully developed planar channel flows.

Refer to caption
Figure 3: Modal decomposition, showing the temporal frequencies (𝐚)𝐚(\mathbf{a})( bold_a ), ω𝜔\omegaitalic_ω, of the two most dynamically relevant modes (b,c)𝑏𝑐(b,c)( italic_b , italic_c ) for five different values of the Reynolds number. The error bars represent the uncertainty of the decomposition algorithm calibration process Note (2). The colored panels 𝐛,𝐜𝐛𝐜\mathbf{b,c}bold_b , bold_c report the typical shapes of the streamwise, u𝑢uitalic_u, and spanwise, w𝑤witalic_w, velocity components of the modes (sampled from the case at Re=50𝑅𝑒50Re=50italic_R italic_e = 50), where positive isosurfaces are colored in purple and negative isosurfaces in orange. The identification of two dominant modes common to all the low Reynolds number cases (consistent with previous stability analysis results) corroborates the unified view of EIT and ET.

To further confirm this scenario, we apply a purely data-driven algorithm to identify and extract the coherent structures in the flow at different Reynolds numbers. In particular, we adopt the higher order dynamic mode decomposition Vega and Le Clainche (2020), a dimensionality reduction algorithm which provides a finite set of modes, each oscillating in time with a specific frequency ω𝜔\omegaitalic_ω. Past stability analysis of viscoelastic two-dimensional channel flows suggests that travelling waves originated at a relatively high-Re𝑅𝑒Reitalic_R italic_e bifurcation Garg et al. (2018) persist in EIT Page et al. (2020); Khalid et al. (2021b); Dubief et al. (2022); Buza et al. (2022a) and ET Buza et al. (2022b); Morozov (2022). These structures are mainly located away from the walls, in the central region of the channel, and are sometimes termed “arrowheads” Page et al. (2020). In ET, these coexist with a recently discovered wall-localised linear instability caused by the presence of small but non-zero diffusivity of the polymer stresses Beneitez et al. (2023); Couchman et al. (2024). The diffusive instability perturbs the flow close to the walls, in the region where the travelling waves convert most of the kinetic energy into elastic energy Beneitez et al. (2024). Our analysis suggests that all the turbulent flows at Re1500𝑅𝑒1500Re\leq 1500italic_R italic_e ≤ 1500 can be described by the superposition of two main sets of modes for all the Reynolds numbers but the highest one, once more corroborating the unified view of EIT and ET suggested by the analysis above. The frequencies of these modes are shown in FIG. 3a𝑎aitalic_a as a function of the Reynolds number, and their typical shapes are reported in FIG. 3b𝑏bitalic_b and c𝑐citalic_c. We observe that the fast modes are mainly located in the center of the channel, while the slow modes are localised close to the walls; their temporal frequencies do not undergo a significant variation with the Reynolds number. The shapes and locations of the modes are reminiscent of the results from simulations Lellep et al. (2024) and stability analysis Buza et al. (2022b); Beneitez et al. (2023).

In this work we have simulated the turbulent flow of a viscoelastic fluid in a planar channel setup across a wide range of Reynolds numbers from TDR to ET, reporting instabilities and flow structures consistent with the available literature Page et al. (2020); Beneitez et al. (2023); Dubief et al. (2023). Our results identify two distinct turbulent regimes. First, one at large Reynolds number dominated by the turbulent shear stresses with the classical energy spectra decay predicted by Kolmogorov’s theory Kolmogorov (1941) Ekx5/3similar-to𝐸superscriptsubscript𝑘𝑥53E\sim k_{x}^{-5/3}italic_E ∼ italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 5 / 3 end_POSTSUPERSCRIPT, which corresponds to the extensively investigated TDR regime Lumley (1969). Next, one at lower Reynolds number with highly elongated polymer chains producing strong polymer stresses with the energy spectra showing a fast fall-off Ekx4similar-to𝐸superscriptsubscript𝑘𝑥4E\sim k_{x}^{-4}italic_E ∼ italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT, consistently with recent results in ET even at Reynolds numbers lower than those here Singh et al. (2024); Lellep et al. (2024). We do not find any other fully-developed state in between these two, and conclude that the peculiar chaotic state attained in viscoelastic planar channel flows remains unchanged by a variation in the Reynolds number once inertial effects are sub-dominant. Thus, within this set-up, fully-developed EIT Dubief et al. (2023) and ET Steinberg (2021) share a similar structure.

Acknowledgements.
The research was supported by the Okinawa Institute of Science and Technology Graduate University (OIST) with subsidy funding from the Cabinet Office, Government of Japan. Part of this work was done during the 2023 Madrid Turbulence Workshop, organised by Prof. J. Jiménez and made possible by ERC, Caust grant ERC-AdG-101018287. M.E.R. acknowledges funding from the Japan Society for the Promotion of Science (JSPS), grant 24K17210. All authors acknowledge the computational resources provided by the Scientific Computing section of the Research Support Division at OIST and by RIKEN, under the HPCI System Research Project grants hp220099 and hp210269. M.E.R. and G.F.R. acknowledge the insightful discussion with Prof. R. R. Kerswell and Dr. M. Beneitez at the University of Cambridge.

References