Hostname: page-component-cb9f654ff-nr592 Total loading time: 0 Render date: 2025-08-23T05:43:17.330Z Has data issue: false hasContentIssue false

Jamming of elastoviscoplastic fluids in elastic turbulence

Published online by Cambridge University Press:  13 August 2025

Christopher Soriano From*
Affiliation:
Department of Chemical Engineering, University of Manchester, Manchester M13 9PL, UK
Vedad Dzanic*
Affiliation:
School of Mechanical, Medical and Process Engineering, Faculty of Engineering, Queensland University of Technology, Brisbane, QLD 4001, Australia
Vahid Niasar
Affiliation:
Department of Chemical Engineering, University of Manchester, Manchester M13 9PL, UK
Emilie Sauret
Affiliation:
School of Mechanical, Medical and Process Engineering, Faculty of Engineering, Queensland University of Technology, Brisbane, QLD 4001, Australia
*
Corresponding authors: Christopher Soriano From, christopher.from@manchester.ac.uk; Vedad Dzanic, v2.dzanic@qut.edu.au
Corresponding authors: Christopher Soriano From, christopher.from@manchester.ac.uk; Vedad Dzanic, v2.dzanic@qut.edu.au

Abstract

Elastoviscoplastic (EVP) fluid flows are driven by a non-trivial interplay between the elastic, viscous and plastic properties, which under certain conditions can transition the otherwise laminar flow into complex flow instabilities with rich space–time-dependent dynamics. We discover that under elastic turbulence regimes, EVP fluids undergo dynamic jamming triggered by localised polymer stress deformations that facilitate the formation of solid regions trapped in local low-stress energy wells. The solid volume fraction $\phi$, below the jamming transition $\phi\lt\phi_J$, scales with $\sqrt {\textit{Bi}}$, where $\textit{Bi}$ is the Bingham number characterising the ratio of yield to viscous stresses, in direct agreement with theoretical approximations based on the laminar solution. The onset of this new dynamic jamming transition $\phi \geqslant \phi _J$ is marked by a clear deviation from the scaling $\phi \sim \sqrt {\textit{Bi}}$, scaling as $\phi \sim \exp {\textit{Bi}}$. We show that this instability-induced jamming transition – analogous to that in dense suspensions – leads to slow, minimally diffusive and rigid-like flows with finite deformability, highlighting a novel phase change in elastic turbulence regimes of complex fluids.

Information

Type
JFM Rapids
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2025. Published by Cambridge University Press

1. Introduction

Complex non-Newtonian fluids are well-known to be subject to unpredictable flow instabilities (Steinberg Reference Steinberg2021; Datta et al. Reference Datta2022), a problem whose motivations originated from industry over 60 years ago due to production quality issues in processing operations involving polymeric fluids (Petrie & Denn Reference Petrie and Denn1976), which has continued to intrigue the scientific and industrial communities since (Dubief, Terrapon & Hof Reference Dubief, Terrapon and Hof2023). In particular, elastoviscoplastic fluids (EVP) characterised by their ability to exhibit elastic, viscous and plastic behaviours depending on applied stress are ubiquitous in various processes in nature (such as lava and landslide flow (Jerolmack & Daniels Reference Jerolmack and Daniels2019; Abdelgawad, Cannon & Rosti Reference Abdelgawad, Cannon and Rosti2023) and certain biological substances (Bertsch et al. Reference Bertsch, Diba, Mooney and Leeuwenburgh2022)) and industry, including materials such as pastes and gels (Balmforth, Frigaard & Ovarlez Reference Balmforth, Frigaard and Ovarlez2014; Nicolas et al. Reference Nicolas, Ferrero, Martens and Barrat2018). The EVP materials, a class of non-Newtonian ‘yield stress’ fluids, behave as elastic solids with finite deformations below their yield criteria and above which flow like complex viscoelastic fluids (Bonn et al. Reference Bonn, Denn, Berthier, Divoux and Manneville2017; Nicolas et al. Reference Nicolas, Ferrero, Martens and Barrat2018). Understanding the dynamics of EVP fluid flows remains an essential area of research that bridges fundamental science and practical applications. A fascinating and challenging aspect of EVP fluid flows is the occurrence of instabilities driven by a delicate balance between this solid-like and liquid-like behaviour that is highly localised and time-dependent (Nicolas et al. Reference Nicolas, Ferrero, Martens and Barrat2018; Varchanis et al. Reference Varchanis, Haward, Hopkins, Syrakos, Shen, Dimakopoulos and Tsamopoulos2020). Flow instabilities in EVP fluids are prominent when subjected to deformation rates near their material yield stress, making their applications unpredictable and difficult to control (Balmforth et al. Reference Balmforth, Frigaard and Ovarlez2014; Nicolas et al. Reference Nicolas, Ferrero, Martens and Barrat2018). Insights into these pave the way for innovations in material design and process optimisation, addressing challenges across a wide range of industries and disciplines, from the production of fast-moving consumer goods (Balmforth et al. Reference Balmforth, Frigaard and Ovarlez2014; Bonn et al. Reference Bonn, Denn, Berthier, Divoux and Manneville2017) to emerging technologies, such as three-dimensional (3-D) printing soft biomaterials (Bertsch et al. Reference Bertsch, Diba, Mooney and Leeuwenburgh2022; Smith & Hashmi Reference Smith and Hashmi2024; Zhang, Dolatshahi-Pirouz & Orive Reference Zhang, Dolatshahi-Pirouz and Orive2024).

Unlike inertial instabilities that arise in Newtonian fluid flows, such as turbulent flows, viscoelastic fluid flow instabilities arise even without inertial effects in the low-Reynolds-number regime $Re=\rho V \ell / \mu \leqslant 1$ (Groisman & Steinberg Reference Groisman and Steinberg2000). Inertialess instabilities can manifest in various forms (Datta et al. Reference Datta2022), including shear banding (Cochran et al. Reference Cochran, Callaghan, Caven and Fielding2024), where regions of differing shear rates develop, and elastic instabilities, where the fluid’s elastic nature leads to flow irregularities. The viscous to elastic effects are measured by the Weissenberg number $Wi=\lambda V/\ell \gg 1$ , and the viscoelastic to inertial effects are characterised by the elasticity number $El\equiv \textit{Wi}/Re = \lambda \mu /(\rho \ell ^2)\gg 1$ . Here, $\lambda$ is the longest polymer relaxation time, the characteristic length scale is $\ell$ , $\rho$ is density, $\mu$ is total viscosity, and the characteristic velocity is $V$ . The additional property, a yield stress criteria $\sigma _y$ , traditionally referred to as a jamming transition in EVP material (i.e. the transition from a jammed solid state to a fluid-like state) (Bonn et al. Reference Bonn, Denn, Berthier, Divoux and Manneville2017), and its ratio to the viscous stress is characterised by the Bingham number $\textit{Bi}=\sigma _y\ell /(\mu V)$ . Time-dependent EVP flows often lead to non-homogeneity with complex flow patterns and transitions (Abdelgawad et al. Reference Abdelgawad, Cannon and Rosti2023; Dzanic, From & Sauret Reference Dzanic, From and Sauret2024). More specifically, the elastic property generates an anisotropic stress contribution, which, in extreme cases (i.e. $Re \leqslant 1$ ), can transition the flow to a chaotic self-sustaining state, known as elastic turbulence (ET), initially discovered to be triggered by linear hoop stress instabilities in curvilinear geometries (Groisman & Steinberg Reference Groisman and Steinberg2000; Steinberg Reference Steinberg2021). Later studies identified ET in rectilinear geometries – such as, notably, in viscoelastic Kolmogorov flow by Boffetta et al. (Reference Boffetta, Celani, Mazzino, Puliafito and Vergassola2005b ) – have recently been shown to be triggered by the mechanism known as a centre-mode instability (Kerswell & Page Reference Kerswell and Page2024; Lewy & Kerswell Reference Lewy and Kerswell2025), giving rise to ‘arrowhead’ structures (Page, Dubief & Kerswell Reference Page, Dubief and Kerswell2020) that interact chaotically, transitioning the flow to ET through a bursting scenario. Plasticity on its own (i.e. viscoplasticity where there is negligible elasticity) can dramatically impact fluid flows. Notably, recent work (Abdelgawad et al. Reference Abdelgawad, Cannon and Rosti2023) demonstrated that plasticity $\textit{Bi}\gt 1$ significantly alters both the energy distribution and intermittency of inertial turbulence $Re\gg 10^3$ , altering Kolmogorov’s well-known inertial and dissipative 5/3 scaling exponent to a new scaling exponent of 2.3. On the other hand, the study of instabilities in inertialess EVP fluid flows, combining both elastic and plastic behaviours, is limited. Our recent work on EVP extensional flows in the ET regime (Dzanic et al. Reference Dzanic, From and Sauret2024) illustrated that the regions of unyielded material increase – indicated by a higher solid volume fraction ( $\phi$ ) – as $\textit{Bi}$ increases. We found that the impact of plasticity on the dynamic behaviour, including transitions to periodic, aperiodic and chaotic regimes, is highly dependent on the flow geometry.

In this work, we discover that EVP fluid flows transition to a jammed state with $\textit{Bi}$ . The connection to jamming has not previously been made for EVP fluid flows. This elastic-plastic-induced jamming phenomenon is notably distinct from the traditional intuition of jamming transitions in EVP fluids defined as the material’s yield stress criterion $\sigma _y$ (Nicolas et al. Reference Nicolas, Ferrero, Martens and Barrat2018; N’Gouamba et al. Reference N’Gouamba, Goyon and Coussot2019). To gain a better understanding of the nature and impact of this phase transition, we demonstrate the direct analogue features of jamming in inertialess EVP flows in the ET regime. We show that intermediate $\phi \hbox{-}$ regimes approaching jamming from below $\phi \rightarrow \phi _J$ follow a square-root scaling behaviour $\phi \sim \sqrt {\textit{Bi}}$ , reminiscent of scaling behaviour in dense suspensions. Beyond the jamming transition $\phi \gt \phi _J$ , the scaling behaviour transitions dramatically – a key signature of a phase transition – growing exponentially as $\phi \sim \exp {\textit{Bi}}$ .

2. Methods

We numerically study inertialess EVP instabilities with the well-known Kolmogorov flow problem in a two-dimensional (2-D) domain $\boldsymbol{x}$ with double periodic boundary conditions. The dimensionless governing equations of the EVP fluid are given by the incompressible Navier–Stokes equation,

(2.1) \begin{equation} \boldsymbol{\nabla } \boldsymbol{\cdot } \boldsymbol{u} = 0, \quad Re\frac {{\textrm D}\boldsymbol{u}}{{\textrm D}t} = -\boldsymbol{\nabla } P + {\beta }\boldsymbol{\Delta }\boldsymbol{u}+ \boldsymbol{\nabla } \boldsymbol{\cdot } \boldsymbol{\sigma } + \boldsymbol{F}_0, \end{equation}

coupled with the polymer stress tensor, $\boldsymbol{\sigma }=Wi^{-1} (1-\beta )(f {\unicode{x1D63E}}-{\unicode{x1D644}}\,)$ , described by a space–time dependent conformation tensor ( $\unicode{x1D63E}\,$ ) constitutive equation,

(2.2) \begin{equation} \frac {{\textrm D}{\unicode{x1D63E}}}{{\textrm D}t} = {\unicode{x1D63E}}\boldsymbol{\cdot }\left (\boldsymbol{\nabla }\boldsymbol{u}\right ) +\left (\boldsymbol{\nabla }\boldsymbol{u}\right )^{\mathsf{T}} \boldsymbol{\cdot } {\unicode{x1D63E}} - \frac {\mathcal{F}}{Wi} (f {\unicode{x1D63E}} - {\unicode{x1D644}}\,)+\kappa \boldsymbol{\Delta }{\unicode{x1D63E}}. \end{equation}

The functions $f = (L^2 - \operatorname {tr}{\unicode{x1D644}}\,)/(L^2 - \textrm{tr}{\unicode{x1D63E}}\,)$ and $\mathcal{F}(\sigma _v,{\textit{Bi}})=\max ( 0,(\sigma _v-{\textit{Bi}})/\sigma _v )$ are constitutive polymer models, namely the finite extensible nonlinear elastic Peterlin (FENE-P) model (Peterlin Reference Peterlin1961) and the Saramito yield stress model (Saramito Reference Saramito2007), for the elastic and plastic non-Newtonian behaviour, respectively. The yield stress is defined by $\sigma _v = \sqrt {\sigma _{J2}}$ (Saramito Reference Saramito2007), where $\sigma _{J2} = ({1}/{2})(\boldsymbol{\sigma }_{d}:\boldsymbol{\sigma }_{d})$ is the second invariant of the deviatoric part of the stress tensor $\boldsymbol{\sigma }_d = \boldsymbol{\sigma } -{\unicode{x1D644}}(\operatorname {tr}\boldsymbol{\sigma } /\operatorname {tr}{\unicode{x1D644}}\,)$ . The model predicts recoverable Kelvin–Voigt viscoelastic deformation in the unyielded state ( $\mathcal{F}=0$ for $\sigma _v\leqslant \textit{Bi}$ ), whereas the FENE-P viscoelastic model is retained beyond yielding ( $0\lneq \mathcal{F} \leqslant 1$ for $\sigma _v\gt \textit{Bi}$ ), and in the purely viscoelastic limit $\textit{Bi}\,{=}\,0\Rightarrow \mathcal{F}\,{=}\,1$ . Here, $L$ is the maximum polymer extensibility ( $L^2 \gt \textrm{tr}{\unicode{x1D63E}}\,$ ), which we set to $L=50$ , characteristic polymer concentration $\beta =\nu _s/(\nu _s+\nu _p)=0.9$ as in Rosti et al. (Reference Rosti, Izbassarov, Tammisola, Hormozi and Brandt2018); Abdelgawad et al. (Reference Abdelgawad, Cannon and Rosti2023), $\nu _{s}$ and $\nu _{p}$ are the solvent and polymer viscosity, respectively, $\unicode{x1D644}$ is the identity tensor ( $\operatorname {tr}{\unicode{x1D644}}=2$ ), $\boldsymbol{u}$ is the velocity field,  $P$ the pressure, and $\boldsymbol{F}_0$ is the external driving force (total force $\boldsymbol{F}= \boldsymbol{F}_p + \boldsymbol{F}_0$ , where $\boldsymbol{F}_p = \boldsymbol{\nabla } \boldsymbol{\cdot } \boldsymbol{\sigma }$ ). Equations (2.1) and (2.2) are solved using a symmetric positive-definite conserving numerical solver developed in-house (Dzanic et al. Reference Dzanic, From and Sauret2024, Reference Dzanic, From and Sauret2022a , Reference Dzanic, From and Sauretb , Reference Dzanic, From and Sauretc , Reference Dzanic, From and Sauretd ), comprising the lattice Boltzmann method coupled with a high-order finite-difference scheme. Let $\boldsymbol{n}$ be the level of periodicity in each direction, where setting $n_x,n_y\gt 1$ results in $n_x\times n_y$ unit cells. For all simulations, we set $n_x=6$ and $n_y=4$ , ensuring unicity is conserved, with $N^2=128^2$ grid points in each unit cell, i.e. spatial resolution of $2\pi /N$ (we present grid convergence test in the supplementary material, § S1.3, figures S1 and S2). Numerical regularity is added to (2.2) through an additional term $ \kappa \boldsymbol{\Delta }{\unicode{x1D63E}}$ with a specified artificial diffusivity $\kappa$ . While this numerical regularity strategy remains highly debated (Dzanic et al. Reference Dzanic, From and Sauret2022b ; Couchman et al. Reference Couchman, Beneitez, Page and Kerswell2024; Yerasi et al. Reference Yerasi, Picardo, Gupta and Vincenzi2024), it is essential to simulate ET flow regimes due to inherent steep polymer stress gradients (Dzanic et al. Reference Dzanic, From and Sauret2022c ; Gupta & Vincenzi Reference Gupta and Vincenzi2019). We minimise $\kappa$ and any associated artefacts by setting the admissible Schmidt number $Sc=\nu _s/\kappa =10^3$ as in Berti et al. (Reference Berti, Bistagnino, Boffetta, Celani and Musacchio2008), resulting in $\kappa = 1.6 \times 10^{-4}$ , which is smaller than that in recent work (Nichols, Guy & Thomases Reference Nichols, Guy and Thomases2025), an order of magnitude of those realistically expected from the diffusion of the polymer’s centre of mass in kinetic theory (Morozov Reference Morozov2022). Full details on the employed methodology are reported in the supplementary material, § S1.

The Kolmogorov shear flow is driven by a constant external force $\boldsymbol{F}_0 = (F_x,F_y)$ , given by $ F_x(y) = F_0 \cos (Ky), \ F_y = 0$ , with an amplitude $F_0=V\nu K^2$ , scaled with the characteristic peak laminar velocity $V$ in the absence of polymer stress diffusion ( $\kappa = 0$ ) imposing a constant pressure gradient $\partial _x P = F_0 \cos (Ky)$ , where $K=2$ is the spatial frequency $\ell = 1/K$ and the turnover time $T = \nu _s K/F_0$ . (Note, the base laminar flow, $\max _{y}(u_{x}(y))$ , is approximately $V$ ; however, the deviation is negligible with the small but finite diffusivity $\kappa =\mathcal{O}(10^{-4})$ since $\max _{y}(u_{x}(y))\rightarrow V$ in the limit $\kappa \rightarrow 0$ .) The flow is in the inertialess ET regime, and, as such, we set $Re = 1$ and $Wi=20$ , such that elastic effects are dominant over inertial effects $El=20$ . In doing so, we limit all the observed changes to the well-known viscoelastic Kolmogorov flow (Boffetta et al. Reference Boffetta, Celani, Mazzino, Puliafito and Vergassola2005b ; Berti et al. Reference Berti, Bistagnino, Boffetta, Celani and Musacchio2008; Berti & Boffetta Reference Berti and Boffetta2010; Kerswell & Page Reference Kerswell and Page2024; Lewy & Kerswell Reference Lewy and Kerswell2025) to be due to the introduction of plasticity, which we vary by $\textit{Bi}$ .

3. Results

When dealing with EVP fluids in practice, a macroscopic point of view of the yield transition is usually adopted, where it is assumed that an applied shear above the yield criteria $\sigma _y$ will sufficiently shear-thin the material, allowing it to flow (Dennin Reference Dennin2008; Bonn et al. Reference Bonn, Denn, Berthier, Divoux and Manneville2017). Here, we will model directly how much of the domain remains solid (unyielded) under the imposed flow field. The solid volume fraction is calculated by $\phi (t)=\mathcal{V}_{\mathcal{F}=0}(t)/\mathcal{V} = \lvert \{k:\mathcal{F}(x_k,t)=0\}\rvert / (n_x n_y N^2)$ where ‘:’ denotes the set formed by the unyielded regions ( $\mathcal{F}=0$ ) of $k$ , and ‘ $\lvert \boldsymbol{\cdot }\rvert$ ’ denotes cardinality (Dzanic et al. Reference Dzanic, From and Sauret2024). The polymer stretching $\operatorname {tr}{\unicode{x1D63E}}$ field at various $\textit{Bi}$ is shown in figure 1. (See supplementary movies 14 for $\textit{Bi}=0,\ 0.5,\ 2$ and $3$ , respectively.) The volume fraction $\phi$ time series and statistics are shown in figure 2(a) and (b,c), respectively. Here, $\phi$ increases with $\textit{Bi}$ non-trivially, with fluctuations of $\phi (t)$ (yielding and unyielding) in the statistically homogenous region $\bar {t}$ varying non-monotonically with $\textit{Bi}$ (figure 2 b,c). At $\textit{Bi}=3$ , the flow slowly transitions to a nearly completely jammed state, arresting the flow (as shown figure 2 a and supplementary movie 4).

Figure 1. The EVP Kolmogorov flow in the ET regime. Representative snapshots of the polymer stretching $\operatorname {tr}{\unicode{x1D63E}}$ field $x=[0,6\times 2\pi )$ and $y=[0,4\times 2\pi )$ at various Bingham numbers, from $\textit{Bi} = 0$ to $ 3$ (columns), at different instances in time (rows): from the initial steady state $\sim\!100T$ (top) to the transition $\sim\!300T$ (middle), and the statistically homogenous regime $\overline {t}\,{\gtrsim}\, 600T$ (bottom). The grey areas represent the instantaneous unyielded $\mathcal{F}=0$ regions. The first column $\textit{Bi} = 0$ corresponds to a purely viscoelastic ET. Note, the colour bar is truncated below the true maximum at $\textrm{tr}(\boldsymbol{C}) = 1500$ for visual clarity.

Figure 2. The solid (unyielded) volume fraction ( ${\textit{a}}$ ) time series $\phi (t)$ and its fluctuation statistics in the statistically homogeneous regime $\overline {t}\,{\gtrsim}\, 600T$ , including ( ${\textit{b}}$ ) violin distribution density superimposed with box-whisker statistics relative to the median $\widetilde {\phi }$ (white line) and ( ${\textit{c}}$ ) the root-mean-squared fluctuations, where $\overline {\phi }$ is the temporal mean. The colour scheme in ( ${\textit{a}}$ ) refers to $\textit{Bi}\gt 0$ as in ( ${\textit{b}}$ ) and ( ${\textit{c}}$ ). In ( ${\textit{b}}$ ), the box plots summarise the lower and upper quartile range of $\phi (t\geqslant \overline {t})$ with violins visualising the density and shape of the distribution. Notably, the strongest and broadest distribution of fluctuations in $\phi (t\geqslant \overline {t})$ is at $\textit{Bi}=2$ . The system energy balance of the spatiotemporal mean for ( ${\textit{d}}$ ) instantaneous kinetic energy (3.1), ( ${\textit{e}}$ ) the viscous dissipation $\epsilon _\nu$ and elastic dissipation $\epsilon _p$ , and ( ${\textit{f}}$ ) their corresponding fluctuations (3.2). The directory including the data and the notebook that generated this figure can be accessed at https://www.cambridge.org/S0022112025104588/JFM-Notebooks/files/Figure_2/Fig2.ipynb.

A notable feature of viscoelastic Kolmogorov flow is the formation of coherent structures (CS) of the stress field, the aforementioned arrowhead structures (Page et al. Reference Page, Dubief and Kerswell2020), which are well-known for the purely viscoelastic case ( $\textit{Bi}=0$ ) at $Wi\gt 10$ (Boffetta et al. Reference Boffetta, Celani, Mazzino, Puliafito and Vergassola2005b ; Berti et al. Reference Berti, Bistagnino, Boffetta, Celani and Musacchio2008; Berti & Boffetta Reference Berti and Boffetta2010; Lewy & Kerswell Reference Lewy and Kerswell2025) and manifest as travelling elastic waves in the streamwise direction $t\,{\gtrsim}\,\, 300T$ (see $\textit{Bi}=0$ case in figure 1 and supplementary movie 1). For EVP cases, similar CS appear due to viscoelastic instabilities of high elasticity, where the addition of spatiotemporal interplay between its viscoelastic-solid and-fluid behaviour leads to further dynamic structural stress deformations of the CS arrowheads (see supplementary movies). These modifications to the CS, which manifest even through the introduction of minimal plasticity at $\textit{Bi}=0.25$ (figure 1), immediately imply modifications to the modes of elastic waves, altering the transition route to the chaotic self-sustaining ET state (Kerswell & Page Reference Kerswell and Page2024; Lewy & Kerswell Reference Lewy and Kerswell2025). The spectral scaling exponent $E\propto k^{-\alpha }$ at $\textit{Bi}=0$ follows the exponent $\alpha =4$ in agreement with Lellep, Linkmann & Morozov (Reference Lellep, Linkmann and Morozov2024), Lewy & Kerswell (Reference Lewy and Kerswell2025) and all cases $\textit{Bi}\leqslant 2.5$ are within the ET regime, as shown in supplemetary material, § S2, figure S3. Notably, increasing $\textit{Bi}$ progressively flattens the scaling exponent $4\gt \alpha \gt 3$ and reduces the length of the inertial subrange. The unyielded (viscoelastic-solid) regions initially ( $t\sim 100T$ , top row in figure 1) form between shear layers (i.e. low-shear-rate regions) before manifesting behind the CS front $t\,{\gtrsim}\, 300T$ . These low-shear regions effectively act as low-energy wells (Donley et al. Reference Donley, Narayanan, Wade, Park, Leheny, Harden and Rogers2023) that facilitate the formation of unyielded regions where, due to mass conservation, the stress is redistributed between shear layers (see the middle row $\sim\!300T$ in figure 1). During the initial transition $\sim\!300T$ , the length of the unyielded regions in the CS increases with $\textit{Bi}$ due to a combination of greater $\phi$ (a consequence of increased yield criteria) and increased elastic effects with increasing $\textit{Bi}$ (at a given constant $\textit{Wi}$ ), causing longer streaks. This behaviour is also apparent in the shape of unyielded regions as they become increasingly deformed and elongated. In the statistically homogeneous regime $\overline {t}\,{\gtrsim}\, 600T$ for $\textit{Bi}=0.25$ to $2.5$ , the unyielded regions manifest as local rearrangements with a broad distribution of sizes and shapes leading to further deformations of the CS (see figure 1 and figure 2 b–c) – concomitant with the view that plastic events lead to a redistribution of elastic stresses in the system (Nicolas et al. Reference Nicolas, Ferrero, Martens and Barrat2018; N’Gouamba et al. Reference N’Gouamba, Goyon and Coussot2019). For $\textit{Bi}\geqslant 1$ , unyielded regions interact across shear layers, merging or splitting each other, where increasing $\textit{Bi}$ increasingly disorganises the base flow until at the extreme $\textit{Bi}=3$ where the polymers are stretched in thin and highly localised regions. Notably, $\textit{Bi}=3$ is the only case with two distinct transitions at $\sim\!300T$ and $\sim\!800T$ , reaching a statistically homogeneous regime $\overline {t}\,{\gtrsim}\, 1000T$ (figure 2 a).

Figure 3. Features of jamming. Spatiotemporal mean of the velocity components: (a) the mean streamise flow profile $\langle \overline {U_x} (y^*)\rangle = (K n_y)^{-1}\sum _{k=0}^{K n_y-1}\lvert \langle \overline {u_x}(y^* + k\pi )\rangle _{{x}} \rvert$ along $y^* \in [0,\pi )$ , (b) $\langle \overline {u_y} (y) \rangle _{{x}}$ and (c) the spatiotemporal mean of the polymer force $\langle \boldsymbol{\overline {\boldsymbol{\nabla }\boldsymbol{\cdot }\sigma }} \rangle _{\boldsymbol{x}} = \langle \overline {\boldsymbol{F}_{p}} \rangle _{\boldsymbol{x}}$ . In (a), velocity profiles are superimposed (grey dashed lines) with base sinusoidal profile scaled by the amplitude of each $\textit{Bi}$ case, i.e. $\max _y(\langle \overline {U_x} (y)\rangle )\operatorname {cos}(Ky)$ . (d) Influence of plasticity on the energy injection rate per unit area, measured by flow resistance $\mathcal{R}$ , the ratio between the power injected in the statistically homogeneous regime to the base laminar fixed point. (e) Volume fraction $\phi (t)$ as a function of Bingham number $\textit{Bi}$ , comparing (diamonds) the temporal mean $\bar {\phi }=\langle \phi \rangle _{\overline {t}}$ numerical simulations and (circles) the theoretically approximated $\phi$ (3.3). For $\phi \lt \phi _J$ , we find that $\phi \sim \sqrt {\textit{Bi}}$ (blue line) with the linear fit $\phi =0.387 \sqrt {\textit{Bi}} -9.7\times 10^{-2}$ with a squared correlation coefficient $R^2=0.991$ . Beyond the jamming transition $\phi _J\simeq 0.54$ at $\textit{Bi}=2$ (red dash-dot line) $\phi \sim \exp {\textit{Bi}}$ (green line) with the linear fit $\phi =3.2\times 10^{-2} \exp {\textit{Bi}} - 0.24$ with ${R}^2=0.999$ . The directory including the data and the notebook that generated this figure can be accessed at https://www.cambridge.org/S0022112025104588/JFM-Notebooks/files/Figure_3/Fig3.ipynb.

Increasing $\textit{Bi}$ decreases the overall energy (figure 2 d), inferring a shift in the instantaneous kinetic energy balance $\partial E_k/\partial t\approx 0$ in the statistically homogeneous regime,

(3.1) \begin{equation} \frac {\partial E_k}{\partial t} = \langle \varepsilon _i\rangle _{\bar {t}}-\langle \varepsilon _{\nu }\rangle _{\bar {t}} - \langle \varepsilon _p\rangle _{\bar {t}} = 0, \quad t\geqslant \bar {t}, \end{equation}

where $\varepsilon _i = \langle \boldsymbol{u}\boldsymbol{\cdot }\boldsymbol{F}_0\rangle _{\boldsymbol{x}}$ is the energy input from the Kolmogorov forcing ( $\boldsymbol{F}_0 = (F_0 \cos (y/\ell _y), 0)$ ), which is dissipated due to contributions from both the viscous Newtonian component $\varepsilon _{\nu }=2\mu _s\langle {\unicode{x1D63F}}:{\unicode{x1D63F}}\rangle _{\boldsymbol{x}}$ and the non-Newtonian polymer component $\varepsilon _p=\langle \boldsymbol{u}\boldsymbol{\cdot }(\boldsymbol{\nabla }\boldsymbol{\cdot }\boldsymbol{\sigma })\rangle _{\boldsymbol{x}}$ . Here, ${\unicode{x1D63F}} = ({1}/{2})(\boldsymbol{\nabla }\boldsymbol{u} + \boldsymbol{\nabla }\boldsymbol{u}^{\mathsf{T}})$ is the rate-of-strain tensor. The contributions of $\varepsilon _p$ (elastic and plastic behaviour) increase with $\textit{Bi}$ and eventually dominate for $\textit{Bi}\geqslant 2.5$ , absorbing more energy than the dissipation of viscous kinetic energy (figure 2 e). Fluctuations of the kinetic energy,

(3.2) \begin{equation} \frac {\partial {E_k}^{\prime }}{\partial t} = \langle {\varepsilon _i}^{\prime }\rangle _{\bar {t}}-\langle {\varepsilon _{\nu }}^{\prime }\rangle _{\bar {t}} - \langle {\varepsilon _p}^{\prime }\rangle _{\bar {t}}= 0, \quad t\geqslant \bar {t}, \end{equation}

in which the production term ${\varepsilon _i}^{\prime } =\langle {u_i}^{\prime }{u_j}^{\prime }({(\partial \overline {u_i})}/{(\partial x_j)})\rangle _{\boldsymbol{x}} \approx 0$ (figure 2 f) due to negligible inertial ( $Re=1$ ) contributions to the velocity fluctuations ${\boldsymbol{u}^{\prime }}=\boldsymbol{u}-\overline {\boldsymbol{u}}$ (Gotoh & Yamada Reference Gotoh and Yamada1984; Boffetta et al. Reference Boffetta, Celani and Mazzino2005a ), where $\overline {\boldsymbol{u}}=\langle {\boldsymbol{u}}\rangle _{\bar {t}}$ . Instead, the fluctuating non-Newtonian polymer term ${\varepsilon _p}^{\prime }=2\mu _s\langle {\unicode{x1D63F}}^{\,\prime }:{\unicode{x1D63F}}^{\,\prime }\rangle _{\boldsymbol{x}}$ is the positive source term to sustain and counteract the fluctuating viscous dissipation ${\varepsilon _{\nu }}^{\prime }=\langle \boldsymbol{u}^{\prime }\boldsymbol{\cdot }(\boldsymbol{\nabla }\boldsymbol{\cdot }\boldsymbol{\sigma }^{\prime })\rangle _{\boldsymbol{x}}$ , which vary non-monotonically with $\textit{Bi}$ (figure 2 f). Interestingly, the viscous ${\varepsilon _{\nu }}^{\prime }$ and elastic dissipation ${\varepsilon _p}^{\prime }$ contributions to the energy budget clearly indicate a transition at $\textit{Bi}=2$ , which aligns directly with the strongest and broadest distribution of fluctuations in $\phi (t)$ (figure 2 b and c). Fluctuations in the dissipation then decrease for $\textit{Bi}\gt 2$ (figure 2 f) due to the polymer dissipation exceeding the viscous dissipation (figure 2 e), i.e. the system becomes unable to sustain the flow due to strong fluctuations. The non-monotonic relationship between $\textit{Bi}$ and the fluctuations $\phi (t)$ along with the impaired flow energy for $\textit{Bi}\geqslant 2$ are clear features of a typical phase transition – such a phase transition from a fluid-like to solid-like state is specifically known as jamming (Bonn et al. Reference Bonn, Denn, Berthier, Divoux and Manneville2017).

A remarkable feature of the Kolmogorov flow is that even in the chaotic ET regime, the mean velocity (figure 3 a) and conformation tensor are accurately described by sinusoidal profiles of the base flow with smaller amplitudes with respect to the laminar fixed point (Boffetta et al. Reference Boffetta, Celani and Mazzino2005a , Reference Boffetta, Celani, Mazzino, Puliafito and Vergassolab ; Berti & Boffetta Reference Berti and Boffetta2010). We observe the $\langle \overline {U_x} \rangle$ profile for all $\textit{Bi}$ retain this feature (see figure S4 in the supplementary material) with the velocity front decreasing as $\textit{Bi}$ increases (figure 3 a). Cases for $\textit{Bi}\leqslant 0.5$ show traces of CS (figure 1) resembling purely elastic flow behaviour. The decreasing front velocity amplitude $\langle \overline {U_x} \rangle$ is minor for $\textit{Bi}\leqslant 1$ but abruptly shifts for $\textit{Bi}\geqslant 2$ decreasing dramatically, approaching static rest $\langle \overline {U_x} \rangle \rightarrow 0$ at $\textit{Bi}=3$ in figure 3 a (see supplemetary movie 4). This infers a shift in the energy balance, in particular energy dissipation (figure 2 d–f), further evident from the increase in the transverse velocity component $\langle \overline {u_y} (y) \rangle _x$ and the polymer force $\boldsymbol{F}_{p} = \boldsymbol{\nabla }\boldsymbol{\cdot }\boldsymbol{\sigma }$ with increasing $\textit{Bi}$ in figures 3(b) and 3(c), respectively. The deformation and locality of plastic events redistribute stresses anisotropically (Nicolas et al. Reference Nicolas, Ferrero, Martens and Barrat2018) as observed in figures 1 and 2, where the growth of these unyielded regions is strongly correlated with secondary flow effects (figures 3 b and 3 c), leading to a departure from the Kolmogorov base sinusoidal flow profile as $\textit{Bi}$ increases (figure 3 a). Notably, the small transverse loads $F_{p,y}$ for $\textit{Bi}\gt 1$ in figure 3(c) arise due to finite deformability (figure 1), which is a well-known feature of traditional jamming in dense suspension flows (see e.g. Cates et al. Reference Cates, Wittmer, Bouchaud and Claudin1998).

We estimate the jamming transition for the present Kolmogorov flow configuration to be at the transition $\textit{Bi}=2$ , $\phi _J\simeq 0.54$ (figure 3 e), a value common in many traditional jammed systems, such as dense suspensions (Peters, Majumdar & Jaeger Reference Peters, Majumdar and Jaeger2016; Bonn et al. Reference Bonn, Denn, Berthier, Divoux and Manneville2017). Approaching the jamming transition from below $\phi \rightarrow \phi _J$ ( $\textit{Bi}\lt 2$ ), we find that $\phi \sim \sqrt {\textit{Bi}}$ in figure 3(e). Above jamming $\phi \geqslant \phi _J$ is a clear deviation from $\phi \sim \sqrt {\textit{Bi}}$ , scaling as $\phi \sim \exp {\textit{Bi}}$ ; such a dramatic change in behaviour further supports the features of phase transition observed in figure 2. To theoretically approximate $\phi$ in figure 3(e), we derive the laminar solution (see supplementary material, § S3),

(3.3) \begin{equation} {\unicode{x1D63E}}_{\textit{lam}}(y) = \begin{pmatrix} f^{-1}\left (1 + \dfrac {2 K^2Wi^2}{\mathcal{F}^2 f} \operatorname {sin}^2(Ky)\right ) & K\dfrac {Wi}{\mathcal{F} f}\operatorname {sin}(Ky) \\[12pt] K\dfrac {Wi}{\mathcal{F} f}\operatorname {sin}(Ky) & 1 \end{pmatrix}\!. \end{equation}

Figure 3(e) shows the scaling behaviour, prior to and above $\phi _J$ , is in direct agreement with $\phi$ approximated by the laminar solution (3.3). The close agreement at intermediate $\textit{Bi}\leqslant 2.5$ is surprising given the drastic flow deformations (figure 1). Interestingly, for $\textit{Bi}=3$ , while (3.3) does not predict $\langle \phi \rangle _{\bar {t}}$ in $\overline {t}\,{\gtrsim}\, 1000T$ (figure 3 e), it is consistent with $\langle \phi \rangle _{t}$ in the initial steady-state region $500T\lesssim t\lesssim 800T$ (figure 2 a). Analogue scaling behaviour $\phi \sim \sqrt {\textit{Bi}}$ for $\phi \leqslant \phi _J$ is commonly observed in traditional jamming transitions of dense suspensions (Peters et al. Reference Peters, Majumdar and Jaeger2016; Bonn et al. Reference Bonn, Denn, Berthier, Divoux and Manneville2017). Recent findings by Abdelgawad et al. (Reference Abdelgawad, Cannon and Rosti2023) showed that EVP fluids in inertial turbulence ( $Re\gg 10^3$ ), with subdominant elastic effects $Wi\lesssim 1$ , plasticity $\textit{Bi}\gt 1$ increases intermittency. We find their results for $\textit{Bi}\gt 1$ agree with our scaling $\phi \sim \sqrt {\textit{Bi}}$ approaching the jamming transition (as shown in supplementary material, figure S5). Whether an analogue exponential scaling beyond the jamming transition $\phi \gt \phi _J$ is relevant in such inertia-dominated turbulent regimes remains to be observed.

The potential implications in practical applications of this elastic-plastic-induced jamming will be inferred by measuring the flow resistance to the power injected (Berti et al. Reference Berti, Bistagnino, Boffetta, Celani and Musacchio2008). The energy injection rate per unit area (power injected), $ \mathcal{P}_{\textit{inj}} = \langle \overline {\boldsymbol{u} \boldsymbol{\cdot } \boldsymbol{F}} \rangle _{\boldsymbol{x}}$ , with $\boldsymbol{F}=\boldsymbol{F}_0+\boldsymbol{F}_{\!p}$ . In the laminar steady state, denoted by $\widehat {\boldsymbol{\cdot }}$ , $\partial _x \widehat {P}$ is constant and $\langle u_x \rangle _{\boldsymbol{x}} \approx \widehat {u}$ , where $\widehat {\mathcal{P}_{\textit{inj}}} = ({1}/{2}) V F_0$ (see supplemetary material, § S3.1). Due to the body force and the energy dissipated by the elastic-plastic-viscous effects, the flow resistance, $\mathcal{R}=\mathcal{P}_{\textit{inj}}/\widehat {\mathcal{P}_{\textit{inj}}} = \langle \overline {\boldsymbol{u} \boldsymbol{\cdot } \boldsymbol{F}} \rangle _{\boldsymbol{x}}/({1}/{2} V F_0)$ , gradually increases for $\textit{Bi}\leqslant 1$ before shifting as a consequence of the jamming phase change (figure 3 d). Moreover, our jamming transition scaling behaviour $\phi \sim \sqrt {\textit{Bi}}$ has a potentially direct connection with other known scaling, namely, the Fanning friction factor in channel flows (Rosti et al. Reference Rosti, Izbassarov, Tammisola, Hormozi and Brandt2018) and the pressure drop in porous media flows (De Vita et al. Reference De Vita, Rosti, Izbassarov, Duffo, Tammisola, Hormozi and Brandt2018).

The macroscopic point of view of jamming in EVP materials is typically considered as the yield criteria, an immediate shift to a flow state once a certain stress threshold is reached, e.g. if the flow applied induces sufficiently high shear (Dennin Reference Dennin2008; Bonn et al. Reference Bonn, Denn, Berthier, Divoux and Manneville2017). We demonstrate the statistically homogenous dynamics of EVP Kolmogorov flows in the ET regime (dominant elastic instabilities) across a range of $\textit{Bi}$ ( $\phi$ ) align with the laminar scaling prediction below and above the transition $\phi _J$ , i.e. $\phi \sim \sqrt {\textit{Bi}}$ and $\phi \sim \exp {\textit{Bi}}$ , respectively (figure 3 e). The connection to jamming in this work has profound implications on EVP fluid flows; it describes plastic events, such as plastic ‘plugs’ in channel flows (Rosti et al. Reference Rosti, Izbassarov, Tammisola, Hormozi and Brandt2018; Izbassarov et al. Reference Izbassarov, Rosti, Brandt and Tammisola2021; Villalba et al. Reference Villalba, Daneshi, Chaparian and Martinez2023), as a phase-change phenomenon, whose consequences in dynamic behaviour are analogous to traditional jamming but different in the sense that it is induced by dynamic structural stress deformations arising from viscoelastic flow instabilities. This dependency on the locality of these plastic-induced stress deformations is all important as these depend on the local geometry and flow field, implying that the nature of this jamming transition $\phi \rightarrow \phi _J$ varies from one configuration to another. For example, for the square-root scaling in the form $\phi \sim c_1 \sqrt {\textit{Bi}} + c_2$ (see e.g. figure 3), the constants $c_1$ and $c_2$ depend on the flow configuration, $Wi$ , and $Re$ . Similarly, the exponential scaling $\phi \sim \exp \textit{Bi}$ may also be influenced by such factors, and its precise form could vary across flow configurations. In support of this, our previous work (Dzanic et al. Reference Dzanic, From and Sauret2024) found that two extensional flow benchmark problems, with the same dimensionless variables and similar flow-type distribution, result in very different dynamic responses to plasticity with different $\phi$ across a range of $\textit{Bi}$ . The consequence of this dependency, for example, is that the bulk shear rheology characterisation of the material (Cheddadi, Saramito & Graner Reference Cheddadi, Saramito and Graner2012; Varchanis et al. Reference Varchanis, Haward, Hopkins, Syrakos, Shen, Dimakopoulos and Tsamopoulos2020) – commonly performed in Couette-type geometries – will be subject to jamming dynamics different from those experienced in the actual flow configuration of the application, making their performance in practice unpredictable. Such issues in translating rheological characterisation to quantify flow performance in actual flow configurations have recently been reported as major challenges in predicting the flow performance of functional EVP materials (Bertsch et al. Reference Bertsch, Diba, Mooney and Leeuwenburgh2022). Predicting and controlling jamming is crucial in, e.g. 3-D extrusion biomaterial printing, where the jamming of soft materials during extrusion has a negative impact on both the print quality and cell viability (Xin et al. Reference Xin, Deo, Dai, Pandian, Chimene, Moebius, Jain, Han, Gaharwar and Alge2021).

4. Discussion

We have studied inertialess instabilities of EVP fluid flows in the ET regime and discovered these to transition towards jamming, featuring rich dynamics with a delicate balance between solid-like and liquid-like behaviour. We show that in the formation of spatiotemporal arrowhead structures, highly localised polymer stress regions act as local low-stress energy wells, facilitating the formation of unyielded solid regions. These localised unyielded structures, in turn, deform and redistribute stresses anisotropically, leading to an interplay between viscoelastic and plastic behaviour, which dominates and absorbs more energy than viscous dissipation. Consequently, increasing plastic effects leads to jamming transition, sharing features directly analogous to traditional jamming of dense suspensions characterised by a drastic change in flow behaviour that is slow, minimally diffusive and rigid-like with finite deformability leading to transverse loads. In particular, we find the volume fraction scales as $\phi \sim \sqrt {\textit{Bi}}$ until the ‘jamming’ phase transition $\phi _J\simeq 0.54$ where the behaviour changes dramatically, scaling as $\phi \sim \exp {\textit{Bi}}$ . Viscoelastic Kolmogorov flow in 2-D has recently been shown to differ from 3-D (Lellep et al. Reference Lellep, Linkmann and Morozov2024), suggesting limited experimental realisability of the plasticity-induced modifications to the arrowhead structures in our simulations due to 3-D dependencies. Nevertheless, key findings in this work are that shear-dependent problems in the ET regime reach a jammed state with increasing $\textit{Bi}$ , scaling as $\phi \sim \sqrt {\textit{Bi}}$ and $\phi \sim \exp {\textit{Bi}}$ beyond the jamming transitions. Moreover, we demonstrated our $\phi \sim \sqrt {\textit{Bi}}$ scaling to hold for inertia-dominated turbulent flows by Abdelgawad et al. (Reference Abdelgawad, Cannon and Rosti2023) (supplementary material, figure S5) and, as such, suspect our findings to hold in shear flows where elasticity and inertia are dominant, i.e. within the elasto-inertial turbulence regime. The exact interactions between the flow and elastic scales with the plastic events remain unclear, and more work is required to understand the effect of elasticity $Wi$ on the scaling behaviour beyond the jamming transition $\phi \gt \phi _J$ . In particular, a rigorous understanding of the $\phi \sim \exp \textit{Bi}$ scaling may benefit from asymptotic analysis or dimensionless modelling in the limit of large $\textit{Bi}$ .

The complex intermittent nature of inertialess EVP fluid flows makes their applications unpredictable and difficult to control, such as with the formation of plugs leading to increased flow resistance in practice. Indeed, previous works on EVP fluid flows describe the increase in volume fraction as a general effect of increasing the Bingham number $\textit{Bi}$ . However, no previous work has connected this phenomenon to jamming – this connection crucially reclassifies it as a phase-change phenomenon – and the analogous features which describe the impact of this transition. The dependence of plasticity-induced stress deformations on local conditions implies that the jamming behaviour observed in one configuration may differ significantly in another. Consequently, bulk shear rheology measurements may fail to predict the jamming dynamics and flow performance of EVP materials in more application-relevant configurations – ranging from industrial food, cosmetic and mining processes (e.g. moulding, extrusion, silo clogging, etc.) to emerging technologies (e.g. printing biomaterials), where unexpected jamming can lead to extreme events, disrupt performance or lead to critical failures.

Supplementary material and movies

Supplementary material, movies and Computational Notebook files are available at https://doi.org/10.1017/jfm.2025.10458. Computational Notebooks can also be found online at https://www.cambridge.org/S0022112025104588/JFM-Notebooks.

Acknowledgements

The authors acknowledge the assistance given by Research IT and the use of the Computational Shared Facility at The University of Manchester, and the High-Performance Computing facilities at Queensland University of Technology. The authors thank the anonymous reviewers whose feedback contributed to the overall improvement of the final manuscript.

Funding

E.S. is supported by the Australian Research Council (ARC-FT200100446).

Declaration of interests

The authors report no conflict of interest.

Data availability statement

The data that support the findings of this study are openly available in the JFM Notebook, reproducing figures 2 and 3.

References

Abdelgawad, M., Cannon, I. & Rosti, M. 2023 Scaling and intermittency in turbulent flows of elastoviscoplastic fluids. Nat. Phys. 19, 15.CrossRefGoogle Scholar
Balmforth, N.J., Frigaard, I.A. & Ovarlez, G. 2014 Yielding to stress: recent developments in viscoplastic fluid mechanics. Annu. Rev. Fluid Mech. 46, 121146.CrossRefGoogle Scholar
Berti, S., Bistagnino, A., Boffetta, G., Celani, A. & Musacchio, S. 2008 Two-dimensional elastic turbulence. Phys. Rev. E 77, 055306.CrossRefGoogle ScholarPubMed
Berti, S. & Boffetta, G. 2010 Elastic waves and transition to elastic turbulence in a two-dimensional viscoelastic Kolmogorov flow. Phys. Rev. E 82, 036314.CrossRefGoogle Scholar
Bertsch, P., Diba, M., Mooney, D.J. & Leeuwenburgh, S.C.G. 2022 Self-healing injectable hydrogels for tissue regeneration. Chem. Rev. 123 (2), 834873.CrossRefGoogle ScholarPubMed
Boffetta, G., Celani, A. & Mazzino, A. 2005 a Drag reduction in the turbulent Kolmogorov flow. Phys. Rev. E 71, 036307.CrossRefGoogle ScholarPubMed
Boffetta, G., Celani, A., Mazzino, A., Puliafito, A. & Vergassola, M. 2005 b The viscoelastic Kolmogorov flow: eddy viscosity and linear stability. J. Fluid Mech. 523, 161170.CrossRefGoogle Scholar
Bonn, D., Denn, M.M., Berthier, L., Divoux, T. & Manneville, S. 2017 Yield stress materials in soft condensed matter. Rev. Mod. Phys. 89, 035005.CrossRefGoogle Scholar
Cates, M.E., Wittmer, J.P., Bouchaud, J.-P. & Claudin, P. 1998 Jamming, force chains, and fragile matter. Phys. Rev. Lett. 81, 18411844.CrossRefGoogle Scholar
Cheddadi, I., Saramito, P. & Graner, F. 2012 Steady Couette flows of elastoviscoplastic fluids are nonunique. J. Rheol. 56 (1), 213239.CrossRefGoogle Scholar
Cochran, J.O., Callaghan, G.L., Caven, M.J.G. & Fielding, S.M. 2024 Slow fatigue and highly delayed yielding via shear banding in oscillatory shear. Phys. Rev. Lett. 132, 168202.CrossRefGoogle ScholarPubMed
Couchman, M.M.P., Beneitez, M., Page, J. & Kerswell, R.R. 2024 Inertial enhancement of the polymer diffusive instability. J. Fluid Mech. 981, A2.CrossRefGoogle Scholar
Datta, S.S. et al. 2022 Perspectives on viscoelastic flow instabilities and elastic turbulence. Phys. Rev. Fluids 7, 080701.CrossRefGoogle Scholar
De Vita, F., Rosti, M.E., Izbassarov, D., Duffo, L., Tammisola, O., Hormozi, S. & Brandt, L. 2018 Elastoviscoplastic flows in porous media. J. Non-Newtonian Fluid Mech. 258, 1021.CrossRefGoogle Scholar
Dennin, M. 2008 Discontinuous jamming transitions in soft materials: coexistence of flowing and jammed states. J. Phys.: Condens. Matt. 20 (28), 283103.Google Scholar
Donley, G.J., Narayanan, S., Wade, M.A., Park, J.D., Leheny, R.L., Harden, J.L. & Rogers, S.A. 2023 Investigation of the yielding transition in concentrated colloidal systems via rheo-XPCS. Proc. Natl Acad. Sci. USA 120 (18), e2215517120.CrossRefGoogle ScholarPubMed
Dubief, Y., Terrapon, V.E. & Hof, B. 2023 Elasto-inertial turbulence. Annu. Rev. Fluid Mech. 55 (2023), 675705.CrossRefGoogle Scholar
Dzanic, V., From, C.S. & Sauret, E. 2022 a Assessment of polymer feedback coupling approaches in simulation of viscoelastic fluids using the lattice Boltzmann method. Comput. Fluids 246, 105629.Google Scholar
Dzanic, V., From, C.S. & Sauret, E. 2022 b Conserving elastic turbulence numerically using artificial diffusivity. Phys. Rev. E 106, L013101.CrossRefGoogle ScholarPubMed
Dzanic, V., From, C.S. & Sauret, E. 2022 c The effect of periodicity in the elastic turbulence regime. J. Fluid Mech. 937, A31.CrossRefGoogle Scholar
Dzanic, V., From, C.S. & Sauret, E. 2022 d A hybrid lattice Boltzmann model for simulating viscoelastic instabilities. Comput. Fluids 235, 105280.CrossRefGoogle Scholar
Dzanic, V., From, C.S. & Sauret, E. 2024 Influence of plasticity on inertialess viscoelastic instabilities in elongational flow regimes. Phys. Rev. Fluids 9, 063301.CrossRefGoogle Scholar
Gotoh, K. & Yamada, M. 1984 Instability of a cellular flow. J. Phys. Soc. Japan 53, 33953398.CrossRefGoogle Scholar
Groisman, A. & Steinberg, V. 2000 Elastic turbulence in polymer solution flow. Nature 405, 5355.CrossRefGoogle ScholarPubMed
Gupta, A. & Vincenzi, D. 2019 Effect of polymer-stress diffusion in the numerical simulation of elastic turbulence. J. Fluid Mech. 870, 405418.CrossRefGoogle Scholar
Izbassarov, D., Rosti, M.E., Brandt, L. & Tammisola, O. 2021 Effect of finite Weissenberg number on turbulent channel flows of an elastoviscoplastic fluid. J. Fluid Mech. 927, A45.CrossRefGoogle Scholar
Jerolmack, D.J. & Daniels, K.E. 2019 Viewing Earth’s surface as a soft-matter landscape. Nat. Rev. Phys. 1, 716730.CrossRefGoogle Scholar
Kerswell, R.R. & Page, J. 2024 Asymptotics of the centre-mode instability in viscoelastic channel flow: with and without inertia. J. Fluid Mech. 991, A13.CrossRefGoogle Scholar
Lellep, M., Linkmann, M. & Morozov, A. 2024 Purely elastic turbulence in pressure-driven channel flows. Proc. Natl Acad. Sci. USA 121 (9).CrossRefGoogle ScholarPubMed
Lewy, T. & Kerswell, R.R. 2025 Revisiting two-dimensional viscoelastic Kolmogorov flow: a centre-mode-driven transition. J. Fluid Mech. 1007, A55.CrossRefGoogle Scholar
Morozov, A. 2022 Coherent structures in plane channel flow of dilute polymer solutions with vanishing inertia. Phys. Rev. Lett. 129, 017801.Google ScholarPubMed
N’Gouamba, E., Goyon, J. & Coussot, P. 2019 Elastoplastic behavior of yield stress fluids. Phys. Rev. Fluids 4, 123301.CrossRefGoogle Scholar
Nichols, J., Guy, R.D. & Thomases, B. 2025 Period-doubling route to chaos in viscoelastic Kolmogorov flow. Phys. Rev. Fluids 10, L041301.CrossRefGoogle Scholar
Nicolas, A., Ferrero, E.E., Martens, K. & Barrat, J.-L. 2018 Deformation and flow of amorphous solids: insights from elastoplastic models. Rev. Mod. Phys. 90, 045006.CrossRefGoogle Scholar
Page, J., Dubief, Y. & Kerswell, R.R. 2020 Exact traveling wave solutions in viscoelastic channel flow. Phys. Rev. Lett. 125, 154501.CrossRefGoogle ScholarPubMed
Peterlin, A. 1961 Streaming birefringence of soft linear macromolecules with finite chain length. Polymer 2, 257264.CrossRefGoogle Scholar
Peters, I.R., Majumdar, S. & Jaeger, H.M. 2016 Direct observation of dynamic shear jamming in dense suspensions. Nature 532, 214217.Google ScholarPubMed
Petrie, C.J.S. & Denn, M.M. 1976 Instabilities in polymer processing. AIChE J. 22 (2), 209236.Google Scholar
Rosti, M.E., Izbassarov, D., Tammisola, O., Hormozi, S. & Brandt, L. 2018 Turbulent channel flow of an elastoviscoplastic fluid. J. Fluid Mech. 853, 488514.CrossRefGoogle Scholar
Saramito, P. 2007 A new constitutive equation for elastoviscoplastic fluid flows. J. Non-Newtonian Fluid Mech. 145, 114.CrossRefGoogle Scholar
Smith, B.T. & Hashmi, S.M. 2024 In situ polymer gelation in confined flow controls intermittent dynamics. Soft Matt. 20 (8), 18581868.CrossRefGoogle ScholarPubMed
Steinberg, V. 2021 Elastic turbulence: an experimental view on inertialess random flow. Annu. Rev. Fluid Mech. 53 (1), 2758.CrossRefGoogle Scholar
Varchanis, S., Haward, S.J., Hopkins, C.C., Syrakos, A., Shen, A.Q., Dimakopoulos, Y. & Tsamopoulos, J. 2020 Transition between solid and liquid state of yield-stress fluids under purely extensional deformations. Proc. Natl Acad. Sci. USA 117, 1261112617.Google ScholarPubMed
Villalba, M.E., Daneshi, M., Chaparian, E. & Martinez, D.M. 2023 Atypical plug formation in internal elastoviscoplastic fluid flows over non-smooth topologies. J. Non-Newtonian Fluid Mech. 319, 105078.CrossRefGoogle Scholar
Xin, S., Deo, K.A., Dai, J., Pandian, N.K.R., Chimene, D., Moebius, R.M., Jain, A., Han, A., Gaharwar, A.K. & Alge, D.L. 2021 Generalizing hydrogel microparticles into a new class of bioinks for extrusion bioprinting. Sci. Adv. 7 (42), eabk3087.Google ScholarPubMed
Yerasi, S.R., Picardo, J.R., Gupta, A. & Vincenzi, D. 2024 Preserving large-scale features in simulations of elastic turbulence. J. Fluid Mech. 1000, A37.CrossRefGoogle Scholar
Zhang, Y.S., Dolatshahi-Pirouz, A. & Orive, G. 2024 Regenerative cell therapy with 3D bioprinting. Science 385, 604606.CrossRefGoogle ScholarPubMed
Figure 0

Figure 1. The EVP Kolmogorov flow in the ET regime. Representative snapshots of the polymer stretching $\operatorname {tr}{\unicode{x1D63E}}$ field $x=[0,6\times 2\pi )$ and $y=[0,4\times 2\pi )$ at various Bingham numbers, from $\textit{Bi} = 0$ to $ 3$ (columns), at different instances in time (rows): from the initial steady state $\sim\!100T$ (top) to the transition $\sim\!300T$ (middle), and the statistically homogenous regime $\overline {t}\,{\gtrsim}\, 600T$ (bottom). The grey areas represent the instantaneous unyielded $\mathcal{F}=0$ regions. The first column $\textit{Bi} = 0$ corresponds to a purely viscoelastic ET. Note, the colour bar is truncated below the true maximum at $\textrm{tr}(\boldsymbol{C}) = 1500$ for visual clarity.

Figure 1

Figure 2. The solid (unyielded) volume fraction (${\textit{a}}$) time series $\phi (t)$ and its fluctuation statistics in the statistically homogeneous regime $\overline {t}\,{\gtrsim}\, 600T$, including (${\textit{b}}$) violin distribution density superimposed with box-whisker statistics relative to the median $\widetilde {\phi }$ (white line) and (${\textit{c}}$) the root-mean-squared fluctuations, where $\overline {\phi }$ is the temporal mean. The colour scheme in (${\textit{a}}$) refers to $\textit{Bi}\gt 0$ as in (${\textit{b}}$) and (${\textit{c}}$). In (${\textit{b}}$), the box plots summarise the lower and upper quartile range of $\phi (t\geqslant \overline {t})$ with violins visualising the density and shape of the distribution. Notably, the strongest and broadest distribution of fluctuations in $\phi (t\geqslant \overline {t})$ is at $\textit{Bi}=2$. The system energy balance of the spatiotemporal mean for (${\textit{d}}$) instantaneous kinetic energy (3.1), (${\textit{e}}$) the viscous dissipation $\epsilon _\nu$ and elastic dissipation $\epsilon _p$, and (${\textit{f}}$) their corresponding fluctuations (3.2). The directory including the data and the notebook that generated this figure can be accessed at https://www.cambridge.org/S0022112025104588/JFM-Notebooks/files/Figure_2/Fig2.ipynb.

Figure 2

Figure 3. Features of jamming. Spatiotemporal mean of the velocity components: (a) the mean streamise flow profile $\langle \overline {U_x} (y^*)\rangle = (K n_y)^{-1}\sum _{k=0}^{K n_y-1}\lvert \langle \overline {u_x}(y^* + k\pi )\rangle _{{x}} \rvert$ along $y^* \in [0,\pi )$, (b) $\langle \overline {u_y} (y) \rangle _{{x}}$ and (c) the spatiotemporal mean of the polymer force $\langle \boldsymbol{\overline {\boldsymbol{\nabla }\boldsymbol{\cdot }\sigma }} \rangle _{\boldsymbol{x}} = \langle \overline {\boldsymbol{F}_{p}} \rangle _{\boldsymbol{x}}$. In (a), velocity profiles are superimposed (grey dashed lines) with base sinusoidal profile scaled by the amplitude of each $\textit{Bi}$ case, i.e. $\max _y(\langle \overline {U_x} (y)\rangle )\operatorname {cos}(Ky)$. (d) Influence of plasticity on the energy injection rate per unit area, measured by flow resistance $\mathcal{R}$, the ratio between the power injected in the statistically homogeneous regime to the base laminar fixed point. (e) Volume fraction $\phi (t)$ as a function of Bingham number $\textit{Bi}$, comparing (diamonds) the temporal mean $\bar {\phi }=\langle \phi \rangle _{\overline {t}}$ numerical simulations and (circles) the theoretically approximated $\phi$ (3.3). For $\phi \lt \phi _J$, we find that $\phi \sim \sqrt {\textit{Bi}}$ (blue line) with the linear fit $\phi =0.387 \sqrt {\textit{Bi}} -9.7\times 10^{-2}$ with a squared correlation coefficient $R^2=0.991$. Beyond the jamming transition $\phi _J\simeq 0.54$ at $\textit{Bi}=2$ (red dash-dot line) $\phi \sim \exp {\textit{Bi}}$ (green line) with the linear fit $\phi =3.2\times 10^{-2} \exp {\textit{Bi}} - 0.24$ with ${R}^2=0.999$. The directory including the data and the notebook that generated this figure can be accessed at https://www.cambridge.org/S0022112025104588/JFM-Notebooks/files/Figure_3/Fig3.ipynb.

Supplementary material: File

From et al. supplementary movie 1

tr C field of purely viscoplastic Kolmogorov flow in the elastic turbulence regime.
Download From et al. supplementary movie 1(File)
File 11 MB
Supplementary material: File

From et al. supplementary movie 2

tr C field of elastoviscoplastic Kolmogorov flow in the elastic turbulence regime at Bi = 0.5.
Download From et al. supplementary movie 2(File)
File 20.2 MB
Supplementary material: File

From et al. supplementary movie 3

tr C field of elastoviscoplastic Kolmogorov flow in the elastic turbulence regime at Bi = 2.
Download From et al. supplementary movie 3(File)
File 18.4 MB
Supplementary material: File

From et al. supplementary movie 4

tr C field of elastoviscoplastic Kolmogorov flow in the elastic turbulence regime at Bi = 3.
Download From et al. supplementary movie 4(File)
File 17.3 MB
Supplementary material: File

From et al. supplementary material 5

From et al. supplementary material
Download From et al. supplementary material 5(File)
File 2.4 MB
Supplementary material: File

From et al. supplementary material 6

From et al. supplementary material
Download From et al. supplementary material 6(File)
File 701.2 KB