Hostname: page-component-54dcc4c588-64p75 Total loading time: 0 Render date: 2025-10-08T01:12:55.634Z Has data issue: false hasContentIssue false

Scaling analysis and self-similarity near breakup of elasto-viscoplastic liquid threads under creeping flow

Published online by Cambridge University Press:  07 October 2025

Pourya Zakeri
Affiliation:
Laboratory of Fluid Mechanics and Rheology, Department of Chemical Engineering, University of Patras, Patras 26504, Greece
Pantelis Moschopoulos
Affiliation:
Laboratory of Fluid Mechanics and Rheology, Department of Chemical Engineering, University of Patras, Patras 26504, Greece
Yiannis Dimakopoulos
Affiliation:
Laboratory of Fluid Mechanics and Rheology, Department of Chemical Engineering, University of Patras, Patras 26504, Greece
John Tsamopoulos*
Affiliation:
Laboratory of Fluid Mechanics and Rheology, Department of Chemical Engineering, University of Patras, Patras 26504, Greece
*
Corresponding author: John Tsamopoulos, tsamo@chemeng.upatras.gr

Abstract

We investigate theoretically the breakup dynamics of an elasto-visco-plastic filament surrounded by an inert gas. The filament is initially placed between two coaxial disks, and the upper disk is suddenly pulled away, inducing deformation due to both constant stretching and capillary forces. We model the rheological response of the material with the Saramito–Herschel–Bulkley (SHB) model. Assuming axial symmetry, the mass and momentum balance equations, along with the constitutive equation, are solved using the finite element framework PEGAFEM-V, enhanced with adaptive mesh refinement with an underlying elliptic mesh generation algorithm. As the minimum radius decreases, the breakup dynamics accelerates significantly. We demonstrate that the evolution of the minimum radius, velocity and axial stress follow a power-law scaling, with the corresponding exponent depending on the SHB shear-thinning parameter, $n$. The scaling exponents obtained from our axisymmetric simulations under creeping flow are verified through asymptotic analysis of the slender filament equations. Our findings reveal three distinct breakup regimes: (a) elasto-plastic, (b) elasto-plasto-capillary, both with finite-time breakup for $n\lt 1$, and (c) elasto-plasto-capillary with no finite-time breakup for $n=1$. We show that self-similar solutions close to filament breakup can be achieved by appropriate rescaling of length, velocity and stress. Notably, the effect of the yield stress becomes negligible in the late stages of breakup due to the local dominance of high elastic stresses. Moreover, the scaling exponents are independent of elasticity, resembling the breakup behaviour of finite extensible viscoelastic materials.

Information

Type
JFM Papers
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

Yield stress (YS) fluids are a class of complex materials that exhibit dual behaviour depending on the magnitude of applied stress. When this magnitude is below a certain threshold they behave like solids, but when it exceeds this threshold, they deform like liquids. This critical threshold is known as the YS (Frigaard Reference Frigaard2019). The boundary separating the solid and fluid regions is referred to as the yield surface. The yield surface becomes particularly significant in free-surface flows, and its accurate determination is essential in several scenarios. For example, during the pinch-off of a viscoplastic filament it influences both the final filament shape and the precise location of the pinch point (Moschopoulos et al. Reference Moschopoulos, Syrakos, Dimakopoulos and Tsamopoulos2020). Moreover, in the flow of a YS fluid down an inclined plate it governs the evolution of the free surface, and its final shape (Chambon et al. Reference Chambon, Freydier, Naaim and Vila2020), while in the viscoplastic dam break set-up it determines the final arrested interface shape (Liu et al. Reference Liu, Balmforth, Hormozi and Hewitt2016). Understanding the flow characteristics of YS materials is crucial, because they are encountered in nature, as in lava and wet avalanches (Freydier, Chambon & Naaim Reference Freydier, Chambon and Naaim2017), and in personal care, cosmetics, food, building, oil and other industries, with products like shaving cream (Huisman, Friedman & Taborek Reference Huisman, Friedman and Taborek2012), toothpaste, ice cream, mayonnaise (Ma & Barbosa-Cánovas, Reference Ma and Barbosa-Chovas1995), fresh concrete and crude oil (Dimitriou & McKinley Reference Dimitriou and McKinley2014). Several models have been developed to describe the transition from solid-like to fluid-like behaviour. The first such model was the Bingham model, which is a generalised Newtonian fluid model with constant viscosity upon yielding. Later, to account for material shear thinning, the Herschel–Bulkley model was introduced (Frigaard Reference Frigaard2019). Despite their relative simplicity, these two models require computationally expensive algorithms to accurately resolve regions where the magnitude of the strain-rate tensor approaches zero (Moschopoulos et al. Reference Moschopoulos, Varchanis, Syrakos, Dimakopoulos and Tsamopoulos2022), otherwise they require a regularisation parameter to avoid the stress singularity developing as the yield surface is approached.

Over the past two decades, several experimental studies have attempted to clarify the behaviour of YS materials using Carbopol (Putz et al. Reference Putz, Burghelea, Frigaard and Martinez2008; Niedzwiedz et al. Reference Niedzwiedz, Arnolds, Willenbacher and Brummer2009; Freydier et al. Reference Freydier, Chambon and Naaim2017; Luu, Philippe & Chambon Reference Luu, Philippe and Chambon2017), an anionic, high molecular weight polymer composed of acrylic acid cross-linked with an ether, which is transparent, making it more appropriate for experimental observations and measurements. Other experimental studies with YS materials have observed distinct patterns, which are characteristic of viscoelastic fluids, such as the inverted teardrop shape of a rising bubble (Mougin, Magnin & Piau Reference Mougin, Magnin and Piau2012) or the generation of a negative wake by a sedimenting sphere (Putz et al. Reference Putz, Burghelea, Frigaard and Martinez2008; Holenberg et al. Reference Holenberg, Lavrenteva, Shavit and Nir2012). To capture theoretically these observations, an elastic component had to be incorporated into the two original YS fluid models by Saramito (Reference Saramito2007, Reference Saramito2009). These new models, known as elasto-visco-plastic (EVP) models, have been used to successfully explain the above-mentioned observations and several others. For example, Fraggedakis, Dimakopoulos & Tsamopoulos (Reference Fraggedakis, Dimakopoulos and Tsamopoulos2016) successfully captured the negative wake and the loss of fore-and-aft symmetry in a sedimenting spherical particle in Carbopol. Moreover, Moschopoulos et al. (Reference Moschopoulos, Spyridakis, Varchanis, Dimakopoulos and Tsamopoulos2021) and Kordalis, Dimakopoulos & Tsamopoulos (Reference Kordalis, Dimakopoulos and Tsamopoulos2024) captured the teardrop shape during the rise of a single or a pair of bubbles in strain-rate-thinning EVP fluids. More recently, Esposito, Dimakopoulos & Tsamopoulos (Reference Esposito, Dimakopoulos and Tsamopoulos2024) studied a buoyancy driven motion of a drop in an EVP material.

The buoyancy-driven motions mentioned above involve mixed, shear and elongation flows. To better isolate and control extensional flow, Varchanis et al. (Reference Varchanis, Haward, Hopkins, Syrakos, Shen, Dimakopoulos and Tsamopoulos2020a ) studied experimentally the flow of a Pluronic aqueous solution (another EVP fluid) in the optimised shape cross-slot extensional rheometer, conducted simulations using Saramito’s model (Saramito Reference Saramito2009) and obtained excellent agreement between experiments and simulations. Filament stretching is another procedure to create a flow field dominated by extension. The way the filament deforms, and breaks is of paramount importance for many industrial processes, such as coating flows, food processing, spraying of pesticides, direct ink-jet printing and additive manufacturing. In all of them, EVP fluids may be involved (Van Der Kolk, Tieman & Jalaal Reference Van Der Kolk, Tieman and Jalaal2023). The easiest way to isolate such a configuration is to place a small amount of material between two coaxial disks of equal radius and pull the upper disk vertically upward. The material elongates, thins and, finally, pinches off primarily driven by capillarity. To probe the filament stretching dynamics, especially when very short relaxation times need to be measured, several instruments have been developed, like the capillary breakup extensional rheometer (Arnolds et al. Reference Arnolds, Buggisch, Sachsenheimer and Willenbacher2010) (CaBER), the Rayleigh Ohnesorge jetting extensional rheometer (Keshavarz et al. Reference Keshavarz, Sharma, Houze, Koerner, Moore, Cotts, Threlfall-Holmes and McKinley2015) and dripping onto substrate (Rosello et al. Reference Rosello, Sur, Barbet and Rothstein2019). The process of filament stretching may be separated into the initial or bulk dynamics and the final or pinching dynamics. In the following we will briefly review the literature that is related either to filament stretching of non-Newtonian, and particularly YS fluids, or to the pinching dynamics.

Filament breakup has been studied with Newtonian (Zhang, Padgett & Basaran Reference Zhang, Padgett and Basaran1996) and generalised Newtonian fluids (Yildirim & Basaran Reference Yildirim and Basaran2001) . In both types of materials, a short neck is formed, and the filament breaks in finite time. Viscoelastic materials have been studied quite extensively and early on under extensional flow (McKinley & Sridhar Reference McKinley and Sridhar2002), due to their widespread use and importance in practical applications, as well as the intriguing phenomena that arise. When a viscoelastic filament thins, strong tensile stresses develop that oppose the necking of the fluid thread. This increases the lifetime of the bridge and leads to the formation of a quite longer and cylindrical thread that connects the two drops that remain in contact with the corresponding disk (Clasen et al. Reference Clasen, Eggers, Fontelos, Li and McKinley2006). Under specific material properties and in the presence of elastic and inertia effects a small satellite drop may develop in the middle of the filament (Bhat et al. Reference Bhat, Appathurai, Harris, Pasquali, McKinley and Basaran2010; Varchanis et al. Reference Varchanis, Dimakopoulos, Wagner and Tsamopoulos2018).

Turning our attention to filaments of YS materials, initially Martinie, Buggisch & Willenbacher (Reference Martinie, Buggisch and Willenbacher2013), and Niedzwiedz, Buggisch & Willenbacher (Reference Niedzwiedz, Buggisch and Willenbacher2010) tried to measure the so-called extensional YS using the CaBER. However, their measurements did not agree with the ideal viscoplastic theory that predicted the extensional YS to be $\sqrt 3$ times the shear YS. This deviation was caused because the elastic stresses in the von Mises criterion were neglected (Varchanis et al. Reference Varchanis, Haward, Hopkins, Syrakos, Shen, Dimakopoulos and Tsamopoulos2020a ). Furthermore, Balmforth, Dubash & Slim (Reference Balmforth, Dubash and Slim2010a ) investigated the primarily extensional flow taking place in drips and bridges of different YS materials and compared their experimental results with the viscoplastic slender theory they developed. Then, they used this theory to study the pinching dynamics of YS materials (Balmforth, Dubash & Slim Reference Balmforth, Dubash and Slim2010b ). Interestingly, they showed that, just before pinching, the breakup dynamics became universal, and was dictated by a simple power law, the exponent of which coincided with the exponent of power-law fluids (Doshi et al. Reference Doshi, Suryo, Yildirim, McKinley and Basaran2003; Suryo & Basaran Reference Suryo and Basaran2006), and that the YS had almost no effect on the scaling law. On the other hand, Huisman et al. (Reference Huisman, Friedman and Taborek2012) and Niedzwiedz et al. (Reference Niedzwiedz, Buggisch and Willenbacher2010) observed experimentally that, although YS stress materials followed the scaling laws of shear-thinning materials, they also formed a longer cylindrical neck connecting the upper and lower conical parts of the bridge, without examining whether the material in the bridge actually exhibited elastic properties. More recently, Moschopoulos et al. (Reference Moschopoulos, Syrakos, Dimakopoulos and Tsamopoulos2020) performed two-dimensional simulations assuming axial symmetry of the bulk and pinching dynamics of a thread with viscoplastic properties. Thus, they confirmed the above-mentioned scaling laws for viscoplastic materials (Balmforth et al. Reference Balmforth, Dubash and Slim2010b ). However, their simulations did not capture the cylindrical neck that had been experimentally observed, (Niedzwiedz et al. Reference Niedzwiedz, Buggisch and Willenbacher2010). This shortcoming was alleviated by Moschopoulos et al. (Reference Moschopoulos, Kouni, Psaraki, Dimakopoulos and Tsamopoulos2023), when they included material elasticity by using the Saramito–Herschel–Bulkley model and predicted the formation of a long and slender neck that connected the upper and lower parts of the Carbopol filament undergoing stretching. Nevertheless, they did not investigate the pinching dynamics in detail and whether it is governed by similarity solutions. Such analysis has been developed for Newtonian, viscoplastic and viscoelastic materials.

Understanding the underlying physics during pinching (Tuladhar & Mackley Reference Tuladhar and Mackley2008; Townsend et al. Reference Townsend, Beck, Gehrke, Berkland and Detamore2019; Van Der Kolk et al. Reference Van Der Kolk, Tieman and Jalaal2023) is crucial for advancing industrial tools, notably those used in inkjet or 3D printing, for controlling the droplet size, which may be created after filament breakup or reducing food residuals during the filling process in designing production lines. While stretching the fluid thread, the mass conservation and capillarity on its free surface shrink the filament radius to zero somewhere along the filament. At this location and time instant the governing equations become singular. Thus, stress and velocity diverge, making the dynamics independent of the boundary conditions governing the earlier bulk dynamics. Hence, the local dynamics in the abruptly thinning filament will exhibit universality, and self-similarity could be achieved at the vicinity of the singular point. This idea of universality was proposed first by Keller & Miksis (Reference Keller and Miksis1983), assuming potential flow in the filament. Since then, various authors have described the pinch-off dynamics of Newtonian liquids (Eggers Reference Eggers1993; Eggers & Dupont Reference Eggers and Dupont1994; Papageorgiou Reference Papageorgiou1995; Brenner, Lister & Stone Reference Brenner, Lister and Stone1996; Chen, Notz & Basaran Reference Chen, Notz and Basaran2002), with or without inertia. Concerning complex fluids, numerous studies have been conducted following and extending the ideas for Newtonian liquids to describe pinching when the rheological model of the filament is a power law (Doshi et al. Reference Doshi, Suryo, Yildirim, McKinley and Basaran2003; Renardy & Renardy Reference Renardy and Renardy2004; Suryo & Basaran Reference Suryo and Basaran2006), viscoelastic (Bechtel, Cao & Forest Reference Bechtel, Cao and Forest1992; Anna & McKinley Reference Anna and McKinley2001; Renardy Reference Renardy2002b ; Fontelos & Li Reference Fontelos and Li2004; Clasen et al. Reference Clasen, Eggers, Fontelos, Li and McKinley2006; Turkoz et al. Reference Turkoz, Lopez-Herrera, Eggers, Arnold and Deike2018) or viscoplastic (Balmforth et al. Reference Balmforth, Dubash and Slim2010b ; Moschopoulos et al. Reference Moschopoulos, Syrakos, Dimakopoulos and Tsamopoulos2020).

However, efforts to describe the bulk and pinch-off dynamics of a bridge composed of an EVP material are very limited. We are aware only of the work by Moschopoulos et al. (Reference Moschopoulos, Kouni, Psaraki, Dimakopoulos and Tsamopoulos2023), who studied the bulk dynamics thoroughly and filament thinning only to some extent but did not examine the asymptotics or self-similarity close to breakup. The present study aims to fill this gap by determining the breakup asymptotics, the existence of self-similar solutions by employing two-dimensional (2-D) axisymmetric simulations coupled with a sharp interface tracking method, while fluid inertia is neglected.

The paper is organised as follows: in § 2, the problem formulation is presented, the equations and boundary conditions governing the axisymmetric flow are introduced, followed by the slender filament (1-D) equations. In § 3, we very briefly review the solution method, postponing to Appendix A certain important techniques about mesh generation, refinement, restructuring and variable interpolation between different meshes. We dedicate § 4 to extracting asymptotic pinch-off scales from the slender filament equations. In the same section we include the approximate solution under the assumption that the filament undergoes a strictly uniaxial extension. In § 5, we present our 2-D results concerning filament pinching, self-similar solutions,- and we conduct a parametric study. Finally, we summarise our findings in § 6.

2. Problem formulation

2.1. General two-dimensional formulation

In what follows, variables or parameters with tilde (˜) over them, are denoted as dimensional while their dimensionless counterparts are presented without tilde. We consider a 2-D, axisymmetric liquid bridge of an EVP material held captive between two parallel and coaxial disks. The initial separation of the two disks is $\tilde{L}_{0}$ and their common radius is $\tilde{h}_{0}$ , both of which are of the order of a few millimetres, at the most, making gravitational effects very small from the beginning of the process. The material is pinned to the disk’s perimeter, so that the initial radius of the fluid thread is also $\tilde{h}_{0}$ . At time $\tilde{t}=0^{+}$ , the upper plate starts to move with constant velocity in the positive $\tilde{z}$ -direction. As time proceeds, the capillary force and material stress components start to grow, and a neck is formed in the filament. We monitor the time evolution of the minimum radius of the filament as it shrinks to zero, $\tilde{h}_{\textit{min}}\rightarrow 0$ , along with the dominant velocity and stress components. The negligible gravity force allows us to assume a symmetry plane at $\tilde{z}=0$ , making it possible to simulate only the top half of the filament. Figure 1 shows a schematic of the axisymmetric EVP bridge being stretched with symmetry plane imposed at $\tilde{z}=0$ .

Figure 1. Schematic of the axisymmetric EVP filament stretching geometry.

The characteristic scales used for making the variables dimensionless are those proposed by Moschopoulos et al. (Reference Moschopoulos, Kouni, Psaraki, Dimakopoulos and Tsamopoulos2023) for the bulk dynamics of an EVP filament stretching. The radius of the disk $\tilde{h}_{0}$ and the capillary force $({\tilde{\sigma }}/{\tilde{h}_{0}})$ , scale all lengths and pressure or stress components, respectively. Balancing capillarity and viscous stresses results in the visco-capillary time, which is used as the characteristic time scale, $\tilde{t}_{vc}=({\tilde{k}\tilde{h}_{0}}/{\tilde{\sigma }})^{({1}/{n})}$ , where $\tilde{k}\, ({\textrm{Pa s}}^{n})$ is the consistency index, and $n$ is the shear-thinning exponent of the Saramito–Herschel–Bulkley (SHB) model. The velocity scale is ${\tilde{h}_{0}}/{\tilde{t}_{vc}}$ . Another form of stress scale, equivalent to the one given earlier, is ${\tilde{k}}/{{(\tilde{t}_{vc}})^{n}}$ . Introducing these characteristic scales in the governing equations results in the dimensionless numbers for the present analysis, which are given in table 1.

Table 1. Dimensionless numbers.

With these characteristic quantities, the dimensionless mass and momentum conservation equations become

(2.1) \begin{align}& Oh^{-2}\frac{D\boldsymbol{u}}{Dt} =-\boldsymbol{\nabla }p+\boldsymbol{\nabla }\boldsymbol{\cdot }\boldsymbol{\tau }+\boldsymbol{\nabla }\boldsymbol{\cdot }\boldsymbol{\tau }_{\boldsymbol{s}}-Bo\, \boldsymbol{e}_{\boldsymbol{z}}, \end{align}
(2.2) \begin{align}&\qquad\qquad\qquad\qquad \boldsymbol{\nabla }\boldsymbol{\cdot }\boldsymbol{u} =0, \end{align}

where ${\textrm{D}}/{{\textrm{D}}t}$ is the substantial derivative, $\boldsymbol{u}$ the velocity vector, $\boldsymbol{\nabla}$ the gradient operator, $\boldsymbol{\tau }_{\boldsymbol{s}}$ the Newtonian solvent stress, $\boldsymbol{\tau }$ the polymeric extra stress tensor, p the pressure and $\boldsymbol{e}_{\boldsymbol{z}}$ the unit vector in the positive z-direction. In this study, the Newtonian solvent contribution is taken to be zero, hence $\boldsymbol{\tau }_{\boldsymbol{s}}=\textbf{0}$ . Furthermore, in this study we will neglect inertial effects, setting $Oh^{-2}=0$ . The polymeric extra stress tensor is described by the SHB (Saramito Reference Saramito2009) EVP model for the rheological response of the material

(2.3) \begin{equation}Ec\, \overset{\boldsymbol{\nabla }}{\boldsymbol{\tau }}+\max \!\left(0,\left| \boldsymbol{\tau}_{\boldsymbol{d}}\right| -Y_{s}\right)^{\frac{1}{n}}\frac{\boldsymbol{\tau }}{\left| \boldsymbol{\tau}_{\boldsymbol{d}}\right| }=\dot{\boldsymbol{\gamma }}.\end{equation}

The first term in this model is the elastic contribution and the second one is the viscoplastic contribution to the total rate of deformation tensor, $\dot{\boldsymbol{\gamma }}=\boldsymbol{\nabla }\boldsymbol{u}+(\boldsymbol{\nabla }\boldsymbol{u})^{T}$ . In the above model, $\overset{\boldsymbol{\nabla }}{\boldsymbol{\tau }}$ is the upper-convected derivative of extra stress tensor, which is defined as

(2.4) \begin{equation}\overset{\boldsymbol{\nabla }}{\boldsymbol{\tau }}=\frac{\partial \boldsymbol{\tau }}{\partial t}+\boldsymbol{u}\boldsymbol{\cdot }\boldsymbol{\nabla }\boldsymbol{\tau }-\left(\boldsymbol{\nabla }\boldsymbol{u}\right)^{T}\boldsymbol{\cdot }\boldsymbol{\tau }-\boldsymbol{\tau }\boldsymbol{\cdot }\boldsymbol{\nabla }\boldsymbol{u}. \end{equation}

Also, $\boldsymbol{\tau}_{\boldsymbol{d}}=\boldsymbol{\tau}-(tr(\boldsymbol{\tau })/tr(\boldsymbol{I}))$ is the deviatoric part of extra stress tensor with magnitude $| \boldsymbol{\tau} _{\boldsymbol{d}}| =\sqrt{({1}/{2})\boldsymbol{\tau }_{\boldsymbol{d}}\colon \boldsymbol{\tau }_{\boldsymbol{d}}}$ , and $(\boldsymbol{\nabla }\boldsymbol{u})^{T}$ indicates the transpose of the velocity gradient. The max term determines the transition from a solid-like to a fluid-like region by incorporating the von Mises criterion. If $| \boldsymbol{\tau}_{\boldsymbol{d}}|$ exceeds $Y_{s}$ , the material yields and deforms like a viscoelastic fluid with shear thinning, otherwise it behaves like a neo-Hookean solid. The free surface of the filament, $S_{f}$ , is in contact with air, which is assumed to have negligible viscosity and density. In our formulation, nodes on $S_{f}$ are treated in a Lagrangian sense, hence the velocity of the mesh and the velocity of the fluid particle on $S_{f}$ in the direction normal to $S_{f}$ must be equal. This is the so-called kinematic boundary condition and serves as a boundary condition on the free surface. Furthermore, the total stress from the fluid side acting on $S_{f}$ must balance capillarity, given that air is dynamically inert. This is the so-called traction boundary condition. These two boundary conditions are

(2.5) \begin{align} \boldsymbol{n}\boldsymbol{\cdot }\left(\boldsymbol{u}-\boldsymbol{u}_{m}\right) = 0,\quad & {\textit{on}}\, S_{f}, \end{align}
(2.6) \begin{align} \boldsymbol{n}\boldsymbol{\cdot}\boldsymbol{T} =\left(2\mathcal{H}\right)\boldsymbol{n},\quad & {\textit{on}}\, S_{f}, \end{align}

where $\boldsymbol{u}_{m}$ is the velocity of the mesh nodes and $\boldsymbol{n}$ the outward unit vector normal to $S_{f}$ ; T the total stress tensor, defined as $\boldsymbol{T}=-p\boldsymbol{I}+\boldsymbol{\tau }$ , with $\boldsymbol{I}$ the unit tensor; $2\mathcal{H}=-\boldsymbol{\nabla} _{s}\boldsymbol{\cdot }\boldsymbol{n}$ is twice the mean curvature of $S_{f}$ and lastly $\boldsymbol{\nabla} _{s}=(\boldsymbol{I}-\boldsymbol{nn})\boldsymbol{\cdot }\boldsymbol{\nabla} \text{is the surface gradient operator}$ (Deen Reference Deen2012). We have assumed axial symmetry with axis at $r$ = 0, or on $S_{A}$ and planar symmetry at $z=0$ , or on $S_{p}$ . On these boundaries the no-shear and no-penetration boundary conditions are applied

(2.7) \begin{align} \boldsymbol{n}\boldsymbol{\cdot }\boldsymbol{T}\boldsymbol{\cdot }\boldsymbol{t} = 0, &\quad {\textrm{on}}\,S_{A}\, {\textrm{and}}\, S_{P}, \end{align}
(2.8) \begin{align} \boldsymbol{n}\boldsymbol{\cdot }\boldsymbol{u} = 0, &\quad {\textrm{on}}\,S_{A}\, {\textrm{and}}\, S_{P}, \end{align}

where $\boldsymbol{t}$ is the unit vector tangential to either $S_{A}$ or $S_{p}$ . The fluid in contact with the two solid disks, initially at rest, follows the no-slip and no-penetration boundary conditions and when t > 0, the upper solid disk starts to move suddenly with dimensionless velocity U

(2.9) \begin{align}& \begin{cases} t=0, \quad {\boldsymbol{u}}=0\\ t \gt 0, \quad {\boldsymbol{u}}=U{\boldsymbol{e}}_{\boldsymbol{z}} \end{cases} \text{at}\quad z=L, \end{align}
(2.10) \begin{align}&\qquad\quad \boldsymbol{u}=\textbf{0}, \quad \text{at} \quad z=0,\end{align}

where $L=L_{0}+Ut$ is the disk separation at time t. The contact point on the intersection between the free surface $S_{f}$ and the solid disks is pinned

(2.11) \begin{equation}h_{0}=1, \quad \text{at}\quad z=L\, \quad {\textrm{and}}\, \quad z=0.\end{equation}

Due to the hyperbolic nature of the SHB model (and similarly for most other constitutive laws for polymeric liquids), the boundary conditions for the extra stress tensor are needed only at the inflow (Van Der Zanden & Hulsen Reference Van Der Zanden and Hulsen1988). In this problem, no inflow conditions exist, and no stress boundary conditions are required, except for $\boldsymbol{n}\boldsymbol{\cdot }\boldsymbol{T}\boldsymbol{\cdot }\boldsymbol{t}=0$ at the axis of symmetry and the symmetry plane along with the force balance at the free interface.

2.2. Slender thread equations

Assuming that to highest order the variation of all flow variables in the radial direction, $r$ , is negligible compared with that in the axial direction, $z$ (except for $u_{r},\, {\textrm{and}}\, \tau _{rz}$ , which vary linearly with $r$ ), the interface $S_{f}$ can be described as a function of $z$ alone, $S_{f}, r=h(z,t)$ , disallowing any overturning of the free surface. Then the previously presented 2-Dl equations can be reduced to a much simpler, 1-D equation set. These are the so-called slender filament equations. The reduced mass and momentum equations, still neglecting solvent viscosity, are (Bechtel et al. Reference Bechtel, Cao and Forest1992; Clasen et al. Reference Clasen, Eggers, Fontelos, Li and McKinley2006)

(2.12) \begin{align}&\qquad\qquad\qquad\quad \frac{\partial h^{2}}{\partial t}+\frac{\partial }{\partial z}\big(u_{z}h^{2}\big)=0, \end{align}
(2.13) \begin{align}& Oh^{-2}\left(\frac{\partial u_{z}}{\partial t}+u_{z}\frac{\partial u_{z}}{\partial z}\right)=\frac{\partial }{\partial z}(2\mathcal{H})+\frac{1}{h^{2}}\frac{\partial }{\partial z}\big(h^{2}(\tau _{\textit{zz}}-\tau _{\textit{rr}})\big), \end{align}
(2.14) \begin{align}&\qquad\qquad\qquad 2\mathcal{H}=\frac{h_{\textit{zz}}}{\left(1+h_{z}^{2}\right)^{\frac{3}{2}}}-\frac{1}{h\left(1+h_{z}^{2}\right)^{\frac{1}{2}}}, \end{align}

where $u_{z}(z,t)$ is the axial velocity, $2\mathcal{H}$ is twice the mean curvature of the free surface, now explicitly written in the cylindrical coordinate system and the subscript $z$ next to $h$ , as in $h_{z}$ , indicates partial differentiation; in this case, $({\partial h}/{\partial z}).$ Please note that in § 2.2 all variables have retained their usual symbols but depend only in the axial direction $z$ and time, $t$ . Equation (2.13) is the reduced form of the dominant momentum balance in the axial direction, in which pressure has been eliminated in favour of the curvature term via the normal force balance and the next-order term of the axial velocity has been eliminated via the tangential force balance. As stated in the previous section, we will analyse the slender filament equations ignoring inertia. Although the computational effort to solve the 1-D equations is considerably smaller than the effort to solve their 2-D counterparts, the former equations are not accurate enough to capture the flow field under certain conditions (Suryo & Basaran Reference Suryo and Basaran2006; Turkoz et al. Reference Turkoz, Lopez-Herrera, Eggers, Arnold and Deike2018). In this study, we avoid conducting 1-D simulations; we will use the slender filament equations only to derive the asymptotic decay of the radius and growth of the stress in the pinch-off regime.

3. Finite element solution

Our numerical algorithm for solving the 2-D equation set is based on the recently developed Petrov–Galerkin finite element formulation for viscoelastic free-surface flows. This method allows us to use equal-order linear interpolants for all variables, circumventing the requirement for mixed finite element formulations for velocity, pressure and stress to satisfy the Ladyzhenskaya–Babuška–Brezzi conditions, and simultaneously does not suffer from the high Weissenberg number problem. To achieve all these advantages, Varchanis et al. (Reference Varchanis, Syrakos, Dimakopoulos and Tsamopoulos2019b ) have carefully chosen specific stabilising terms for the weak form of the governing equations. Coupling this method with the quasi-elliptic mesh generator (Dimakopoulos & Tsamopoulos Reference Dimakopoulos and Tsamopoulos2003; Chatzidai et al. Reference Chatzidai, Giannousakis, Dimakopoulos and Tsamopoulos2009) allows us to obtain stable numerical solutions even in highly deformed meshes and for very high values of fluid elasticity (Varchanis et al. Reference Varchanis, Syrakos, Dimakopoulos and Tsamopoulos2020b ). This method has been tested in benchmark flows and many other studies with steady and moving boundary problems, confirming its accuracy and robustness (Moschopoulos et al. Reference Moschopoulos, Spyridakis, Varchanis, Dimakopoulos and Tsamopoulos2021; Varchanis et al. Reference Varchanis, Pettas, Dimakopoulos and Tsamopoulos2021; Kordalis et al. Reference Kordalis, Pema, Androulakis, Dimakopoulos and Tsamopoulos2023; Moschopoulos et al. Reference Moschopoulos, Spyridakis, Dimakopoulos and Tsamopoulos2024; Kordalis et al. Reference Kordalis, Dimakopoulos and Tsamopoulos2024)

To solve the set of partial differential equations in this problem, the domain is discretised into elements, altogether forming the mesh. According to Moschopoulos et al. (Reference Moschopoulos, Kouni, Psaraki, Dimakopoulos and Tsamopoulos2023), we prefer linear three-node triangular elements to linear quadrangles for their better representation of the free surface. Additionally, we utilise right-angle triangles for mesh construction, as they exhibit better convergence behaviour compared with other types of domain triangulation methods. For accurately tracking the displacement of free surface, we employ the arbitrary Lagrangian Eulerian formulation. In this method the normal displacement of the free surface is dictated by its Lagrangian movement, whereas all other nodes inside the liquid domain are positioned arbitrarily to achieve a smooth mesh for solution accuracy. To optimise the mesh within the domain, we selected the quasi-elliptic grid generation scheme proposed by Dimakopoulos & Tsamopoulos (Reference Dimakopoulos and Tsamopoulos2003). As suggested in previous studies (Notz & Basaran Reference Notz and Basaran2004; Suryo & Basaran Reference Suryo and Basaran2006; Bhat et al. Reference Bhat, Appathurai, Harris, Pasquali, McKinley and Basaran2010), using a highly non-uniform mesh where elements are densely concentrated in the necking region is essential for accurate and efficient computation of breakup problems. However, even in a densely packed mesh generated by the quasi-elliptic grid generation scheme, the elements will become highly distorted, leading to divergence of the numerical algorithm as $h_{\textit{min}}\rightarrow 0$ . To address this issue, we introduce additional nodes on radial cross-sections to maintain the quality of the mesh as breakup is approached. Computational experiments indicate that full domain remeshing introduces significantly larger errors during solution transfer from the previous to the new mesh than simply adding nodes. Therefore, we do not use full domain remeshing. This approach allows us to start computations with a relatively coarse mesh and progressively refine it as $h_{\textit{min}}\rightarrow 0$ , while ensuring a smooth mesh and an accurate solution.

For the time integration and iterative scheme to solve the matrix system assembled by the finite element formulation, we employ the same procedure reported by Moschopoulos et al. (Reference Moschopoulos, Kouni, Psaraki, Dimakopoulos and Tsamopoulos2023). Convergence of Newton’s iterations at each time step is accepted if the Euclidean norm of both the residual and the Newton’s correction vector falls below of $10^{-6}$ . Certain important ideas about mesh refinement and variable interpolation between meshes are given in Appendix A. In the supplementary material SM1, we present the evolution of the mesh either for the entire domain, or in the neck where continuous mesh refinement takes place.

4. Asymptotic analysis for an EVP thread

4.1. Asymptotic analysis near filament pinching at the symmetry plane via the slender thread equations following the SHB model

Since we are seeking a self-similar solution in the vicinity of the singular point, where $h_{\textit{min}}\rightarrow 0$ , we introduce two new local variables, $Z$ and $T$ , which are defined as

(4.1) \begin{align} Z&=z-z_{0}, \end{align}
(4.2) \begin{align} T&=t_{0}-t. \end{align}

In the above expressions, $Z$ denotes the axial distance from the singular point (the location of $h_{\textit{min}})$ , the axial coordinate of which is $z_{0}$ . Having set $Oh^{-2}=0$ , to neglect inertia in the momentum equation, we find $z_{0}=0$ in all our simulations and $z$ is probed where $h=1.01h_{\textit{min}}$ . Here, $T$ is the time until the breakup time, which is denoted as $t_{0}$ . Preliminary 2-D results indicate that the time evolution of the minimum radius follows the pattern shown in figure 2.

Figure 2. The evolution of minimum radius vs. time, for parameter values $n=0.45,\, Ec=0.14,\, Y_{s}=1.25,\, U=0.39$ .

This figure clearly demonstrates that the breakup time of the filament is finite, a trend observed in all our simulations for values of $n$ less than unity. This should have been expected, because the SHB model predicts a finite extensible material, which is known to have a finite breakup time as opposed to the Oldroyd-B fluid, which approaches breakup exponentially (Renardy Reference Renardy1995; McKinley Reference McKinley2005). Hence, we introduce self-similar forms with power-law functions for the axial position, axial velocity and axial component of extra stress tensor

(4.3) \begin{align} h\left(Z,T\right)&=T^{{\alpha _{1}}}\, \phi \left(\xi \right), \end{align}
(4.4) \begin{align} u_{z}\left(Z,T\right)&=T^{{\alpha _{2}}}\, \psi \left(\xi \right), \end{align}
(4.5) \begin{align} \tau _{\textit{zz}}\left(Z,T\right)&=T^{{\alpha _{3}}}{\Omega} \left(\xi \right), \end{align}
(4.6) \begin{align} \xi &=\frac{Z}{T^{\delta }}. \\[6pt] \nonumber \end{align}

Please note that in § 4 all variables have retained their previous symbols but now are functions of $Z,T$ . Moreover, $\xi$ denotes the similarity variable and the exponent $\delta$ controls the time dependence of the extent of the similarity region. When the filament radius becomes extremely small, both $u_{z}(Z,T)$ and $\tau _{\textit{zz}}(Z,T)$ in the vicinity of the singular point grow increasingly fast. Under these conditions, the nonlinear terms in the EVP constitutive equation become dominant. From now on, we will refer to this regime as the highly nonlinear regime. Next, we substitute the self-similar forms of $h(Z,T)$ and $u_{z}(Z,T)$ , in the continuity equation, (2.12), and enforce time invariance to obtain

(4.7) \begin{equation}\delta -\alpha _{2}=1,\quad \phi '=\phi \frac{\alpha _{1}-\dfrac{\psi '}{2}}{\psi +\delta \xi }.\end{equation}

Given that the axial momentum equation is the dominant momentum balance and $\tau _{\textit{zz}}$ the dominant stress component, we analyse the $z$ -component of the constitutive equation, which according to the SHB model is

(4.8) \begin{equation}\frac{\partial \tau _{\textit{zz}}}{\partial t}+u_{z}\frac{\partial \tau _{\textit{zz}}}{\partial z}=2\frac{\partial u_{z}}{\partial z}\tau _{\textit{zz}}+\frac{2}{Ec}\frac{\partial u_{z}}{\partial z}-\frac{\max \!\left(0,\, \left| \boldsymbol{\tau} _{\boldsymbol{d}}\right| -Y_{s}\right)^{\frac{1}{n}}}{Ec}\frac{{\tau} _{\textit{zz}}}{\left| \boldsymbol{\tau} _{\boldsymbol{d}}\right| }.\end{equation}

According to its definition, $| \boldsymbol{\tau}_{\boldsymbol{d}}|$ is

(4.9) \begin{equation}\left| \boldsymbol{\tau} _{\boldsymbol{d}}\right| =\sqrt{\frac{1}{2}\left(\left(\tau _{\textit{rr}}-\frac{tr\left(\boldsymbol{\tau} \right)}{3}\right)^{2}+2\tau _{rz}^{2}+\left(\tau _{\textit{zz}}-\frac{tr\left(\boldsymbol{\tau} \right)}{3}\right)^{2}+\left(\tau _{\theta \theta }-\frac{tr\left(\boldsymbol{\tau} \right)}{3}\right)^{2}\right)}.\end{equation}

In regions of high nonlinearity $tr(\boldsymbol{\tau })\approx \tau _{\textit{zz}}$ , yielding

(4.10) \begin{equation}\left| \boldsymbol{\tau} _{\boldsymbol{d}}\right| \approx \sqrt{\left(\frac{1}{2}\right)\left[\left(-\frac{\tau _{\textit{zz}}}{3}\right)^{2}+\left(\frac{2\tau _{\textit{zz}}}{3}\right)^{2}+\left(-\frac{\tau _{\textit{zz}}}{3}\right)^{2}\right]}=\sqrt{\frac{1}{3}\tau _{\textit{zz}}^{2}}=\frac{1}{\sqrt{3}}\left| \tau _{\textit{zz}}\right|.\end{equation}

Moreover, in the same region $| \tau _{\textit{zz}}| \gg Y_{s}$ , while in the upper half of the thread that we examine $\tau _{\textit{zz}}\geq 0$ , simplifying the max term in the EVP constitutive equation to $({1}/{\sqrt{3}}\tau _{\textit{zz}})^{{1}/{n}}$ . Therefore, the YS does not enter the dominant balance in the constitutive law and the entire thread is yielded in all examined cases. By assuming that the normal $z$ -component of the stress tensor in the EVP thread is constant, we neglect the convective part of the substantial derivative in (4.8). The rest of the terms in the EVP constitutive equation are approximated following ideas in Renardy (Reference Renardy1995) and Fontelos & Li (Reference Fontelos and Li2004), resulting in

(4.11) \begin{equation}\frac{\partial \tau _{\textit{zz}}}{\partial t}=2\frac{\partial u_{z}}{\partial z}\tau _{\textit{zz}}-\frac{\left(\frac{1}{\sqrt{3}}\tau _{\textit{zz}}\right)^{\frac{1}{n}}}{Ec}\sqrt{3}.\end{equation}

Substituting the assumed forms for the similarity variables and enforcing time invariance yields

(4.12) \begin{equation}\alpha _{3}=\frac{n}{n-1},\,\, -\alpha _{3}{\Omega} +\delta \xi {\Omega}^{\prime} =2\phi^{\prime}{\Omega} -\frac{1}{Ec\left(\sqrt{3}\right)^{n-1}}{\Omega} ^{\frac{1}{n}}.\end{equation}

Considering inertialess physics, the slender momentum equation, (2.13), becomes

(4.13) \begin{equation}\frac{\partial }{\partial z}\big(h+h^{2}(\tau _{\textit{zz}}-\tau _{\textit{rr}})\big)=0.\end{equation}

For this balance to be consistent, $\tau _{\textit{zz}}$ must grow like 1/h so we can conclude that

(4.14) \begin{equation}\alpha _{1}=-\alpha _{3}=\frac{n}{1-n}.\end{equation}

Turning our attention again to the mass conservation equation, we obtain

(4.15) \begin{equation} \phi^{\prime}=\phi \frac{\alpha _{1}-\dfrac{\psi^{\prime}}{2}}{\psi +\delta \xi }.\end{equation}

The denominator will be zero where $\psi =-\delta \xi$ , for $\xi =\xi _{0}$ . Therefore, as in the Newtonian (Doshi et al. Reference Doshi, Suryo, Yildirim, McKinley and Basaran2003; Eggers Reference Eggers1993) or the power-law material (Doshi et al. Reference Doshi, Suryo, Yildirim, McKinley and Basaran2003), a smooth solution exists only if, at $\xi =\xi _{0}$ , the following hold:

(4.16) \begin{align} \psi \left(\xi _{0}\right)&=-\delta \xi _{0}, \end{align}
(4.17) \begin{align} \psi^{\prime} \left(\xi \right)&=2\alpha _{1}. \end{align}

Hence, a Taylor series expansion in $\xi$ about $\xi =\xi _{0}$ as in Doshi et al. (Reference Doshi, Suryo, Yildirim, McKinley and Basaran2003) results in

(4.18) \begin{align} \phi \left(\xi -\xi _{0}\right)&=\phi _{0}+\phi _{2}\left(\xi -\xi _{0}\right)^{2}+\phi _{4}\left(\xi -\xi _{0}\right)^{4}+\ldots, \end{align}
(4.19) \begin{align} \psi \left(\xi -\xi _{0}\right)&=-\delta \xi _{0}+2\alpha _{1}\left(\xi -\xi _{0}\right)+\psi _{3}\left(\xi -\xi _{0}\right)^{3}+\ldots, \end{align}

where $\phi _{0},\, \phi _{2},\, \ldots $ and $\psi _{3},\, \ldots$ are the coefficients of the expansion. In all our simulations, the value of the axial velocity is zero at the plane of symmetry, which arises at $\xi _{0}=0$ for inertialess flow physics. Therefore, $\phi$ and $\psi$ are even and odd functions, respectively. The minimum value of $\phi$ is $\phi _{0}$ . Hence, the minimum radius takes the following power-law form:

(4.20) \begin{equation}h_{\textit{min}}\left(t\right)=\phi _{0}\, T^{{\alpha _{1}}},\end{equation}

where $\phi _{0}$ depends on physical properties and can be determined numerically. In this study, because of the dual nature of the material, and particularly the max term in the constitutive model, the flow is more complicated, and we do not determine analytically the value of $\phi _{0}$ . This could be pursued in future studies. The main focus of this study is the breakup asymptotic forms and the existence of self-similarity. In a nutshell, the results for the scalings in the capillary breakup regime are given by (4.7a ) and (4.14).

The asymptotes derived above are similar with those derived by Renardy, (Reference Renardy2002a ) for a generalised form of the viscoelastic Phan–Thien–Tanner (G-PTT) model. The G-PTT fluid described in Renardy, (Reference Renardy2002a ) in dimensional form can be written as follows:

(4.21) \begin{equation}\tilde{\lambda }\overset{\boldsymbol{\nabla }}{\tilde{\boldsymbol{\tau }}}+\tilde{\boldsymbol{\tau }}+\frac{\nu \tilde{\lambda }^{a-1}}{\tilde{\eta }_{p}^{a-1}}\left(tr\left(\tilde{\boldsymbol{\tau }}\right)\right)^{a-1}\tilde{\boldsymbol{\tau }}=\tilde{\eta }_{p}\widetilde{\dot{\boldsymbol{\gamma }}},\end{equation}

where $a\gt 1$ is the exponent of the G-PTT model, $\tilde{\lambda }$ is the relaxation time, $\tilde{\eta }_{p}$ is the polymeric viscosity and $\nu$ is a positive dimensionless number, playing the same role as the mobility factor in the Giesekus model. The $z$ -component of G-PTT model where $\tau _{\textit{zz}}$ is the dominant stress component is

(4.22) \begin{equation}\frac{\partial \tilde{\tau }_{\textit{zz}}}{\partial \tilde{t}}=2\frac{\partial \widetilde{u_{z}}}{\partial \tilde{z}}\tilde{\tau }_{\textit{zz}}-\frac{\nu \tilde{\lambda }^{a-2}}{\tilde{\eta }_{p}^{a-1}}\tilde{\tau }_{\textit{zz}}^{a}.\end{equation}

Equation (4.22) reduces to (4.11) if the exponent $a$ in (4.22) is replaced by $({1}/{n})$ . Hence, it is expected that the G-PTT model will have the same asymptotes as the SHB model in the breakup regime. According to Renardy, (Reference Renardy2002a ), the asymptotes of the G-PTT model are

(4.23) \begin{align} \alpha _{1}&=\frac{1}{a-1}, \quad \alpha _{2}=-1, \quad \alpha _{3}=\frac{1}{1-a}, \!\!\qquad\text{if }1\lt a\lt \frac{7}{3}, \end{align}
(4.24) \begin{align} \alpha _{1}&=\frac{1}{4a-8}, \quad \alpha _{2}=-1, \quad \alpha _{3}=\frac{1}{1-a}, \quad \text{ if }\, \frac{7}{3}\lt a\lt 3. \end{align}

For $({7}/{3})\lt a\lt 3\, (0.33\lt n\lt 0.43),$ surface tension has no influence on breakup, and pinching enters a new regime, the so-called elastic breakup regime. Our numerical results are consistent with this finding, as we will demonstrate in § 5.5.2. Integration of (4.13) yields

(4.25) \begin{equation}h+h^{2}\tau _{\textit{zz}}=F\left(t\right),\end{equation}

where $F(t)$ is the total force acting on the filament cross-section. In the elastic breakup regime, the first term on the left-hand side is associated with capillarity and is subdominant. Substituting $\alpha _{1}$ and $\alpha _{3}$ , as given in (4.24), into the asymptotic forms of (4.3) and (4.5), one obtains

(4.26) \begin{equation}F\left(t\right)\sim h^{2}\tau _{\textit{zz}}\sim \, \left(t_{0}-t\right)^{{2\alpha _{1}}+{\alpha _{3}}}\sim \left(t_{0}-t\right)^{\frac{2n-6n^{2}}{\left(n-1\right)\left(4-8n\right)}}.\end{equation}

For the force to be finite, requires that $({2n-6n^{2}})/{(n-1)(4-8n)}\gt 0$ which yields $n\gt {1}/{3}$ . Indeed, our code did not converge for values $n\lt 0.33$ . Furthermore, Renardy suggests that $\delta =0,\, \alpha _{2}=-1$ , but our 2-D numerical results slightly deviate from this value. The reason for this difference is probably caused by the fact that Renardy has neglected any radial dependence of $u_{z}, p,\, \tau _{\textit{zz}},\, \tau _{rr\, }\, {\textrm{and}}\, \tau _{\theta \theta }$ and completely neglected $u_{r}, \tau _{rz}$ , which is not entirely correct, as will be seen in § 5. On the other hand, our solution of the 2-D equations may include numerical error. We calculate the following scaling exponents for axial velocity and axial position numerically:

(4.27) \begin{equation}\alpha _{2}=\frac{n}{n-1}\left(\lt 0\right),\, \delta =\frac{n}{n-1}+1(\lt 0), \quad \textrm{if}\quad n\gt 0.5,\end{equation}

where and

(4.28) \begin{equation}\alpha _{2}\rightarrow -1, \delta \rightarrow 0,\quad \textrm{if}\quad n\leq 0.5\, \quad {\textrm{and}} \quad n\rightarrow 0.4.\end{equation}

In late stages of breakup, since $\tau _{\textit{zz}}$ is dominant over $\tau _{\textit{rr}}$ , the asymptotic behaviour of the transient extensional viscosity $\eta _{ext}=(\tau _{\textit{zz}}-\tau _{\textit{rr}})/\dot{\epsilon }_{\textit{local}}$ yields

(4.29) \begin{equation}\eta _{ext}\sim \left(t_{0}-t\right)^{\frac{2n-1}{n-1}}.\end{equation}

Clearly, for $n\gt 0.5$ , $\eta _{ext}$ diverges to infinity, while for $n\lt 0.5$ , $\eta _{ext}$ tends to 0, indicating transient extension-rate-thickening and -thinning behaviours, respectively. It is noteworthy that the exponents in EVP fluids are different from those in power-law or viscoplastic fluids. Similar deviations have been observed in experiments with shear-thinning viscoelastic materials where scalings do not follow expected asymptotes of a power-law material (Moon et al. Reference Moon, Yamani, McKinley and Lee2024). In all our simulations, the scaling exponent for the axial position, denoted as δ, is found to be negative. We are probing the axial position at $z$ where $h=1.01h_{\textit{min}}$ . The negative sign of δ implies that the neck length increases as breakup is approached, instead of decreasing, as is usually the case with $\delta \gt 0$ . Moreover, the magnitude of the scaling exponent of the minimum radius, |α1|, consistently exceeds |δ| in our simulations. Hence, the thinning of the thread occurs faster than the growth of the length of the neck. Consequently, although breakup still occurs at $z=0$ , the axial extent of fast thinning increases. This has been called ‘breakup at a finite axial length’ by Renardy, (Reference Renardy2002a ) who found that $\delta =0$ , analytically for the G-PTT fluid. This implies that the neck length remains constant. Consequently, the inertialess breakup of EVP fluids occurs over a finite length, whereas in power-law or viscoplastic fluids, it takes place at a single point. These ideas will be discussed and supported further in § 5.5.4.

4.2. Ideal uniaxial extension of an EVP material under highly nonlinearity conditions

To further validate the derived scaling exponents for $h$ and $\tau _{\textit{zz}}$ , we solve numerically an ideal uniaxial extension flow with the EVP constitutive equation. This simplification reduces the slender jet equations to those for a perfectly cylindrical liquid jet under extension. Thus, the strain rate becomes

(4.30) \begin{equation}\dot{\epsilon }=\frac{\partial u_{z}}{\partial z}=-\frac{2}{h}\frac{{\textrm{d}}h}{{\textrm{d}}t}.\end{equation}

Assuming the power-law asymptote for the radius, the local strain rate reduces to

(4.31) \begin{equation}h=\phi _{0}\left(t_{0}-t\right)^{{\alpha _{1}}}\, \Longrightarrow \, \dot{\epsilon }=\frac{2\alpha _{1}}{t_{0}-t}\approx \frac{2}{T}.\end{equation}

We note that assuming $({2\alpha _{1}})/({t_{0}-t})\approx {2}/{T}$ will not affect the high nonlinearity trend. In uniaxial extension

(4.32) \begin{equation}u_{r}=-0.5\dot{\epsilon }r, \quad u_{\theta }=0, \quad u_{z}=\dot{\epsilon }z.\end{equation}

In this ideal flow, the velocity gradient $\boldsymbol{\nabla }\boldsymbol{u}$ and $\boldsymbol{\dot{\gamma}}$ are

(4.33) \begin{equation}\boldsymbol{\nabla }\boldsymbol{u}=\dot{\epsilon }\, \textit{diag}\left(-\frac{1}{2},-\frac{1}{2}\, ,1\right), \quad \dot{\boldsymbol{\gamma} }=\boldsymbol{\nabla }\boldsymbol{u}+\left(\boldsymbol{\nabla }\boldsymbol{u}\right)^{T}=\dot{\epsilon }\, \textit{diag}\left(-1,-1\, ,2\right).\end{equation}

The problem is reduced to solving the two components of the momentum balance in the $r\hbox{-}$ and $z\hbox{-}$ directions, noting that there is no spatial dependence under this idealised flow condition

(4.34) \begin{align} \frac{{\textrm{d}}\tau _{\textit{zz}}}{{\textrm{d}}t}&=2\dot{\epsilon }\tau _{\textit{zz}}+\frac{2}{Ec}\dot{\epsilon }-\frac{\max \!\left(0,\, \left| \boldsymbol{\tau} _{\boldsymbol{d}}\right| -Y_{s}\right)^{\frac{1}{n}}}{Ec}\frac{\tau _{\textit{zz}}}{\left| \boldsymbol{\tau} _{\boldsymbol{d}}\right| }, \end{align}
(4.35) \begin{align} \frac{{\textrm{d}}\binom{\tau _{\textit{rr}}}{\tau _{\theta \theta }}}{{\textrm{d}}t} &=-\dot{\epsilon }\binom{\tau _{\textit{rr}}}{\tau _{\theta \theta }}-\frac{1}{Ec}\dot{\epsilon }-\frac{\max \! \left(0,\left| \boldsymbol{\tau} _{\boldsymbol{d}}\right| -Y_{s}\right)^{\frac{1}{n}}}{Ec}\frac{\binom{\tau _{\textit{rr}}}{\tau _{\theta \theta }}}{\left| \boldsymbol{\tau} _{\boldsymbol{d}}\right| }. \end{align}

The system of the three ordinary differential equations is solved numerically. The time evolution of $\tau _{\textit{zz}}$ is shown on figure 3, for $n=0.45$ . The important finding from this simple analysis is that $\tau _{\textit{zz}}$ scales like $(t_{0}-t)^{n/n-1}$ , irrespective of other parameters, like $Ec,\, \text{ or }Y_{S}$ in extreme nonlinearity, in other words it depends only on $n$ .

Figure 3. Evolution of the axial normal stress, $\tau _{\textit{zz}}$ , vs. time to pinch in uniaxial extension flow with $\dot{\epsilon }={2}/{t_{0}-t}$ . The other EVP parameters are $n=0.45,\, Y_{S}=1$ , resulting in ${n}/{n-1}=-0.8181$ .

The same figure demonstrates that elasticity through $Ec$ affects only the prefactor of the asymptotic form, or, in other words, the intersection points of the asymptotic lines with the ordinate. Reducing $Ec$ will require longer time to reach the region of high nonlinearity, where $\tau _{\textit{zz}}$ scales as we calculated, but at the same point in time the axial stress is smaller.

Moreover, multiple tests were conducted by varying the exponent $n$ with the same extension rate $\dot{\epsilon }=({2}/{T})$ , and with $Ec=0.1$ and $Y_{s}=1$ to determine if the exponent in the scale $(t_{0}-t)^{{\alpha _{3}}}$ will agree with our prediction of $({n}/{n-1})$ . Figure 4, clearly shows that this is the case for $n\in [0.3,\, 0.8].$

Figure 4. The scaling exponent of axial normal stress, $\tau _{\textit{zz}}$ , vs. the SHB strain-rate-thinning exponent (n). The other EVP parameters are $Ec=0.1,\, Y_{S}=1$ .

5. Solution of the two-dimensional equations

We begin this section by validating the numerical algorithm we developed through a test case. Following this, we present the computed solution for the base material, including detailed examination of stress and velocity fields near the pinch point. We analyse the presence of universality and self-similar profiles. We conclude it by examining the impact of material parameters on the asymptotic behaviour.

Figure 5. Pinching of a power-law thread with $n=0.4,\, Oh^{-2}=0,\, L=2,\, U=0$ ; (a) interface profile of the entire filament just before the breakup, (b) close-up near the pinch-off point, (c) minimum radius, axial position and velocity plotted as a function of time to pinch. The calculated slopes are in perfect agreement with the ones reported in Suryo & Basaran (Reference Suryo and Basaran2006).

5.1. Accuracy of the numerical method

To validate our numerical method, we compare our simulation results with those reported by Suryo & Basaran (Reference Suryo and Basaran2006) for a power-law material. In such a material, the viscosity is expressed as

(5.1) \begin{equation}\tilde{\mu }\left(\widetilde{\dot{\gamma }}\right)=\tilde{k}\left| \widetilde{\dot{\gamma }}\right| ^{n-1},\end{equation}

where $\tilde{\mu }$ is the apparent power-law viscosity, $\tilde{k}$ is the consistency index and $n$ is the power-law exponent. For flow without inertia, the aforementioned study predicts different flow regimes, depending on the value of $n$ . Among them the most challenging one to capture is the so-called non-slender viscous power-law regime arising for $n\lt 0.54$ , because the filament loses its slenderness and as $n$ decreases, the strongly shear-thinning behaviour of the material leads to very fast pinch-off of the thread under capillary action. In this regime, the asymptotic scales are

(5.2) \begin{align} h\, &\sim \, T^{n}, \end{align}
(5.3) \begin{align} Z\, &\sim \, T^{n}, \end{align}
(5.4) \begin{align} u_{z}\, &\sim \, T^{n-1}. \end{align}

We select $n=0.4,\, Oh^{-2}=0,\, L=2$ to validate our numerical framework. The initial interface height is set as $h(z,0)=1-0.8\cos ({\pi z}/{2})$ . This perturbed interface suffices to initiate instability, even with a stationary upper disk ( $u_{\textit{plate}}=0$ ), because of the absence of plasticity. Figure 5(a) shows the interface profile just before the breakup obtained with our numerical method and figure 5(b) illustrates a close-up view of the interface near the pinch point. Furthermore, figure 5(c) depicts the scaling of the minimum radius, axial velocity and axial position for the power-law thread. All scales in this figure are in perfect agreement with those reported in Suryo & Basaran (Reference Suryo and Basaran2006). This demonstrates the validity and accuracy of our calculations.

5.2. General time Evolution of an EVP filament

We use the 0.2 % Carbopol solution produced and rheologically characterised by Lopez et al. (Reference Lopez, Naccache, De and Mendes2017), as the material for our base case study. We used a nonlinear regression to determine the material parameters from their flow curve and frequency sweep test at 1 % strain amplitude, and we achieved close agreement with the rheological measurements. For a more detailed description, the reader is referred to Moschopoulos et al. (Reference Moschopoulos, Spyridakis, Varchanis, Dimakopoulos and Tsamopoulos2021, Reference Moschopoulos, Kouni, Psaraki, Dimakopoulos and Tsamopoulos2023). The resulting parameter values for the base case study, are given in table 2, except for setting $Oh^{-2}=0$ , to simulate flow physics without inertia.

Table 2. Parameters for the base case study.

Figure 6 illustrates the time evolution of the EVP filament. Once the disks begin to move, the material yields in most of its length, as seen for example at $t=1.32$ . Near the disks though, the material only deforms elastically as a Kelvin–Voigt solid, because its initial velocity and stresses were set to zero everywhere and the no-slip and no-penetration conditions adhere the material to the disks, preventing the development of normal stresses as well. To identify the yielded regions, we employ the von Mises criterion, as detailed in § 2.1. Yielding occurs first at the plane of symmetry, where the stress magnitude is maximised and with the assistance of capillarity they form a neck there. At all times, the filament deformation retains the plane of symmetry at $z=0$ , because of the absence of gravity and inertia. As stretching continues, the neck gets thinner there making the velocity and stress gradients more pronounced. Simultaneously, rate thinning decreases the viscous resistance there, making the deformation larger and more local, $\, t=2.75$ . Consequently, deformation has to decrease away from the mid-plane, reducing the velocity and stress gradients. Where the stress magnitude drops below the yield condition again, the material will solidify, permitting deformation in a decreasing length of the bridge. This sequence of events can be seen in the first three panels of figure 6.

Figure 6. Time evolution of filament under constant stretching with $U$ = 0.39. The red and blue areas indicate yielded and unyielded regions, respectively. The other parameters are $Oh^{-2}=0,\, n=0.45,\, Y_{s}=1.25,\, Ec=0.14,\, L_{0}=2$ .

When $h_{\textit{min}}$ reaches approximately $O(10^{-2})$ , the filament begins to deform even more rapidly, now driven primarily by capillarity. The faster escape of the material from the plane of symmetry through the increasingly smaller cross-section of the neck produces a jet-like, larger axial velocity away from the pinching region. With its increased axial velocity, this material now penetrates further inside the unyielded material away from the necking region, and toward the top moving solid disk, compressing it more effectively. In this area, the local increase in the radius is attributed to the continuous injection of material. This complex kinematics leads to material yielding away from the neck with a small unyielded area in between at time $t=2.8978$ . Before turning into the dynamics, we note the following: Because of axial symmetry, the radial velocity is zero at the axis of symmetry, but away from it and near the pinch point, at $z=0$ , it is negative following the decrease in radius, whereas in the corner region (close to highest curvature of the interface), it becomes positive. Consequently, close to the pinch point, $\tau _{\textit{rr}}$ is negative but it turns positive in the vicinity of the corner region, explaining the sign change of $\tau _{\textit{rr}}$ in this area; see figure 7 below. The same reasoning applies to the change of the sign of $\tau _{\textit{zz}}$ and $\tau _{rz}$ . The sign change of all stresses indicates that, close to the corner region, the components of the deviatoric stress tensor will approach zero. Hence, the transition region contains unyielded material that separates the yielded neck from the yielded island further up. As the filament approaches breakup, velocity and stress magnitude increase even faster and the unyielded area in the transition region shrinks in size and eventually disappears, as shown in figure 6, at $t=2.8984$ . It is worth mentioning that this is observable in EVP materials, because they can deform even in unyielded regions.

Figure 7. Contours of velocity and stress components of the filament under constant stretching with $U$ = 0.39 just before pinch-off, at $t=2.8984$ . Other parameters are $Oh^{-2}=0, n=0.45, Y_{s}=1.25, Ec=0.14, L_{0}=2$ .

Figures 7(a) and 8(a) present contours of velocity and extra stress tensor components, along with the second invariant of the extra stress tensor, just before pinch-off at $t=2.8984$ . These plots clearly illustrate that extremely high velocity and stress magnitudes arise in the necking region. Moving away from this region, all values rapidly decrease until they match the boundary conditions imposed on the solid plate. Figures 7(b) and 7(c) depict contours of velocity components focusing at the corner and necking regions of the interface, respectively. In the necking region, the axial velocity $u_{z}$ dominates, becoming three orders of magnitude larger than the radial velocity $u_{r}$ . This dominance decreases when approaching the corner region, with its magnitude being only one order of magnitude higher than that of $u_{r}$ .

Figure 8. Contours of pressure and stress components of the filament under constant stretching with $U$ = 0.39 just before pinch-off, at $t=2.8984$ . Other parameters are $Oh^{-2}=0, n=0.45, Y_{s}=1.25, Ec=0.14, L_{0}=2$ .

As demonstrated in figures 7(c) and 7(d), the magnitudes of $\tau _{\textit{rr}}$ and $\tau _{\theta \theta }$ are nearly identical in both magnified regions, with their extrema occurring at the necking region. This is consistent with the fact that $({{\textrm{d}}u_{r}}/{{\textrm{d}}r})$ and $({u_{r}}/{r})$ also attain their maximum value in this region. The pressure magnitude in the necking region is proportional to $({1}/{h})$ , implying that the high pressure within this region decreases progressively as the distance from the symmetry plane increases. This trend is clearly illustrated in figures 8(b) and 8(c), which show pressure contours in the magnified corner and necking regions, respectively. The shear stress component $\tau _{rz}$ reaches its extremum in the corner region, approximately at the point of highest interface curvature. Figure 8(b) further reveals that $\tau _{rz}$ exhibits a magnitude comparable or even slightly higher than $\tau _{\textit{rr}}$ and $\tau _{\theta \theta }$ in this region. The extension-dominated nature of filament breakup is evident in the contours of $\tau _{\textit{zz}}$ in necking and corner regions, as shown in figure 8(c) and 8(d). Similar to the axial velocity, $\tau _{\textit{zz}}$ is three orders of magnitude higher than all other stress components in the necking region but rapidly decreases until it becomes comparable to $\tau _{rz}$ at the corner. The magnitude of the second invariant of the extra stress tensor in the thread is approximately $({1}/{\sqrt{3}})| \tau _{\textit{zz}}|$ , see figure 8(c); However, this relation no longer holds beyond the neck region, as depicted in figure 8(d). The variation in the radial direction inside the neck for all variables is negligible. One may observe a small radial variation of $u_{r}, \tau _{rz }\text{ and}\, \tau _{\textit{rr}}$ , but only because their magnitude is much smaller than that of the other variables and the range of colours in each panel depends on the range of variation of the corresponding variable. In the supplementary material SM2, we present the time evolution of the axial stress component $\tau _{\textit{zz}}$ and the axial velocity $u_{z}$ , where one can easily observe that they increase from $O(1)$ to $O(10^{5})$ without any oscillations in their contours, during the same simulation.

All these justify the assumptions made to derive the slender filament equations in the necking region. However, we will not employ the slender filament equations for simulating this flow. As stated by Moschopoulos et al. (Reference Moschopoulos, Syrakos, Dimakopoulos and Tsamopoulos2020), in stretching of viscoplastic filament, the deviation between 2-D and slender filament simulations increases as the viscoplastic character of the flow becomes more pronounced. We believe this applies for EVP material as well and verifying the accuracy of the solution obtained by the slender equations requires extensive analysis, particularly for different combinations of $Y_{s}$ and $Ec$ . Moreover, aside from the necking region, the filament is not slender throughout, particularly in the corner region, necessitating a full 2-D analysis for accurate results. Additionally, the slender filament approximation, neglects $\tau _{rz}$ , which is inappropriate given that $\tau _{rz}$ is of the same order of magnitude or larger than $\tau _{\textit{rr}}$ and $\tau _{\theta \theta }$ in this region; see also Eggers, Herrada & Snoeijer (Reference Eggers, Herrada and Snoeijer2020).

5.3. Determination of pinch-off time, ${t}_{{0}}$

Accurate approximation of the pinch-off time, $t_{0}$ , is most important, because even a very small error in determining its value will lead to incorrect calculation of the asymptotes. Moreover, in our arbitrary Lagrangian Eulerian formulation, breakup cannot occur without external intervention, which must take place as close as numerically possible and physically allowed to the pinch-off time. Computational experiments as well as other studies suggest that, in order to accurately determine it, simulations have to be conducted until the minimum radius becomes approximately five orders of magnitude smaller than the initial radius, or, in the present dimensionless form, when $h_{\textit{min}}=10^{-5}$ . Then the very last time instant of the simulation can be considered as pinch-off time and the scalings are correctly determined. On the other hand, when the minimum radius reaches approximately $O(10^{-3})$ , the EVP stresses grow very fast because of the viscoelastic nature of the EVP material, making the resulting system of equations extremely stiff to solve. Although reaching values as small as $h_{\textit{min}}=10^{-5}$ may not be feasible, we have employed a least-squares, nonlinear regression method to extrapolate the pinch-off time.

The extrapolation to the breakup time is based on the asymptotic form obtained in § 4.1 for $h_{\textit{min}}=\phi _{0}(t_{0}-t)^{{\alpha _{1}}}$ . Treating $\phi _{0}, t_{0}$ and $\alpha _{1}$ as fitting parameters, we apply this power-law function to the minimum radius data over the last 100 computed time steps. Figure 9 depicts the regression applied to the minimum radius as a function of time for the base case study. The computed $t_{0}$ for this case is found to be $2.8984063$ , while $\varPhi _{o }\approx 10.7$ and $\alpha _{1}=0.8$

Figure 9. Nonlinear regression employed to extract the pinch-off time, $t_{0}$ , from the numerical results of the base case study, for $Oh^{-2}=0, n=0.45, Y_{s}=1.25, Ec=0.14, L_{0}=2$ .

The calculated $t_{0}$ is then used to generate plots of $h_{\textit{min}}, z, u_{z}, \tau _{\textit{zz}}$ versus $T$ . According to § 4.1, the axial stress $\tau _{\textit{zz}}$ should scale as $(t_{0}-t)^{({n}/{n-1})}$ . Owing to conservation of momentum and the balance between capillarity and axial stresses, the minimum radius should decrease asymptotically as $({1}/{\tau _{\textit{zz}}})$ , in other words as $h_{min }\sim (t_{0}-t)^{({n}/{1-n})}.$ Furthermore, to satisfy the slender jet continuity equation, the axial position and velocity must scale like $(t_{p}-t)^{\delta }$ and $(t_{p}-t)^{1-\delta }$ , respectively. Figure 10 illustrates in log–log scales $h_{\textit{min}}, z, u_{z}, {\textrm{and}}\, \tau _{\textit{zz}}$ versus $t_{0}-t$ for $n=0.45, Ec=0.14, Y_{s}=1.25, U=0.39$ . The linear slope in figure 10 verifies that, indeed, a power-law relation exists for all variables. The calculated slopes are $h_{\textit{min}}, Z, u_{z}\, {\textrm{and}}\, \tau _{\textit{zz}}$ are $0.80\, (\approx ({n}/{1-n})),-0.08\, (=\delta ), -1.08\, (=\delta -1)$ and $-0.81\, (\approx {n}/{n-1})$ , respectively. These slopes represent the exponents of power-law scaling functions for the base material and they are in perfect agreement with the asymptotes we derived with the elasto-capillary scalings, as discussed earlier.

Figure 10. Decay of the minimum radius, growth of axial velocity, axial position and axial stresses as pinching time is approached for $Oh^{-2}=0, n=0.45, Ec=0.14, Y_{s}=1.25, U=0.39$ ( $({n}/{1-n})=0.818$ . The slopes determine the corresponding exponents of the power-law scaling functions.

5.4. Self-similar forms of the two-dimensional solutions

Given a distinct set of parameters $(n,Ec,Y_{s})$ , first we calculate the scaling exponents for minimum radius, axial position, axial velocity and stress, as shown in figure 10, then we rescale these variables accordingly at different times to verify their collapse onto a single curve, completing in this way the demonstration of the existence of self-similarity. We have observed that this holds for a range of power-law exponents, but, for conciseness, next we present only a single set of parameters corresponding to the base case study. The inset to figure 11 shows interface profiles for five different values of minimum radius $h_{\textit{min}}$ . Figure 11 depicts that these transient interface profiles obtained by 2-D simulations collapsing onto a single self-similar interface profile by rescaling the interface height $\overline{h}=({h}/{(t_{0}-t)^{({n}/{1-n})}})$ and the axial position $\overline{Z}=({z-z_{0}}/{(t_{0}-t)^{\delta }})$ , where $\delta \approx -0.08$ . Please note that the rescaled variables are denoted by an overbar ( $\overline{})$ over the corresponding variable symbol.

Figure 11. (Inset) Transient interface shape obtained by 2-D simulations at five different $h_{\textit{min}}$ values. (Main figure) Rescaled interface using the numerically determined scaling laws for the same five $h_{\textit{min}}$ values. The main figure illustrates convergence to a self-similar interface profile. The very large difference in the magnitudes of $h$ before and after scaling is noteworthy.

The inset to figure 12 shows the variation of axial velocity $u_{z}$ at the interface versus the axial coordinate $z$ for five different values of minimum radius $h_{\textit{min}}$ . Figure 12 depicts that by rescaling the axial velocity as $\overline{u}_{z}= ({u_{z}}/{(t_{0}-t)^{{\alpha _{2}}}})$ , with $\alpha _{2}=-1.08$ , the transient velocity profiles collapse onto a single self-similar profile as the minimum radius $h_{\textit{min}}\rightarrow 0$ .

Figure 12. (Inset) Transient velocity variation along axial coordinate (z) at the interface obtained by 2-D simulations at five different $h_{\textit{min}}$ values. (Main figure) Rescaled velocity plotted against rescaled axial coordinate using the numerically determined scaling laws for the same five $h_{\textit{min}}$ values. The main figure demonstrates convergence to a self-similar velocity profile. The very large difference in the magnitudes of the velocity before and after scaling is noteworthy.

Finally, the inset to figure 13 shows the axial stress $\tau _{\textit{zz}}$ versus the axial coordinate extracted from nodal values of 2-D calculations at the interface for five different values of $h_{\textit{min}}$ . The rescaled axial stress $\overline{\tau }_{\textit{zz}}=({\tau _{\textit{zz}}}/{(t_{0}-t)^{({n}/{n-1})}})$ vs. rescaled axial coordinate does converge to a self-similar profile, as shown on figure 13.

Figure 13. (Inset) Transient axial stress variation along axial coordinate (z) at the interface obtained by 2-D simulations at five different $h_{\textit{min}}$ values. (Main figure) Rescaled axial stress plotted against rescaled axial coordinate using the numerically determined scaling laws for the same five $h_{\textit{min}}$ values. The main figure demonstrates convergence to a self-similar axial stress profile. The very large difference in the magnitudes of stress before and after scaling is noteworthy.

The plots in this section show good collapse near $\overline{Z}=0$ ; however, the quality of the collapse diminishes as $\overline{Z}$ increases. This can be attributed to the negative value of the exponent $\delta$ , which leads to smaller values of $\overline{Z}$ compared with cases where $\delta \gt 0$ . As illustrated in the scaling plot of figure 10, the axial position increases as breakup is approached. Consequently, in self-similarity plots, the span of $\overline{Z}$ is reduced relative to other constitutive equations, such as Newtonian or power-law fluids, for which $\delta \gt 0$ .

Moreover, we have extracted numerically scaling laws for the variables in the corner region. However, they do not follow a universal scaling law, we were unable to derive analytically these scalings nor could we obtain self-similar solutions. It is possible that this is caused by the fact that the corner region involves in the thread side yielded material until the very late stages of breakup, much later than when asymptotic behaviour of the neck starts, but unyielded material in the other side toward the disc. Consequently, derivation of the asymptotics is not straightforward, because the material exhibits dual behaviour in this region. Additionally, since the solvent viscosity is set to zero in this study, the dominant balance and self-similar form of the governing equations become more complex than those reported by Fontelos & Li (Reference Fontelos and Li2004) for the Giesekus fluid.

5.5. Parametric analysis

We will separate this section based on the value of the power-law index. When $0.43\lt n\lt 1$ , surface tension is the main driving force causing breakup. This range of values is examined in § 5.5.1. Although for $n\gt 0.5$ , the SHB model predicts extension rate hardening, which is not expected for Carbopol (Kordalis et al. Reference Kordalis, Varchanis, Ioannou, Dimakopoulos and Tsamopoulos2021), we will examine values of $n$ up to one for a more complete study. Values of $0.33\lt n\lt 0.43$ will be examined separately in § 5.5.2, because then capillarity plays no role in pinching-off, while $n=1$ will be examined in § 5.5.3, also separately, because, as we will demonstrate, pinching-off does not take place in the Saramito–Bingham model (Saramito Reference Saramito2007). We will examine the effect of $n$ and $Ec$ , where appropriate. Here, $Y_{s}$ has no effect, because the thread during pinching-off is yielded throughout.

5.5.1. Elasto-plasto-capillary breakup regime, ${0.43}\lt {n}\lt {1}$

Although the YS does not enter the dominant balance determining the asymptotic behaviour of the thread in this regime, the title includes the term ‘plasto’, because part of the included elastic term is what remains from the viscoplastic contribution in the constitutive law, see (3.11).

5.5.1.1. Effect of strain-rate-thinning exponent (n)

We proceed by examining the effect of the strain-rate-thinning exponent, $n$ , on the pinch-off dynamics. Figure 14(a) depicts interface profiles for four different values of $n$ , recorded when $h_{\textit{min}}=1.5\times 10^{-3}$ .

Figure 14. Profiles of the entire filament (a) and closeup views in the corner (b) and necking (c) regions for four different values of $n$ . In (b) we did not include the entire ‘local’ interface for $n=0.8$ , because it would reduce the clarity of the rest. The remaining parameter values are $Y_{s}=0.1, Ec=0.14, U=0.1$ .

As $n$ increases, strain-rate thinning decreases, resulting in increasing elastic stresses, including its radial normal component, which opposes capillarity. This leads to a longer thread around the middle of the filament for the same $h_{\textit{min}}$ , as depicted more clearly in figure 14(b). Given that the initial dimensions of the filament for all cases are the same, a longer neck causes a slight increase in the total height of the filament (see figure 14 a), when it has reached the same $h_{\textit{min}}$ under the same constant stretching velocity. Additionally, the increased opposing elastic force slows down the breakup rate, causing the EVP thread to break later in time as $n$ increases. Figure 14(c) shows that the interface profile close to pinch off is almost a perfect cylinder for all values of $n$ .

Figure 15. Effect of the exponent of strain-rate thinning on the scaling with respect to $(t_{o}-t)$ of minimum radius (a), axial stress (b), axial velocity (c) and axial length of neck (d). The rest of the parameter values are $Y_{s}=0.1, Ec=0.14, U=0.1$ .

The dependence on time to pinch of the minimum radius, the axial normal stress and axial velocity, as well as the length of the neck up to which its radius increases by only 1 % from its minimum value, are presented in figure 15. Figures 15(a) and 15(b) depict the $h_{min }$ decrease and the $\tau _{\textit{zz}}$ increase as a function of time to pinch $t_{o}-t$ . Focusing on their values when they reach their asymptotic behaviour, one readily observes that they follow very closely the power-law dependence on $n$ , namely $\pm n/(1-n)$ for all four values of $n$ . The calculated slopes, which represent the exponents of the power-law scaling functions, are almost in perfect agreement with the proposed asymptotes. An error of $\pm \, 2\,\%$ from the earlier predicted asymptotes may be attributed to inaccuracies of the numerical method and a slight radial dependence of the variables in the numerical solution of the 2-D equations. As these exponents require, when $n$ increases, the slopes of the asymptotes of both variables increase. Figure 15(a) also shows that increasing $n$ brings $h_{min }$ faster to a certain small value, although the time to breakup is delayed, since this occurs at larger values of $(t_{o}-t)$ . Similarly, figure 15(c), demonstrates that the slope of the asymptotic increase of $v_{z}$ increases with $n$ , although for $n\lt 0.5$ , the exponent does not have an analytic relation with it; see (3.27) and (3.28). Finally, figure 15(d) indicates that the thread of the filament becomes longer at a faster rate, when $n$ increases. This length extends from the plane of symmetry to the axial position where $h=1.01h_{min }$ , and it is denoted as $Z$ .

Altogether, figure 15 demonstrates that, for smaller $n$ values, the asymptotic behaviour is reached when calculations are carried out to times much closer to breakup. This leads to a thinner neck, driven by higher velocity produced by higher axial stress. Another important observation from these four panels is the sharp increase in the extreme values reached by $h_{\textit{min}}, v_{z}\, {\textrm{and}}\, \tau _{\textit{zz}}$ as $n$ decreases. These values are $\sim 2\times 10^{-4}, 2\times 10^{4}\, \text{and }5\times 10^{4}$ , respectively, for $n=0.45$ , while their initial values are 1, 0 and 0, respectively. This makes plain how challenging these calculations are and the importance of the new ideas adopted in the particular numerical algorithm.

5.5.1.2. Effect of elastic modulus

We move on by examining the effect of the elasto-capillary number, $Ec$ , on the pinch-off dynamics. Figure 16 illustrates the different shapes of the filament when $h_{\textit{min}}=4.5\times 10^{-4}$ .

Figure 16. Profiles of the entire filament (a) and closeup views in the corn (b) and necking (c) regions for three different values of $\, Ec$ . The rest of the parameter values are $Y_{s}=0.1, n=0.45, U=0.1$ .

Focusing on figure 16(a), we observe that increasing $Ec$ generates both a shorter filament and, hence, one that breaks faster, but a longer thread around the plane of symmetry. This seemingly contradictory trend is explained as follows: increasing $Ec$ affects both the unyielded and yielded regimes of the filament in different ways. To start with, when $Ec$ increases, the filament deforms more as an elastic solid prior to yielding, allowing for a more localised viscoelastic deformation in the region around the plane of symmetry. This leads to earlier formation of the neck and earlier breakup of the filament, although the opposing elastic forces also increase. All these lead to a shorter total filament height at pinch-off time, figure 16(a). This increased elastic yielding has been reported also by Moschopoulos et al. (Reference Moschopoulos, Kouni, Psaraki, Dimakopoulos and Tsamopoulos2023), where the bulk dynamics has been studied. The same increase in $Ec$ increases $\tau _{\textit{zz}}$ , which accelerates the evacuation of the neck and the decrease of its radius. At the same time, the larger normal stresses generate a longer neck, because they oppose neck thinning by capillarity, as shown on closeup view at the corner region, figure 16(b). Figure 16(c) shows that the neck profile close to pinch off forms a nearly perfect cylinder for all values of $Ec$ .

The trends in $h_{\textit{min}}$ decay and $\tau _{\textit{zz}}, v_{z},\text{ and}\, Z$ growth are presented in figure 17. The slopes of all four variables with respect to the time to pinch $(t_{o}-t)$ in the asymptotic regime depend only on $n$ , but are independent from $Ec$ . The elastocapillary number affects only the intersection with the ordinate, via affecting the coefficient $\phi _{0}$ in (3.20). Even this effect is weak for all variables, except for $Z$ . Indeed, figure 17(d) indicates that the axial extent over which this scaling holds is longer, or, in other words, the neck is longer for higher $Ec$ , although its growth rate remains the same. The calculated slopes, which represent exponents of the power-law scaling functions, are almost in perfect agreement with the proposed asymptotes.

Figure 17. Effect of $Ec$ on the scaling with respect to $(t_{o}-t)$ of minimum radius (a), axial stress (b), axial velocity(c) and axial length of neck (d). The rest of the parameter values are $Y_{s}=0.1, n=0.45, U=0.1$ .

5.5.2. Elasto-plastic breakup regime, ${0.33}\lt {n}\lt {0.43}$

In this section we will examine the pinching dynamics of materials with smaller power-law exponent $0.33\lt n\lt 0.43$ and with as low $n$ values (always positive) as our algorithm allows with convergence in a reasonable time. As in the previous regime, the YS does not enter the dominant balance in this regime either, but its title again includes the term ‘plastic’, because part of the appearing elastic term is what remains from the viscoplastic term in the constitutive law. A small decrease in $n$ below 0.45 leads to a significant deviation from the asymptotic result for the minimum radius. For the values of $0.33\lt n\lt 0.43$ , the breakup mechanism is no longer governed by capillarity. Instead, it is dictated purely by elastic stresses, which drive the breakup after the neck is formed. This type of pinching-off has been identified by Renardy (Reference Renardy2002a ) for the generalised form of PTT materials that he introduced. To further investigate this regime, we carried out filament stretching simulations for three different values of $n$ where $0.33\lt n\lt 0.43.$ In the elasto-plastic breakup regime, the minimum radius follows a different asymptotic scaling $h_{\textit{min}}\sim \, (t_{0}-t)^{({n}/{4-8n})}$ . However, since surface tension does not appear in the constitutive equation, the scaling for $\tau _{\textit{zz}}$ remains unchanged, following $\tau _{\textit{zz}}\sim (t_{0}-t)^{({n}/{n-1})}$ .

Figure 18. Effect of the exponent of the strain-rate thinning (n) in elastic breakup regime on the scaling with respect to $t_{0}-t$ of minimum radius (a), axial stress (b), axial velocity (c) and axial length of neck (d). The rest of the parameter values are $Y_{s}=0.1, Ec=0.14, U=0.1$ .

This is shown in figures 18(a) and 18(b), respectively. Additionally, our calculations indicate that, by decreasing the value of $n$ , the axial velocity and axial position scaling exponents, $\alpha _{2}$ and $\delta$ approach the values of −1 and 0, respectively, as illustrated in figures 18(c) and 18(d), respectively. The exponents in this regime do not depend on $Ec$ , similar to the capillary-driven regime, and the corresponding results are omitted for conciseness. It is noteworthy that reaching the asymptotic range now requires getting even closer to the breakup time $(t_{0}-t)\lt 10^{-5}$ and fully capturing it requires approaching small $T$ values more than ever before, down to $(t_{0}-t)\sim 10^{-8}$ . All these necessitate extremely small time steps for accurate predictions. Furthermore, the extreme values of $h_{\textit{min}}, v_{z}\, {\textrm{and}}\, \tau _{\textit{zz}}$ now become $\sim 3\times 10^{-4}, 5\times 10^{5}\, \text{and }2\times 10^{5}$ , respectively, with the last two approximately one order of magnitude larger than their values for $n=0.45$ . All these observations reveal that the calculations become increasingly difficult as $n$ decreases. The deviation in the scaling exponent $\alpha _{1}$ for $n=0.4$ is most pronounced among all cases studied. Specifically, our numerical results yield $\alpha _{1} = 0.45$ , whereas the asymptotic scale predicted from slender filament equations is 0.5. Even with a more refined mesh in the necking region, including more than 10 000 axial elements results in the same value of $\alpha _{1}=0.45$ . We believe this deviation arises by the factors discussed earlier.

5.5.3. Collapse of filament shapes and radial velocity asymptotes for all cases with, ${n}\lt {1}$

Having analysed the scalings where the SHB exponent, $n$ , is less than unity, we now present self-similar interface profiles obtained by 2-D simulations for various values of $n$ , shown in figure 19. We have shown that $z$ scales as $(t_{0}-t)^{\delta }$ and that when $n$ increases, $\delta $ increases in absolute value, remaining negative. Consequently, the rescaled axial coordinate $\overline{Z}$ spans a narrower region as $n$ increases. Similarly, the span of $\overline{h}$ decreases. Both are clearly seen in figure 19. The local slenderness of the necking region decreases progressively with decreasing $n$ . This trend can be understood by examining the ratio of radial to axial length scale, given by $({h_{\textit{min}}}/{z})\sim (t_{0}-t)^{{\alpha _{1}}-\delta }$ . The variation of the exponent $\alpha _{1}-\delta$ is provided in table 3, revealing a sharp decrease from ∼ 7 to 0.45 as $n$ decreases from $0.8$ to $0.4$ . Consequently, the breakup dynamics in the necking region exhibits an abrupt transition from a fully acceptable slender filament one to one that increasingly approaches a non-slender breakup at $n=0.4$ .

Table 3. Values of exponents $\alpha _{1}, \delta$ and $\alpha -\delta$ for five different SHB exponents. The rest of parameters are $Y_{s}=0.1, Ec=0.14, U=0.1$ .

Figure 19. Self-similar interface profiles of EVP fluids with five different strain-rate-thinning exponents. The rest of the parameters are $Y_{s}=0.1, Ec=0.14, U=0.1$ .

Figure 20 shows the evolution of the radial velocity at the interface on the symmetry plane as a function of remaining time to pinch, $t_{0}-t$ . Results from 2-D simulations reveal that, for all values of the SHB exponent $(n)$ less than unity, the scale of $u_{r}$ is $\alpha _{1}-1$ . So, the radial velocity exhibits an asymptotic scale $u_{r}\, \sim ({h_{\textit{min}}}/{(t_{0}-t}))$ , being valid for all cases where $n\lt 1$ . Interestingly, as shown in figure 20, for SHB exponents $n\gt 0.5$ , the absolute radial velocity $| u_{r}|$ at the interface on the symmetry plane decreases considerably as $t_{0}-t\rightarrow 0$ , indicating an increase in material’s resistance to breakup, which is a characteristic of extension-rate-thickening behaviour. Conversely, for $n\lt 0.5$ , $| u_{r}|$ increases toward breakup, reflecting a significant reduction in elastic resistance, corresponding to extension-rate-thinning behaviour.

Figure 20. Variation of the radial velocity magnitude at the intersection between the symmetry plane and the interface with $t_{0}-t$ . The rest of the parameters are $Y_{s}=0.1, Ec=0.14, U=0.1$ .

5.5.4. Discrimination between finite-length and point breakup

In this section we elaborate further on the discussion in § 4.1 concerning the finite-length breakup. We investigate this phenomenon by examining the filament thickness $h$ at two axial locations; one at the plane of symmetry, where $h$ attains its minimum value, and another one slightly above it at $z^{+}=0.01$ . Initially, we attempted to demonstrate finite-length breakup by plotting $h_{z+}$ and $h_{\textit{min}}$ as functions of time $t$ or time to breakup $t_{0}-t$ . However, the near perfect overlap of these curves hindered any definitive conclusion. Next, we employed an alternative metric by plotting the normalised difference of filament thickness probed at these two axial locations versus time to breakup

(5.5) \begin{equation}\chi _{h}=\frac{h_{{z^{+}}}-h_{\textit{min}}}{h_{\textit{min}}}.\end{equation}

This quantity may also be experimentally accessible. If a line is drawn connecting these two points on the interface, one at $z=0$ and the other at $z=z^{+}=0.01$ , the difference $h_{{z^{+}}}-h_{\textit{min}}$ is proportional to the absolute slope of that line, since ${\unicode[Arial]{x0394}} z$ is fixed. A decrease in $h_{{z^{+}}}-h_{\textit{min}}$ indicates that the filament shape is becoming more cylindrical. Meanwhile, the decay rate of $h_{\textit{min}}$ characterises the rate at which breakup occurs at the symmetry plane. Therefore, an increase in $\chi _{h}$ as $t_{0}-t\rightarrow 0$ denotes that the rate of filament thinning leading to point breakup exceeds the rate at which a cylindrical ‘inner thread’, if present, is formed. In this context, a negative slope of $\chi _{h}$ plotted against $t_{0}-t\rightarrow 0$ in log–log plot implies that breakup will asymptotically precede the formation of a cylindrical neck, if such neck is formed at all. Conversely, a positive or zero slope suggests finite-length breakup, wherein the filament retains a finite axial extent close to breakup. Asymptotically, this implies that the cylindrical neck forms either before breakup, if $\chi _{h}\rightarrow$ 0, or simultaneously with it, if $\chi _{h}$ approaches a constant. As shown in figure 21, $\chi _{h}$ increases near breakup for both a power-law material with $n=0.4$ analysed in § 5.1 and a Newtonian fluid. Conversely, in an EVP filament with $n$ varying from 0.4 to 0.6, analysed in §§ 5.5.1 and 5.5.2, $\chi _{h}$ exhibits a decreasing or constant trend, with a plateau at O( $10^{-2}-10^{-4})$ as breakup is approached. These observations suggest that the EVP filament retains an almost cylindrical shape very near pinch-off, consistent with finite-length breakup behaviour.

Figure 21. Evolution of $\chi _{h}=({h_{{z^{+}}}-h_{\textit{min}}})/{h_{\textit{min}}}$ plotted against time to breakup $t_{0}-t$ for EVP, power-law and Newtonian fluids. Other parameters for the EVP fluid are $Y_{s}=0.1, Ec=0.14, U=0.1$ , and those for the power-law fluid are provided in § 5.1. Here, $z^{+}=0.01$ .

5.5.5. Analysis of the local curvature at the symmetry plane

In addition to analysing the scaling laws for minimum radius, axial stress or axial velocity at the interface on the symmetry plane $z=0$ , it is often important to examine the interfacial curvature and its asymptotic behaviour at this location. This becomes particularly relevant when it is important to distinguish between a SHB material with $n=0.5$ and a Newtonian fluid, because both exhibit a similar linear decay of the minimum radius. As highlighted by Du et al. (Reference Du, Ohtani, Ellwood and McKinley2024), evaluation of the curvature at the mid-plane offers a more comprehensive characterisation of the filament shape than the decay of the minimum radius alone. Following this approach, we define the curvature ratio $\chi$ at the interface on the symmetry plane as

(5.6) \begin{equation}\chi = \displaystyle\frac{\kappa _{z}}{\kappa _{r}}= \displaystyle\frac{\left| \displaystyle\frac{h_{\textit{zz}}}{\left(1+h_{z}^{2}\right)^{\frac{3}{2}}}\right| }{\left| \displaystyle\frac{1}{h\left(1+h_{z}^{2}\right)^{\frac{1}{2}}}\right| }.\end{equation}

Substituting (3.3) and (3.6) into the above expression yields a general asymptotic scaling form for $\chi$

(5.7) \begin{equation}\chi \sim \left(t_{0}-t\right)^{{2\alpha _{1}}-2\delta \, }\sim \, \left(h_{\textit{min}}\right)^{2-\frac{2\delta }{\alpha _{1}}}.\end{equation}

Figure 22 shows the evolution of the curvature ratio, $\chi$ , as a function of the of minimum radius $h_{\textit{min}}$ , neglecting inertial effects. It includes SHB fluids in the range of exponents $n\in [0.4,0.6]$ , as well a power-law fluid with $n=0.4$ and a Newtonian fluid.

Figure 22. Evolution of the curvature ratio against minimum radius for EVP, power-law and Newtonian fluids. Other parameters for EVP fluid are $Y_{s}=0.1, Ec=0.14, U=0.1$ , and those for the power-law fluid are provided in § 5.1.

According to Renardy (Reference Renardy1994), the inertialess breakup asymptotic scaling exponents of a Newtonian fluid are $\alpha _{1}=1\, \text{and }\delta =0.175$ , while for a power-law fluid with $n=0.4$ , they are $\alpha _{1}=\delta =0.4$ , as discussed in § 5.1. Consequently, the asymptotic scaling of $\chi$ is expected to follow $(h_{\textit{min}})^{1.65}$ for Newtonian fluid and remain constant for power-law fluids under the assumption of negligible inertia. Notably, figure 22 reveals that the EVP fluid with $n=0.5$ exhibits a numerically determined scaling exponent of approximately 2.35, which is significantly higher than that of the Newtonian case despite the fact that both fluids exhibit a similar linear decay in the minimum radius. This demonstrates that the evolution of the curvature ratio with respect to the minimum radius can effectively distinguish between the two cases, even though the formation of an elastic neck in the EVP fluids may not be fully observable in experiments. It is noteworthy that, having computed the curvature numerically, to evaluate curvature ratio $\chi$ , which involves the second derivative of the interface shape, we have calculated the $\phi _{2}$ term in (3.18) as well. We opted not to report it explicitly for conciseness.

5.5.6. Elasto-plasto-capillary regime with no breakup in finite time, ${n}={1}$

Increasing $n$ to unity to examine numerically the asymptotics of the Saramito–Bingham model, we revert to the asymptotes of the simpler Oldroyd-B fluids, which are known to predict that all variables and in particular the minimum radius have an exponential dependence on time to breakup. This is demonstrated in figure 23, where the minimum radius depends exponentially on time t, and the coefficient of the exponent is the same as the one for an Oldroyd-B fluid, $-1/(3\times Ec\times \ln (10))$ , while the coefficient of $\tau _{\textit{zz}}$ has the opposite sign. The other two variables also have an exponential dependence on $t$ with the same coefficient between them, as required by the asymptotics. The final values of all variables required to cover the asymptotic range are much smaller now than in cases with $n\lt 1$ , revealing that indeed the previous calculations are the most challenging ones. Once again, the YS plays no apparent role in the asymptotes.

Figure 23. Effect of $Ec$ on the scaling with respect to time $(t)$ of minimum radius (a), axial stress (b), axial velocity (c) and axial length of neck (d). The rest of the parameters are $Y_{s}=0.1, n=1., U=0.1$ .

6. Concluding remarks

In this study, we examined the dynamics of EVP filament breakup, while neglecting inertial effects as a first approach to understanding it. By utilising the SHB constitutive model, we derived the asymptotic behaviour and self-similar forms of the variables as the minimum filament radius approaches extremely small values. We show that the value of the YS is irrelevant when examining the late-stage asymptotic behaviour. Interestingly, we observed that the SHB model exhibits a similar asymptotic form as the G-PTT model, previously analysed by Renardy, (Reference Renardy2002a ). We identified three distinct asymptotic regimes: (a) elasto-plasto-capillary and (b) elasto-plastic in both of which breakup takes place in finite time and (c) the elasto-plasto-capillary regime with no breakup in finite time. The former two cases arise when n < 1 and the latter when n = 1. We did not examine cases with n > 1, corresponding to shear-rate-thickening behaviour of the material. A comprehensive summary of the derived asymptotes is presented in table 4. By employing our recently developed numerical algorithm, PEGAFEM-V, we conducted 2-D axisymmetric simulations to capture the breakup process. To achieve minimum radius values as small as $O(10^{-4}-10^{-5})$ , we developed a novel stabilised adaptive finite element method capable of simulating any breakup scenario we encountered in this study. Using the 2-D simulation results, we derived self-similar solutions for the interface shape, axial velocity and axial stress by appropriately rescaling the variables, by scales presented in table 4.

Table 4. Summary of scales of all variables for breakup of an EVP filament when $Oh^{-2}=0$ .

We demonstrated that both the YS and the elasticity, represented by the coefficient of the upper-convected derivative in the constitutive equation, do not affect the determination of the asymptotic behaviour. The effect of $Y_{s}$ is implicit for all examined $n$ values, while the effect of $Ec$ is implicit for $n\lt 1$ , affecting only the intercept of the asymptotes and explicit for $n=1$ . We expect that other variants of Saramito’s model, such as the Saramito–Giesekus model, will also exhibit the asymptotic behaviour of their viscoelastic counterpart during the late-stage breakup, as the YS becomes negligible in the dominant stress balance in the constitutive equation. We presented a scaling analysis that offers insights into experiments of extension of YS fluids with elastic properties, particularly where thixotropic effects are negligible. This analysis is very relevant to extensional rheology, where key quantities such as minimum filament radius or extensional viscosity are commonly reported.

We did not perform numerical simulations with the 1-D slender EVP filament. Following the findings of Moschopoulos et al. (Reference Moschopoulos, Syrakos, Dimakopoulos and Tsamopoulos2020) the slenderness of the filament may be influenced by the presence of YS, even though YS does not affect the late-time asymptotic behaviour. The final breakup shape and time might differ in 1-D slender simulations due to the YS affecting the filament’s initial deformation up to the point very close to breakup. Future studies could explore the comparison between the results derived from 2-D and 1-D simulations across the full parameter space of YS and elasticity.

In future work, we will extend this study to provide a comprehensive understanding of EVP filament breakup including inertia. We will assess whether this leads to interfacial instabilities and the formation of bead-on-string structures in EVP filaments. Additionally, we plan to investigate the impact of thixotropy on breakup dynamics through full 2-D axisymmetric simulations of thixo-elasto-visco-plastic constitutive models, such as those developed by Spyridakis et al. (Reference Spyridakis, Moschopoulos, Varchanis, Dimakopoulos and Tsamopoulos2024) and Varchanis et al. (Reference Varchanis, Makrigiorgos, Moschopoulos, Dimakopoulos and Tsamopoulos2019a).

Supplementary movies

Supplementary movies are available at https://doi.org/10.1017/jfm.2025.10652.

Acknowledgements

The authors express their gratitude to G. H. McKinley for insightful discussions and enlightening suggestions. Additionally, we acknowledge the valuable contributions of G. D’Avino and S. Varchanis for their helpful suggestions on the numerical approach.

Funding

This research has received funding from the European Union’s Horizon 2020 research and innovation programme under the Marie Skłodowska-Curie grant agreement No 955605.

Declaration of interests

The authors report no conflict of interest.

Data availability

Data will be made available upon request.

Author contributions

Pourya Zakeri: Software development and testing, Data curation, Investigation, Validation, Writing original draft. Pantelis Moschopoulos: Development of initial software, testing and validation, Yannis Dimakopoulos: Conceptualisation, Investigation, Methodology, Supervision, funding acquisition, Writing reviewing & editing. John Tsamopoulos: Conceptualisation, Investigation, Methodology, Validation, Supervision, Funding acquisition, Writing, reviewing & editing.

Appendix A

A.1. Mesh adaptation method

The initial mesh is a unit square of side length 1. Concerning the elliptic grid generation scheme, the computational domain remains the unit square, and the physical domain deforms based on the imposed velocity on the top plate and the deformation of the free surface. The initial unit square is discretised into 500 axial elements and 30 radial elements leading to 15 000 nodes and 135 000 unknowns. As the filament stretches and the neck forms, the elements near and inside the neck will become skewed. To address this, we add nodes in the radial cross-section in the middle of every element in the computational domain if the minimum angle of the corresponding element in the physical domain falls below 25 degrees. Figure 24 shows schematically the h-adaptive mesh refinement after 1 step of node insertion. The mesh is shown in the computational domain for simplicity. However, adding nodes only based on the angle criterion may not be enough to have good accuracy of the solution. To address this, and motivated by our Petrov–Galerkin formulation, we define a characteristic pressure gradient for element $k$ as follows:

(A1) \begin{equation}\boldsymbol{\nabla }p_{k}=\frac{{\Sigma} _{i=1}^{n_{en}}\left\| \boldsymbol{\nabla }\!P_{i}^{h}\right\| }{n_{en}},\end{equation}

Figure 24. Schematic of h-adaptive mesh refinement. Nodes are added radially where any element exceeds the adaptation criteria.

where $n_{en}$ denotes the number of nodes in the element and $\| \boldsymbol{\nabla }P_{i}^{h}\|$ is the L2-norm of the pressure gradient at each node. We use a pressure gradient-based error indicator defined as

(A2) \begin{equation}\eta _{k}=\boldsymbol{\nabla }p_{k}\boldsymbol{\cdot }(\textrm{Area})_{k}.\end{equation}

The error indicator $\eta _{k}$ is normalised by the averaged error indication $\, \overline{\eta }_{k}$ over the domain

(A3) \begin{equation}\overline{\eta }_{k}=\frac{{\Sigma} _{i=1}^{n_{el}}\eta _{{k_{i}}}}{n_{el}}.\end{equation}

where $n_{el}$ is the total number of elements. A node is added in the middle of every element if the ratio $({\eta _{k}}/{\overline{\eta }_{k}})$ exceeds 2. An additional criterion is developed by trial and error to prevent excessive node addition. Employing this error indicator results in less than 5 iterations per time step to reach the convergence criterion and achieve robust accuracy of the computed solution. Figure 25 depicts the adapted mesh just before breakup which contains more than $4\times 10^{5}$ nodes and $3.6\times 10^{6}$ unknowns.

Figure 25. The adapted mesh with more than 400 000 nodes for the stretched filament; here $n=0.45, Y_{s}=1.25, EC=0.14\, \textrm{and}\, U=0.39$ .

A.2. Solution transfer method

After every adaptation step, values on the newly generated nodes must be computed. One straightforward method is to use the interpolants of the parent element to calculate these values. However, this method is not accurate enough, as we experienced in our calculations. This linear interpolation introduces inaccuracy, particularly in the pressure field, leading to spurious oscillations despite the use of a stabilised finite element method. Hence, to calculate the flow variables on the newly generated nodes, we use the conservative interpolation method introduced by Farrell & Maddison (Reference Farrell and Maddison2011). This method is a local L2-Galerkin projection which tries to minimise the L2 norm of a field between the old mesh and the new mesh by constructing a local supermesh. Supermesh is the mesh of the intersections of the new and the old mesh. For further details, the reader is referred to Farrell & Maddison (Reference Farrell and Maddison2011). The minimisation statement and the weak form of interpolation between the two meshes are as follows:

(A4) \begin{equation}\textit{Minimize}\, \left| \left| q_{old}-q_{\textit{new}}\right| \right| _{2}\,\rightarrow \, \int _{{{\Omega} _{\textit{new}}}}q_{\textit{new}}{\textrm{d}}{\Omega} =\int _{{{\Omega} _{\textit{supermesh}}}}q_{\textit{old}}{\textrm{d}}{\Omega}.\end{equation}

The integration on the left- and right-hand sides is conducted over the new mesh and the supermesh, respectively. This method ensures local conservation of integrals at the element level resulting in a smooth and continuous computed solution over time. The h-adaption method described here is straightforward, simple, and eliminates the need for the intersecting algorithm described in Farrell & Maddison (Reference Farrell and Maddison2011) to construct the supermesh. Figure 26 illustrates the supermesh constructed after one step of the h-adaptive mesh refinement. As the nodes of the old mesh, the new mesh and their local supermesh are predefined, implementing a code to generate the supermesh is both intuitive and computationally efficient.

Figure 26. Constructed supermesh after one step of an h-adaptive mesh refinement.

References

Anna, S.L. & McKinley, G.H. 2001 Elasto-capillary thinning and breakup of model elastic liquids. J. Rheol. 45 (1), 115138.10.1122/1.1332389CrossRefGoogle Scholar
Arnolds, O., Buggisch, H., Sachsenheimer, D. & Willenbacher, N. 2010 Capillary breakup extensional rheometry (CaBER) on semi-dilute and concentrated polyethyleneoxide (PEO) solutions. Rheol. Acta 49 (11), 12071217.10.1007/s00397-010-0500-7CrossRefGoogle Scholar
Balmforth, N.J., Dubash, N. & Slim, A.C. 2010a Extensional dynamics of viscoplastic filaments: i. Long-wave approximation and the Rayleigh instability. J. Non-Newtonian Fluid Mech. 165 (19–20), 11391146.10.1016/j.jnnfm.2010.05.012CrossRefGoogle Scholar
Balmforth, N.J., Dubash, N. & Slim, A.C. 2010b Extensional dynamics of viscoplastic filaments: II. Drips and bridges. J. Non-Newtonian Fluid Mech. 165 (19–20), 11471160.10.1016/j.jnnfm.2010.06.004CrossRefGoogle Scholar
Bechtel, S.E., Cao, J.Z. & Forest, M.G. 1992 Practical application of a higher order perturbation theory for slender viscoelastic jets and fibers. J. Non-Newtonian Fluid Mech. 41 (3), 201273.10.1016/0377-0257(92)87001-RCrossRefGoogle Scholar
Bhat, P.P., Appathurai, S., Harris, M.T., Pasquali, M., McKinley, G.H. & Basaran, O.A. 2010 Formation of beads-on-a-string structures during break-up of viscoelastic filaments. Nat. Phys. 6 (8), 625631.10.1038/nphys1682CrossRefGoogle Scholar
Brenner, M.P., Lister, J.R. & Stone, H.A. 1996 Pinching threads, singularities and the number 0.0304. Phys. Fluids 8 (11), 28272836.10.1063/1.869086CrossRefGoogle Scholar
Chambon, G., Freydier, P., Naaim, M. & Vila, J.-P. 2020 Asymptotic expansion of the velocity field within the front of viscoplastic surges: comparison with experiments. J. Fluid Mech. 884, A43.10.1017/jfm.2019.943CrossRefGoogle Scholar
Chatzidai, N., Giannousakis, A., Dimakopoulos, Y. & Tsamopoulos, J. 2009 On the elliptic mesh generation in domains containing multiple inclusions and undergoing large deformations. J. Comput. Phys. 228 (6), 19802011.10.1016/j.jcp.2008.11.020CrossRefGoogle Scholar
Chen, A.U., Notz, P.K. & Basaran, O.A. 2002 Computational and experimental analysis of pinch-off and scaling. Phys. Rev. Lett. 88 (17), 174501.10.1103/PhysRevLett.88.174501CrossRefGoogle ScholarPubMed
Clasen, C., Eggers, J., Fontelos, M.A., Li, J. & McKinley, G.H. 2006 The beads-on-string structure of viscoelastic threads. J. Fluid Mech. 556, 283308.10.1017/S0022112006009633CrossRefGoogle Scholar
Deen, W.M. 2012 Analysis of Transport Phenomena. 2nd edn. Oxford University Press.Google Scholar
Dimakopoulos, Y. & Tsamopoulos, J. 2003 A quasi-elliptic transformation for moving boundary problems with large anisotropic deformations. J. Comput. Phys. 192 (2), 494522.10.1016/j.jcp.2003.07.027CrossRefGoogle Scholar
Dimitriou, C.J. & McKinley, G.H. 2014 A comprehensive constitutive law for waxy crude oil: a thixotropic yield stress fluid. Soft Matt. 10 (35), 66196644.10.1039/C4SM00578CCrossRefGoogle ScholarPubMed
Doshi, P., Suryo, R., Yildirim, O.E., McKinley, G.H. & Basaran, O.A. 2003 Scaling in pinch-off of generalized Newtonian fluids. J. Non-Newtonian Fluid Mech. 113 (1), 127.10.1016/S0377-0257(03)00081-8CrossRefGoogle Scholar
Du, J., Ohtani, H., Ellwood, K. & McKinley, G.H. 2024 Capillarity-driven thinning and breakup of weakly rate-thickening fluids. J. Non-Newtonian Fluid Mech. 331, 105294.10.1016/j.jnnfm.2024.105294CrossRefGoogle Scholar
Eggers, J. 1993 Universal pinching of 3D axisymmetric free-surface flow. Phys. Rev. Lett. 71 (21), 3458.10.1103/PhysRevLett.71.3458CrossRefGoogle ScholarPubMed
Eggers, J. & Dupont, T.F. 1994 Drop formation in a one-dimensional approximation of the Navier–Stokes equation. J. Fluid Mech. 262, 205221.10.1017/S0022112094000480CrossRefGoogle Scholar
Eggers, J., Herrada, M.A. & Snoeijer, J.H. 2020 Self-similar breakup of polymeric threads as described by the Oldroyd-B model. J. Fluid Mech. 887, A19.10.1017/jfm.2020.18CrossRefGoogle Scholar
Esposito, G., Dimakopoulos, Y. & Tsamopoulos, J. 2024 Buoyancy induced motion of a Newtonian drop in elastoviscoplastic materials. J. Rheol. 68 (5), 815835.10.1122/8.0000883CrossRefGoogle Scholar
Farrell, P.E. & Maddison, J.R. 2011 Conservative interpolation between volume meshes by local Galerkin projection. Comput. Meth. Appl. Mech. 200 (1–4), 89100.10.1016/j.cma.2010.07.015CrossRefGoogle Scholar
Fontelos, M.A. & Li, J. 2004 On the evolution and rupture of filaments in Giesekus and FENE models. J. Non-Newtonian Fluid Mech. 118 (1), 116.10.1016/j.jnnfm.2004.02.002CrossRefGoogle Scholar
Fraggedakis, D., Dimakopoulos, Y. & Tsamopoulos, J. 2016 Yielding the yield stress analysis: a thorough comparison of recently proposed elasto-visco-plastic (EVP) fluid models. J. Non-Newtonian Fluid Mech. 238, 170188.10.1016/j.jnnfm.2016.11.007CrossRefGoogle Scholar
Freydier, P., Chambon, G. & Naaim, M. 2017 Experimental characterization of velocity fields within the front of viscoplastic surges down an incline. J. Non-Newtonian Fluid Mech. 240, 5669.10.1016/j.jnnfm.2017.01.002CrossRefGoogle Scholar
Frigaard, I. 2019 Simple yield stress fluids. Curr. Opin. Colloid Interface Sci. 43, 8093.10.1016/j.cocis.2019.03.002CrossRefGoogle Scholar
Holenberg, Y., Lavrenteva, O.M., Shavit, U. & Nir, A. 2012 Particle tracking velocimetry and particle image velocimetry study of the slow motion of rough and smooth solid spheres in a yield-stress fluid. Phys. Rev. E - Stat. Nonlinear Soft Matt. Phys. 86 (6), 066301.10.1103/PhysRevE.86.066301CrossRefGoogle Scholar
Huisman, F.M., Friedman, S.R. & Taborek, P. 2012 Pinch-off dynamics in foams, emulsions and suspensions. Soft Matt. 8 (25), 67676774.10.1039/c2sm25240fCrossRefGoogle Scholar
Keller, J.B. & Miksis, M.J. 1983 Surface tension driven flows. SIAM J. Appl. Maths 43 (2), 268277.10.1137/0143018CrossRefGoogle Scholar
Keshavarz, B., Sharma, V., Houze, E.C., Koerner, M.R., Moore, J.R., Cotts, P.M., Threlfall-Holmes, P. & McKinley, G.H. 2015 Studying the effects of elongational properties on atomization of weakly viscoelastic solutions using Rayleigh Ohnesorge Jetting Extensional Rheometry (ROJER). J. Non-Newtonian Fluid Mech. 222, 171189.10.1016/j.jnnfm.2014.11.004CrossRefGoogle Scholar
Kordalis, A., Dimakopoulos, Y. & Tsamopoulos, J. 2024 Hydrodynamic interaction between coaxially rising bubbles in elasto-visco-plastic materials: bubbles with a wide range of relative sizes. Phys. Rev. Fluids 9 (9), 093301.10.1103/PhysRevFluids.9.093301CrossRefGoogle Scholar
Kordalis, A., Pema, D., Androulakis, S., Dimakopoulos, Y. & Tsamopoulos, J. 2023 Hydrodynamic interaction between coaxially rising bubbles in elastoviscoplastic materials: equal bubbles. Phys. Rev. Fluids 8 (8), 083301.10.1103/PhysRevFluids.8.083301CrossRefGoogle Scholar
Kordalis, A., Varchanis, S., Ioannou, G., Dimakopoulos, Y. & Tsamopoulos, J. 2021 Investigation of the extensional properties of elasto-visco-plastic materials in cross-slot geometries. J. Non-Newtonian Fluid Mech. 296, 104627.10.1016/j.jnnfm.2021.104627CrossRefGoogle Scholar
Liu, Y., Balmforth, N.J., Hormozi, S. & Hewitt, D.R. 2016 Two–dimensional viscoplastic dambreaks. J. Non-Newtonian Fluid Mech. 238, 6579.10.1016/j.jnnfm.2016.05.008CrossRefGoogle Scholar
Lopez, W.F., Naccache, M.F., De, P.R. & Mendes, S. 2017 Rising bubbles in yield stress materials. J. Rheol. 62 (1), 209.10.1122/1.4995348CrossRefGoogle Scholar
Luu, L.-H., Philippe, P. & Chambon, G. 2017 Flow of a yield-stress fluid over a cavity: experimental study of the solid–fluid interface. J. Non-Newtonian Fluid Mech. 245, 2537.10.1016/j.jnnfm.2017.04.011CrossRefGoogle Scholar
Ma, L. & Barbosa-Chovas, G.V. 1995 Rheological characterization of mayonnaise. Part II: flow and viscoelastic properties at different oil and Xanthan gum concentrations. J. Food Engng 25 (3), 409425.10.1016/0260-8774(94)00010-7CrossRefGoogle Scholar
Martinie, L., Buggisch, H. & Willenbacher, N. 2013 Apparent elongational yield stress of soft matter. J. Rheol. 57 (2), 627646.10.1122/1.4789785CrossRefGoogle Scholar
McKinley, G.H. 2005 Visco-elasto-capillary thinning and break-up of complex fluids. Annu. Rheol. Rev., https://dspace.mit.edu/handle/1721.1/18085.Google Scholar
McKinley, G.H. & Sridhar, T. 2002 Filament-stretching rheometry of complex fluids. Annu. Rev. Fluid Mech. 34 (1), 375415.10.1146/annurev.fluid.34.083001.125207CrossRefGoogle Scholar
Moon, H., Yamani, S., McKinley, G.H. & Lee, J. 2024 Tuning the shear and extensional rheology of semi-flexible polyelectrolyte solutions, arXiv Preprint arXiv:2410.15132.Google Scholar
Moschopoulos, P., Kouni, E., Psaraki, K., Dimakopoulos, Y. & Tsamopoulos, J. 2023 Dynamics of elastoviscoplastic filament stretching. Soft Matt. 19 (25), 47174736.10.1039/D3SM00143ACrossRefGoogle ScholarPubMed
Moschopoulos, P., Spyridakis, A., Dimakopoulos, Y. & Tsamopoulos, J. 2024 Unravelling the existence of asymmetric bubbles in viscoelastic fluids. J. Fluid Mech. 985, A30.10.1017/jfm.2024.316CrossRefGoogle Scholar
Moschopoulos, P., Spyridakis, A., Varchanis, S., Dimakopoulos, Y. & Tsamopoulos, J. 2021 The concept of elasto-visco-plasticity and its application to a bubble rising in yield stress fluids. J. Non-Newtonian Fluid Mech. 297, 104670.10.1016/j.jnnfm.2021.104670CrossRefGoogle Scholar
Moschopoulos, P., Syrakos, A., Dimakopoulos, Y. & Tsamopoulos, J. 2020 Dynamics of viscoplastic filament stretching. J. Non-Newtonian Fluid Mech. 284, 104371.10.1016/j.jnnfm.2020.104371CrossRefGoogle Scholar
Moschopoulos, P., Varchanis, S., Syrakos, A., Dimakopoulos, Y. & Tsamopoulos, J. 2022 S-PAL: a stabilized finite element formulation for computing viscoplastic flows. J. Non-Newtonian Fluid Mech. 309, 104883.10.1016/j.jnnfm.2022.104883CrossRefGoogle Scholar
Mougin, N., Magnin, A. & Piau, J.M. 2012 The significant influence of internal stresses on the dynamics of bubbles in a yield stress fluid. J. Non-Newtonian Fluid Mech. 171–172, 4255.Google Scholar
Niedzwiedz, K., Arnolds, O., Willenbacher, N. & Brummer, R. 2009 41969-1 how to characterize yield stress fluids with capillary breakup extensional rheometry (CaBER)? Appl. Rheol. 19 (4), 41969).Google Scholar
Niedzwiedz, K., Buggisch, H. & Willenbacher, N. 2010 Extensional rheology of concentrated emulsions as probed by capillary breakup elongational rheometry (CaBER). Rheol. Acta 49 (11), 11031116.10.1007/s00397-010-0477-2CrossRefGoogle Scholar
Notz, P.K. & Basaran, O.A. 2004 Dynamics and breakup of a contracting liquid filament. J. Fluid Mech. 512, 223256.10.1017/S0022112004009759CrossRefGoogle Scholar
Papageorgiou, D.T. 1995 On the breakup of viscous liquid threads. Phys. Fluids 7 (7), 15291544.10.1063/1.868540CrossRefGoogle Scholar
Putz, A.M.V., Burghelea, T.I., Frigaard, I.A. & Martinez, D.M. 2008 Settling of an isolated spherical particle in a yield stress shear thinning fluid. Phys. Fluids 20 (3), 033102.10.1063/1.2883937CrossRefGoogle Scholar
Renardy, M. 1994 Some comments on the surface-tension driven break-up (or the lack of it) of viscoelastic jets. J. Non-Newtonian Fluid Mech. 51 (1), 97107.10.1016/0377-0257(94)85005-4CrossRefGoogle Scholar
Renardy, M. 1995 A numerical study of the asymptotic evolution and breakup of Newtonian and viscoelastic jets. J. Non-Newtonian Fluid Mech. 59 (2–3), 267282.10.1016/0377-0257(95)01375-6CrossRefGoogle Scholar
Renardy, M. 2002 a Self-similar jet breakup for a generalized PTT model. J. Non-Newtonian Fluid Mech. 103 (2–3), 261269.10.1016/S0377-0257(02)00007-1CrossRefGoogle Scholar
Renardy, M. 2002 b Similarity solutions for jet breakup for various models of viscoelastic fluids. J. Non-Newtonian Fluid Mech. 104 (1), 6574.10.1016/S0377-0257(02)00016-2CrossRefGoogle Scholar
Renardy, M. & Renardy, Y. 2004 Similarity solutions for breakup of jets of power law fluids. J. Non-Newtonian Fluid Mech. 122 (1–3), 303312.10.1016/j.jnnfm.2004.01.026CrossRefGoogle Scholar
Rosello, M., Sur, S., Barbet, B. & Rothstein, J.P. 2019 Dripping-onto-substrate capillary breakup extensional rheometry of low-viscosity printing inks. J. Non-Newtonian Fluid Mech. 266, 160170.10.1016/j.jnnfm.2019.03.006CrossRefGoogle Scholar
Saramito, P. 2007 A new constitutive equation for elastoviscoplastic fluid flows. J. Non-Newtonian Fluid Mech. 145 (1), 114.10.1016/j.jnnfm.2007.04.004CrossRefGoogle Scholar
Saramito, P. 2009 A new elastoviscoplastic model based on the Herschel–Bulkley viscoplastic model. J. Non-Newtonian Fluid Mech. 158 (1–3), 154161.10.1016/j.jnnfm.2008.12.001CrossRefGoogle Scholar
Spyridakis, A., Moschopoulos, P., Varchanis, S., Dimakopoulos, Y. & Tsamopoulos, J. 2024 Thixo-elastoviscoplastic modeling of human blood. J. Rheol. 68 (1), 123.10.1122/8.0000711CrossRefGoogle Scholar
Suryo, R. & Basaran, O.A. 2006 Local dynamics during pinch-off of liquid threads of power law fluids: scaling analysis and self-similarity. J. Non-Newtonian Fluid Mech. 138 (2–3), 134160.10.1016/j.jnnfm.2006.04.008CrossRefGoogle Scholar
Townsend, J.M., Beck, E.C., Gehrke, S.H., Berkland, C.J. & Detamore, M.S. 2019 Flow behavior prior to crosslinking: the need for precursor rheology for placement of hydrogels in medical applications and for 3D bioprinting. Prog. Polym. Sci. 91, 126140.10.1016/j.progpolymsci.2019.01.003CrossRefGoogle ScholarPubMed
Tuladhar, T.R. & Mackley, M.R. 2008 Filament stretching rheometry and break-up behaviour of low viscosity polymer solutions and inkjet fluids. J. Non-Newtonian Fluid Mech. 148 (1–3), 97108.10.1016/j.jnnfm.2007.04.015CrossRefGoogle Scholar
Turkoz, E., Lopez-Herrera, J.M., Eggers, J., Arnold, C.B. & Deike, L. 2018 Axisymmetric simulation of viscoelastic filament thinning with the Oldroyd-B model. J. Fluid Mech. 851, R2.10.1017/jfm.2018.514CrossRefGoogle Scholar
Van Der Kolk, J., Tieman, D. & Jalaal, M. 2023 Viscoplastic lines: printing a single filament of yield stress material on a surface. J. Fluid Mech. 958, A34.10.1017/jfm.2023.118CrossRefGoogle Scholar
Van Der Zanden, J. & Hulsen, M. 1988 Mathematical and physical requirements for successful computations with viscoelastic fluid models. J. Non-Newtonian Fluid Mech. 29, 93117.10.1016/0377-0257(88)85052-3CrossRefGoogle Scholar
Varchanis, S., Dimakopoulos, Y., Wagner, C. & Tsamopoulos, J. 2018 How viscoelastic is human blood plasma? Soft Matt. 14 (21), 42384251.10.1039/C8SM00061ACrossRefGoogle ScholarPubMed
Varchanis, S., Haward, S.J., Hopkins, C.C., Syrakos, A., Shen, A.Q., Dimakopoulos, Y. & Tsamopoulos, J. 2020 a Transition between solid and liquid state of yield-stress fluids under purely extensional deformations. Proc. Natl Acad. Sci. 117 (23), 1261112617.10.1073/pnas.1922242117CrossRefGoogle ScholarPubMed
Varchanis, S., Makrigiorgos, G., Moschopoulos, P., Dimakopoulos, Y. & Tsamopoulos, J. 2019 a Modeling the rheology of thixotropic elasto-visco-plastic materials. J. Rheol. 63 (4), 609639.10.1122/1.5049136CrossRefGoogle Scholar
Varchanis, S., Pettas, D., Dimakopoulos, Y. & Tsamopoulos, J. 2021 Origin of the Sharkskin instability: nonlinear dynamics. Phys. Rev. Lett. 127 (8), 088001.10.1103/PhysRevLett.127.088001CrossRefGoogle ScholarPubMed
Varchanis, S., Syrakos, A., Dimakopoulos, Y. & Tsamopoulos, J. 2019 b A new finite element formulation for viscoelastic flows: circumventing simultaneously the LBB condition and the high-Weissenberg number problem. J. Non-Newtonian Fluid Mech. 267, 7897.10.1016/j.jnnfm.2019.04.003CrossRefGoogle Scholar
Varchanis, S., Syrakos, A., Dimakopoulos, Y. & Tsamopoulos, J. 2020 b PEGAFEM-v: a new Petrov–Galerkin finite element method for free surface viscoelastic flows. J. Non-Newtonian Fluid Mech. 284, 104365.10.1016/j.jnnfm.2020.104365CrossRefGoogle Scholar
Yildirim, O.E. & Basaran, O.A. 2001 Deformation and breakup of stretching bridges of newtonian and shear-thinning liquids: comparison of one-and two-dimensional models. In Chemical Engineering Science, vol. 56.Google Scholar
Zhang, X., Padgett, R.S. & Basaran, O.A. 1996 Nonlinear deformation and breakup of stretching liquid bridges. J. Fluid Mech. 329, 207245.10.1017/S0022112096008907CrossRefGoogle Scholar
Figure 0

Figure 1. Schematic of the axisymmetric EVP filament stretching geometry.

Figure 1

Table 1. Dimensionless numbers.

Figure 2

Figure 2. The evolution of minimum radius vs. time, for parameter values $n=0.45,\, Ec=0.14,\, Y_{s}=1.25,\, U=0.39$.

Figure 3

Figure 3. Evolution of the axial normal stress, $\tau _{\textit{zz}}$, vs. time to pinch in uniaxial extension flow with $\dot{\epsilon }={2}/{t_{0}-t}$. The other EVP parameters are $n=0.45,\, Y_{S}=1$, resulting in ${n}/{n-1}=-0.8181$.

Figure 4

Figure 4. The scaling exponent of axial normal stress, $\tau _{\textit{zz}}$, vs. the SHB strain-rate-thinning exponent (n). The other EVP parameters are $Ec=0.1,\, Y_{S}=1$.

Figure 5

Figure 5. Pinching of a power-law thread with $n=0.4,\, Oh^{-2}=0,\, L=2,\, U=0$; (a) interface profile of the entire filament just before the breakup, (b) close-up near the pinch-off point, (c) minimum radius, axial position and velocity plotted as a function of time to pinch. The calculated slopes are in perfect agreement with the ones reported in Suryo & Basaran (2006).

Figure 6

Table 2. Parameters for the base case study.

Figure 7

Figure 6. Time evolution of filament under constant stretching with $U$ = 0.39. The red and blue areas indicate yielded and unyielded regions, respectively. The other parameters are $Oh^{-2}=0,\, n=0.45,\, Y_{s}=1.25,\, Ec=0.14,\, L_{0}=2$.

Figure 8

Figure 7. Contours of velocity and stress components of the filament under constant stretching with $U$ = 0.39 just before pinch-off, at $t=2.8984$. Other parameters are $Oh^{-2}=0, n=0.45, Y_{s}=1.25, Ec=0.14, L_{0}=2$.

Figure 9

Figure 8. Contours of pressure and stress components of the filament under constant stretching with $U$ = 0.39 just before pinch-off, at $t=2.8984$. Other parameters are $Oh^{-2}=0, n=0.45, Y_{s}=1.25, Ec=0.14, L_{0}=2$.

Figure 10

Figure 9. Nonlinear regression employed to extract the pinch-off time, $t_{0}$, from the numerical results of the base case study, for $Oh^{-2}=0, n=0.45, Y_{s}=1.25, Ec=0.14, L_{0}=2$.

Figure 11

Figure 10. Decay of the minimum radius, growth of axial velocity, axial position and axial stresses as pinching time is approached for $Oh^{-2}=0, n=0.45, Ec=0.14, Y_{s}=1.25, U=0.39$ ($({n}/{1-n})=0.818$. The slopes determine the corresponding exponents of the power-law scaling functions.

Figure 12

Figure 11. (Inset) Transient interface shape obtained by 2-D simulations at five different $h_{\textit{min}}$ values. (Main figure) Rescaled interface using the numerically determined scaling laws for the same five $h_{\textit{min}}$ values. The main figure illustrates convergence to a self-similar interface profile. The very large difference in the magnitudes of $h$ before and after scaling is noteworthy.

Figure 13

Figure 12. (Inset) Transient velocity variation along axial coordinate (z) at the interface obtained by 2-D simulations at five different $h_{\textit{min}}$ values. (Main figure) Rescaled velocity plotted against rescaled axial coordinate using the numerically determined scaling laws for the same five $h_{\textit{min}}$ values. The main figure demonstrates convergence to a self-similar velocity profile. The very large difference in the magnitudes of the velocity before and after scaling is noteworthy.

Figure 14

Figure 13. (Inset) Transient axial stress variation along axial coordinate (z) at the interface obtained by 2-D simulations at five different $h_{\textit{min}}$ values. (Main figure) Rescaled axial stress plotted against rescaled axial coordinate using the numerically determined scaling laws for the same five $h_{\textit{min}}$ values. The main figure demonstrates convergence to a self-similar axial stress profile. The very large difference in the magnitudes of stress before and after scaling is noteworthy.

Figure 15

Figure 14. Profiles of the entire filament (a) and closeup views in the corner (b) and necking (c) regions for four different values of $n$. In (b) we did not include the entire ‘local’ interface for $n=0.8$, because it would reduce the clarity of the rest. The remaining parameter values are $Y_{s}=0.1, Ec=0.14, U=0.1$.

Figure 16

Figure 15. Effect of the exponent of strain-rate thinning on the scaling with respect to $(t_{o}-t)$ of minimum radius (a), axial stress (b), axial velocity (c) and axial length of neck (d). The rest of the parameter values are $Y_{s}=0.1, Ec=0.14, U=0.1$.

Figure 17

Figure 16. Profiles of the entire filament (a) and closeup views in the corn (b) and necking (c) regions for three different values of$\, Ec$. The rest of the parameter values are $Y_{s}=0.1, n=0.45, U=0.1$.

Figure 18

Figure 17. Effect of $Ec$ on the scaling with respect to $(t_{o}-t)$ of minimum radius (a), axial stress (b), axial velocity(c) and axial length of neck (d). The rest of the parameter values are $Y_{s}=0.1, n=0.45, U=0.1$.

Figure 19

Figure 18. Effect of the exponent of the strain-rate thinning (n) in elastic breakup regime on the scaling with respect to $t_{0}-t$ of minimum radius (a), axial stress (b), axial velocity (c) and axial length of neck (d). The rest of the parameter values are $Y_{s}=0.1, Ec=0.14, U=0.1$.

Figure 20

Table 3. Values of exponents $\alpha _{1}, \delta$ and $\alpha -\delta$ for five different SHB exponents. The rest of parameters are $Y_{s}=0.1, Ec=0.14, U=0.1$.

Figure 21

Figure 19. Self-similar interface profiles of EVP fluids with five different strain-rate-thinning exponents. The rest of the parameters are $Y_{s}=0.1, Ec=0.14, U=0.1$.

Figure 22

Figure 20. Variation of the radial velocity magnitude at the intersection between the symmetry plane and the interface with $t_{0}-t$. The rest of the parameters are $Y_{s}=0.1, Ec=0.14, U=0.1$.

Figure 23

Figure 21. Evolution of $\chi _{h}=({h_{{z^{+}}}-h_{\textit{min}}})/{h_{\textit{min}}}$ plotted against time to breakup $t_{0}-t$ for EVP, power-law and Newtonian fluids. Other parameters for the EVP fluid are $Y_{s}=0.1, Ec=0.14, U=0.1$, and those for the power-law fluid are provided in § 5.1. Here, $z^{+}=0.01$.

Figure 24

Figure 22. Evolution of the curvature ratio against minimum radius for EVP, power-law and Newtonian fluids. Other parameters for EVP fluid are $Y_{s}=0.1, Ec=0.14, U=0.1$, and those for the power-law fluid are provided in § 5.1.

Figure 25

Figure 23. Effect of $Ec$ on the scaling with respect to time $(t)$ of minimum radius (a), axial stress (b), axial velocity (c) and axial length of neck (d). The rest of the parameters are $Y_{s}=0.1, n=1., U=0.1$.

Figure 26

Table 4. Summary of scales of all variables for breakup of an EVP filament when $Oh^{-2}=0$.

Figure 27

Figure 24. Schematic of h-adaptive mesh refinement. Nodes are added radially where any element exceeds the adaptation criteria.

Figure 28

Figure 25. The adapted mesh with more than 400 000 nodes for the stretched filament; here $n=0.45, Y_{s}=1.25, EC=0.14\, \textrm{and}\, U=0.39$.

Figure 29

Figure 26. Constructed supermesh after one step of an h-adaptive mesh refinement.

Supplementary material: File

Zakeri et al. supplementary movie 1

Evolution of the mesh during the simulation of the base case: the entire domain is shown on the left, and the neck region, where continuous mesh refinement occurs, is shown on the right.
Download Zakeri et al. supplementary movie 1(File)
File 16.5 MB
Supplementary material: File

Zakeri et al. supplementary movie 2

Time evolution of the filament interface during the base case simulation: the right side shows contours of the axial velocity uz, while the left side shows contours of the axial stress component τzz.
Download Zakeri et al. supplementary movie 2(File)
File 6.8 MB