Hostname: page-component-5b777bbd6c-gcwzt Total loading time: 0 Render date: 2025-06-19T15:09:43.871Z Has data issue: false hasContentIssue false

A modelling framework for jet penetration into soft gels

Published online by Cambridge University Press:  22 May 2025

S.P. Mousavi
Affiliation:
Department of Chemical Engineering, Université Laval, Québec, QC G1V 0A6, Canada
H. Hassanzadeh
Affiliation:
Department of Chemical Engineering, Université Laval, Québec, QC G1V 0A6, Canada
F. Larachi
Affiliation:
Department of Chemical Engineering, Université Laval, Québec, QC G1V 0A6, Canada
C.D. Ohl
Affiliation:
Department Soft Matter, Institute of Physics & Faculty of Natural Sciences, Otto-von-Guericke-University, Universitatsplatz 2, Magdeburg 39106, Germany
S.M. Taghavi*
Affiliation:
Department of Chemical Engineering, Université Laval, Québec, QC G1V 0A6, Canada
*
Corresponding author: S.M. Taghavi, Seyed-Mohammad.Taghavi@gch.ulaval.ca

Abstract

Jet penetration into soft gels is essential for optimising fluid delivery in medical therapies, biomedical engineering, and soft robotics. In this work, we visualise the jet flow of a Newtonian fluid into a soft viscoplastic gel using camera imaging and time-resolved tomographic particle image velocimetry (PIV) systems. The flow is primarily governed by the Reynolds number ($Re = 350-5000$) and the effective viscosity ratio ($m$ up to 22). We observe three flow regimes – mixing, jellyfish, and fingering – with transitions between them quantified in the $Re-m$ plane. An experimentally informed, systematic, practical, semi-analytical modelling framework is developed to estimate jet penetration depth over time, incorporating PIV results and an approximate functional decomposition approach to describe the velocity distribution and Reynolds stress contributions. The model provides reasonable estimations across all three regimes.

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

1. Introduction

The study of jets has been a long-standing area of interest, beginning with da Vinci’s foundational observations (da Vinci Reference da Vinci1510) and advanced by Stokes (Reference Stokes1851), Rayleigh (Reference Rayleigh1878), and Reynolds (Reference Reynolds1962). These early investigations laid the groundwork for understanding jet dynamics, which continue to drive modern research into complex jet interactions with soft gels. Understanding jet penetration into soft gels (Bantawa et al. Reference Bantawa, Keshavarz, Geri, Bouzid, Divoux, McKinley and Del Gado2023; Li & Gong Reference Li and Gong2024), whose viscoplasticity endows them with the unique ability to behave as both fluids and solids (Balmforth et al. Reference Balmforth, Frigaard and Ovarlez2014; Bonn et al. Reference Bonn, Denn, Berthier, Divoux and Manneville2017), is crucial for optimising fluid delivery across various fields. In medical applications, the penetration depth of jets in soft gels used as scaffolds for tissue engineering (Bailey & Appel Reference Bailey and Appel2024) determines the efficacy of therapeutic delivery systems (Taheri et al. Reference Taheri, Bao, He, Mohammadi, Ravanbakhsh, Lessard, Li and Mongeau2022), including needle-free injections (Jones et al. Reference Jones, Shen, Walter, LaBranche, Wyatt, Tomaras, Montefiori, Moss, Barouch and Clements2019; Schoppink & Rivas Reference Schoppink and Rivas2022). Jet penetration is also essential for bio-printing (Xie et al. Reference Xie, Shi, Zhang, Ge, Zhang, Chen, Fu, Xie and He2022), liquid-in-liquid printing (Bazazi et al. Reference Bazazi, Stone and Hejazi2022; Xie et al. Reference Xie, Xu, Yu, Jiang, Li and Feng2023), and improving precision in soft robotics for tasks such as gripping and manipulating delicate objects (Cianchetti et al. Reference Cianchetti, Laschi, Menciassi and Dario2018). Despite these advancements, a fundamental question remains: how does jet penetration evolve over time, and to what extent can a jet penetrate a non-Newtonian soft gel upon injection?

Numerical models and velocimetry techniques, such as particle image velocimetry (PIV), have been crucial in analyzing jet dynamics in complex scenarios, including turbulent and non-Newtonian flows (Philippe et al. Reference Philippe, Raufaste, Kurowski and Petitjeans2005; Pickering et al. Reference Pickering, Rigas, Schmidt, Sipp and Colonius2021; Usta et al. Reference Usta, Ahmad, Pathikonda, Khan, Gillis, Ranjan and Aidun2023). PIV studies, in particular, have provided detailed insights into jet instabilities, mixing, and unsteady behaviours in various media (Davies et al. Reference Davies, Fisher and Barratt1963; Dombrowski et al. Reference Dombrowski, Lewellyn, Pesci, Restrepo, Kessler and Goldstein2005; Vessaire et al. Reference Vessaire, Varas, Joubaud, Volk, Bourgoin and Vidal2020; Gauding et al. Reference Gauding, Bode, Brahami, Varea and Danaila2021; Hassanzadeh et al. Reference Hassanzadeh, Frigaard and Taghavi2023). Recent research, albeit focused on Newtonian fluids, also shows that fluid viscosity significantly influences jet penetration dynamics (Guyot et al. Reference Guyot, Cartellier and Matas2020). The dynamics of single-fluid viscoelastic jets have long been studied (Hosokawa et al. Reference Hosokawa, Kamamoto, Watanabe, Kusuno, Kobayashi and Tagawa2023), but research on fast jets in viscoplastic fluids – defined by yield stress and complex rheology – is still nascent and rapidly advancing (Jalaal et al. Reference Jalaal, Schaarsberg, Visser and Lohse2019). However, there is currently no model in the literature that predicts jet penetration dynamics into soft non-Newtonian materials, in particular viscoplastic ones. In this context, we conduct jet flow experiments in this study, injecting a Newtonian fluid into a soft viscoplastic gel (Balmforth et al. Reference Balmforth, Frigaard and Ovarlez2014; Thompson & Soares Reference Thompson and Soares2016), identifying the flow regimes, and developing an experimentally informed, simplified, semi-analytical model to estimate jet penetration depth based on PIV data.

2. Experimental setting

The jet was generated using a gear pump (ISMATEC MCP-Z Standard, 1 % accuracy) to inject fluid vertically through a centrally positioned, long cylindrical nozzle (diameter, $\hat {D}$ , of $0.432$ mm and length of 0.0508 m) into a transparent rectangular tank (20 $\times$ 10 $\times$ 10 cm $^3$ ); see figure 1. In this study, dimensionless quantities are hatless to distinguish them from the dimensional hatted quantities. The Newtonian jet fluid (dyed deionised water) was injected into a soft viscoplastic gel (transparent Carbopol solution (Carbomer 940), Making Cosmetics Co.); both were miscible and had a density of $\hat {\rho } \approx 997$ kg m $^{-3}$ , measured using a high-precision density metre (Anton Paar, DMA 35N).

Figure 1. Schematic of experimental setup with camera imaging and PIV techniques.

The Carbopol solutions were assumed to follow the viscoplastic Herschel–Bulkley model (Balmforth et al. Reference Balmforth, Frigaard and Ovarlez2014):

(2.1) \begin{equation} \begin{cases} \hat {\tau }=\hat {\tau }_y+ \hat {\kappa }\hat {\dot {\gamma }}^n ,\hspace {5 mm} \hat {\tau }\gt \hat {\tau }_y,\\ \hat {\dot {\gamma }}=0,\hspace {16 mm} \hat {\tau }\leqslant \hat {\tau }_y, \end{cases} \end{equation}

confirmed via rheometry (DHR-3, TA Instruments). In Equation (2.1), $\hat {\tau }$ , $\hat {\dot {\gamma }}$ , $\hat {\tau }_y$ , $\hat {\kappa }$ , and $n$ represent the shear stress, shear rate, yield stress ( $0-4.3$ Pa), fluid consistency index ( $0.001-1.64$ Pa $\cdot$ s $^n$ ), and power-law index ( $0.4-1$ ), respectively. Accordingly, the effective viscosity of the ambient viscoplastic fluid is defined using the jet characteristic shear rate $ ( {\hat {V}_0}/{\hat {D}} )$ : $\hat{\mu}_{a} = {\hat \tau _y} ({{\hat V}_0}/{\hat D} )^{-1} + \hat \kappa ({{\hat V}_0}/{\hat D} )^{n - 1}$ , where the mean injection velocity ( $\hat {V}_0$ ) ranged from 0.9 m to 11 ms−1, resulting in $\hat {\mu}_a$ ranging from 0.001–0.022 Pa $\cdot$ s. The jet fluid viscosity ( $\hat {\mu}_j$ ) was 0.001 Pa $\cdot$ s.

Our backlit setup, featuring light-emitting diode arrays and a digital camera (Basler acA2040–90um), captured jet flow images at 25 frames per second, which were processed using in-house codes to determine the temporal jet penetration depth. A time-resolved tomographic PIV system (LaVision) (Buzzaccaro et al. Reference Buzzaccaro, Secchi and Piazza2013; Hassanzadeh et al. Reference Hassanzadeh, Frigaard and Taghavi2023) analyzed the velocity fields by seeding both jet and soft gel with polyamide tracer particles (60 $\unicode{x03BC}$ m diameter, 1030 kg m $^{-3}$ density). A high-speed pulsed Nd laser (532 nm, 30 mJ per pulse) created a 5 cm laser illumination volume, with images captured by four high-speed CMOS cameras (Phantom VEO-E 340L) with 60 mm lenses (Nikon Micro Nikkor) and synchronised by a PTU-X unit. The system was calibrated with a 3D calibration plate, and 3D voxel volumes were reconstructed from particle intensity data, with velocity fields extracted via 3D cross-correlation. PIV images were processed (using DaVis 10 software) on a supercomputer (Micro Logo).

The key dimensionless numbers governing the jet flow reduce to the Reynolds number ( $Re$ ):

(2.2) \begin{equation} Re = \frac {\hat {\rho } \hat {V}_0 \hat {D}}{\hat {\mu}_j}, \end{equation}

which ranges from 350–5000, and the effective viscosity ratio ( $m$ ), obtained by balancing the characteristic viscous stresses in the jet and ambient fluids:

(2.3) \begin{align} {\hat{\mu}_{a}} \left ( \frac {\hat V_0}{\hat D} \right ) \sim \hat \tau _y \left ( \frac {\hat V_0}{\hat D} \right )^{-1} + \hat{\kappa} \left ( \frac {\hat V_0}{\hat D} \right )^{n - 1} \Rightarrow m &= \frac {\hat{\mu}_{a}}{\hat{\mu}_{j}} = \frac {\hat \tau _y \left ( \frac {\hat V_0}{\hat D} \right )^{-1} + \hat \kappa \left ( \frac {\hat V_0}{\hat D} \right )^{n - 1}}{\hat{\mu} _{j}} \nonumber\\&\equiv Bn + \frac {\hat \kappa }{\hat{\mu}_{j}} \left ( \frac {\hat V_0}{\hat D} \right )^{n - 1}, \end{align}

where the modified Bingham number is defined as $Bn = ({\hat {\tau }_y}/{\hat {\mu}_j (\hat {V}_0 / \hat {D}))}$ , ranging from 0 to 2, and $m$ spans from 1 to 22. In other words, $m$ provides a measure of how the viscosity and yield stress effects of the ambient fluid influence the jet flow, balancing inertial and viscous forces through their Reynolds number ratios, i.e., $m \equiv {Re}/{Re^\dagger }$ , where $Re^\dagger = {\hat {\rho }\hat {V}_0\hat {D}}/{\hat {\mu}_a}$ defines the ambient fluid’s Reynolds number. Note that, as $n$ is already embedded in the definitions of both $Bn$ and $m$ , its influence as a separate parameter may be less significant. Thus, $m$ and $Re^\dagger$ mainly characterise the flow dynamics; nevertheless, they do not fully span the dimensionless space, but numerical computations, free from experimental constraints, can systematically explore their effects (Thompson & Soares Reference Thompson and Soares2016), a task for future work.

Figure 2. (a) Sequence of experimental snapshots of mixing (Re $\approx$ 1250, m $\approx$ 3), jellyfish (Re $\approx$ 1600, m $\approx$ 4), and fingering (Re $\approx$ 1000, m $\approx$ 13) regimes. Snapshots at $\hat {t} = 0.29, 0.80, 1.42, 2.65$ , and $3.18$ seconds (mixing regime, supplementary video 1); $\hat {t} = 0.63, 3.26, 5.04, 7.05$ , and $9.03$ seconds (jellyfish regime, supplementary video 2); and $\hat {t} = 0.91, 3.46, 11.91, 18.85$ , and $25.79$ seconds (fingering regime, supplementary video 3). (b) Regime classification in $Re- ({Re}/{m})$ plane, showing mixing, jellyfish, and fingering regimes, with dashed (3.1) and solid (3.2) line transition boundaries. Triangle-square and square-circle symbols mark transitions between fingering-jellyfish and jellyfish-mixing regimes, respectively. Pink dotted line represents an alternative relation using a third-order expansion of (3.1), given by $Re_c^{\textit{mixing} \to \textit{jellyfish}} = ({Re}/{m}) + ({1}/{120}) ( ({Re}/{m})^2) - ({4}/{10^6}) (({Re}/{m})^3)$ .

3. Experimental results

In a typical experiment (figure 2 a), dyed jet fluid is injected from a nozzle into transparent viscoplastic fluid, penetrating into it as the jet advances. The maximum axial distance reached at a given time is defined as the jet penetration depth ( $\hat {L}_p$ ). Results are presented in dimensionless form using $\hat {D}$ for lengths, $\hat {V}_0$ for velocities, and ${\hat {D}}/{\hat {V}_0}$ for times, unless otherwise stated. Our results reveal the existence of three regimes: mixing, jellyfish, and fingering. The upper row of figure 2(a) shows the mixing regime, which occurs at high $Re$ and low $m$ , where significant mixing between the jet and soft gel is observed, along with an initial stable region. At higher $m$ and lower $Re$ , the middle row illustrates the jellyfish regime, i.e., a newly identified flow state, reported for the first time in viscoplastic fluids, and characterised by a vortex ring around the jet tip caused by instabilities. This vortex grows, and the jet radius expands transversely due to the higher $m$ . With further increases in $m$ and decreases in $Re$ , the lower row shows the fingering regime, where the jet fluid initially penetrates the yield-stress fluid before becoming trapped (Hassanzadeh et al. Reference Hassanzadeh, Frigaard and Taghavi2023), eventually forming evolving fingers. Moreover, our PIV analysis reveals fluctuations within each regime, with an average fluctuation intensity – defined as the ratio of turbulent to total kinetic energy – at $y \approx 40$ of approximately 60 %, 30 %, and 55 % for the mixing, jellyfish, and fingering regimes, respectively.

As shown in figure 2(b), the three morphological regimes – mixing, jellyfish, and fingering – can be classified using $Re$ and $m$ . Here, the transition from mixing (at high $Re^\dagger$ ) to jellyfish, and then to fingering (as $Re^\dagger$ decreases), is influenced by a combination of inertial and effective viscous forces. The critical transition between the mixing and jellyfish regimes is given by:

(3.1) \begin{equation} Re_c^{\textit {mixing} \to \textit {jellyfish}} = 200\;m \left (m-1\right ), \;\;\; 1\lesssim m\lesssim 22,\,\,\,O(10^2)\lesssim Re \lesssim O(10^4). \end{equation}

While the simplified relation in (3.1) provides a convenient approximation, a more precise transition can be obtained through higher-order expansions, as shown by the dotted line ( $Re_c^{\textit {mixing} \to \textrm {jellyfish}} = ({Re}/{m}) + ({1}/{120}) ( ({Re}/{m})^2) - ({4}/{10^6}) ( ({Re}/{m})^3)$ ) in figure 2(b).

The fingering regime emerges as $m$ increases, particularly with higher yield stress in the soft gel. The critical transition Reynolds number between the jellyfish and fingering regime is given by:

(3.2) \begin{equation} Re_c^{\kern1pt \textit {jellyfish} \to \textit {fingering}} = 15\;m \left (m-1\right ), \;\;\; 1\lesssim m\lesssim 22,\,\,\,O(10^2)\lesssim Re \lesssim O(10^4). \end{equation}

Figure 3. Modelling framework and results: (a) Newtonian jet injection into a viscoplastic fluid, showing jet centreline (dashed-dot), coordinates, and penetration depth ( $\hat {L}_p$ ) at $Re \approx 1000$ and $m \approx 4.5$ . Normalised axial velocity (b), axial Reynolds stress (c), and radial-axial Reynolds stress (d) versus $\eta$ at different axial distances, with brighter symbols indicating greater distances ( $30 \lesssim y \lesssim 100$ ). Each row corresponds to mixing, jellyfish, and fingering regimes (left to right). Fitted velocity curves (dashed-dotted) are $\cosh (\eta )^{-1.539}$ (mixing), $\cosh (\eta )^{-1.300}$ (jellyfish), and $\cosh (\eta )^{-0.864}$ (fingering), consistent with (Pope Reference Pope2000). Fitted axial Reynolds stress curves (dashed) are $\cosh (\eta )^{-2.226}$ , $\cosh (\eta )^{-1.594}$ , and $\cosh (\eta )^{-0.623}$ , and fitted radial-axial Reynolds stress curves (solid) are $0.22\sinh (\eta )\cosh (\eta )^{-2.5}$ , $0.12\sinh (\eta )\cosh (\eta )^{-3}$ , and $0.2\sinh (\eta )\cosh (\eta )^{-2.6}$ for the respective regimes. Mean squared errors between fitted and measured velocity profiles are $0.38\,\%$ (mixing), $0.08\,\%$ (jellyfish), $4.47\,\%$ (fingering), with Reynolds stress errors in a comparable range. (e) $S(y)$ , versus $y$ , with fitted curves $0.0039e^{-0.018y}$ (solid), $0.0044e^{-0.043y}$ (dashed), and $0.0031e^{-0.045y}$ (dotted).

4. Model development and comparison with experiments

We develop an experimentally guided, semi-analytical model to estimate the jet penetration depth into the soft gel over time. The model is based on dimensionless motion equations for momentum and continuity in a cylindrical coordinate system (with $(r, \theta , y)$ denoting radial, tangential, and axial directions; see figure 3 a):

(4.1) \begin{align} Re^\dagger \left ( {\frac {{\partial \mathbf{{v}}}}{{\partial t}} + \mathbf{{v}} \cdot \nabla \mathbf{{v}}} \right ) &= - \nabla p + \nabla \cdot \boldsymbol{\tau }, \end{align}
(4.2) \begin{align} \nabla \cdot \mathbf{v}&=0, \end{align}

where $\mathbf{{v}} = (v_r, v_\theta , v_y )$ represents the velocity field, $p$ the pressure, and $\boldsymbol{\tau }$ the stress tensor. Using $Re^\dagger$ in (4.1) simplifies the analysis by encapsulating the viscoplastic fluid’s rheology into an effective viscosity, reflecting the dominant inertial-to-viscous force ratio. This approach captures the gel’s resistance properties, governed by its yield stress and characteristic shear rate, which critically influence the jet penetration depth and flow morphology. While it aligns with the observed flow transitions (mixing, jellyfish, and fingering, as shown in figure 2 b), it neglects local variations in viscosity, detailed mixing mechanisms, and secondary flow dynamics, such as vortex and finger formation.

We assume that the jet dynamics develop over a thin, elongated layer with thickness $\zeta$ and length $\ell$ . This allows us to define an arbitrary small aspect ratio $\varepsilon = \zeta /\ell$ , and we rescale our variables as follows:

(4.3) \begin{equation} \varepsilon y = {y^*}\,\,, \varepsilon t = {t^*}\,\,,\varepsilon p = {p^ * }\,\,,{v_r} = \varepsilon v_r^*\,\,,{v_\theta } = \varepsilon v_\theta ^*. \end{equation}

Therefore, the motion equations in the leading order can be found as:

(4.4) \begin{align} &\qquad\qquad\qquad\qquad O\big ( {R{e^\dagger }{\varepsilon ^3}} \big ) = - \frac {{\partial {p^*}}}{{\partial r}} + O\big ( {{\varepsilon ^2}} \big ), \end{align}
(4.5) \begin{align} &\qquad\qquad\qquad\qquad O\big ( {R{e^\dagger }{\varepsilon ^3}} \big ) = - \frac {{\partial {p^*}}}{{\partial \theta }} + O\big ( {{\varepsilon ^2}} \big ), \end{align}
(4.6) \begin{align} &R{e^\dagger }\varepsilon \left ( {\frac {{\partial {v_y}}}{{\partial {t^*}}} + v_r^*\frac {{\partial {v_y}}}{{\partial r}} + \frac {{v_\theta ^*}}{r}\frac {{\partial {v_y}}}{{\partial \theta }} + {v_y}\frac {{\partial {v_y}}}{{\partial {y^*}}}} \right ) = - \frac {{\partial {p^*}}}{{\partial {y^*}}} + \frac {1}{r}\frac {\partial }{{\partial r}}\left ( {r\frac {\partial v_y }{{\partial r}}} \right ) \nonumber\\&\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\,\, + \frac {1}{{{r^2}}}\frac {{{\partial ^2}{v_y}}}{{\partial {\theta ^2}}} + O\big ( {{\varepsilon ^2}} \big ), \end{align}
(4.7) \begin{align} &\qquad\qquad\qquad\qquad \frac {1}{r}\frac {{\partial (rv_r^*)}}{{\partial r}} + \frac {1}{r}\frac {{\partial v_\theta ^*}}{{\partial \theta }} + \frac {{\partial {v_y}}}{{\partial {y^*}}} = 0, \end{align}

which, considering $\varepsilon \ll 1$ with a fixed $\varepsilon Re^\dagger$ (implying $Re^\dagger \propto {1}/{\varepsilon } \gg 1$ ), leads to the following axial momentum equation:

(4.8) \begin{equation} R{e^\dagger }\varepsilon \left ( {\frac {{\partial {v_y}}}{{\partial {t^*}}} + v_r^*\frac {{\partial {v_y}}}{{\partial r}} + \frac {{v_\theta ^*}}{r}\frac {{\partial {v_y}}}{{\partial \theta }} + {v_y}\frac {{\partial {v_y}}}{{\partial {y^*}}}} \right ) = - \frac {{\partial {p^*}}}{{\partial {y^*}}} + \frac {1}{r}\frac {\partial }{{\partial r}}\left ( {r\frac {\partial v_y }{{\partial r}}} \right ) + \frac {1}{{{r^2}}}\frac {{{\partial ^2}{v_y}}}{{\partial {\theta ^2}}}, \end{equation}

where $p^*=p^*(y)$ only. Now, it can be transformed back to the original variable scaling to reach:

(4.9) \begin{equation} Re^\dagger \left ({\frac {{\partial v_y}}{{\partial {t}}} + {v_r}\frac {{\partial v_y}}{{\partial r}} + \frac {{v_\theta }}{r}\frac {{\partial v_y}}{{\partial \theta }} + v_y\frac {{\partial v_y}}{{\partial {y}}}} \right ){\textrm { }} = \frac {1}{r}\frac {\partial }{{\partial r}}\left ( {r\frac {\partial v_y }{{\partial r}}} \right ) + \frac {1}{{{r^2}}}\frac {{{\partial ^2}v_y}}{{\partial {\theta ^2}}}, \end{equation}

in which the pressure gradient term is neglected, following the fact that pressure does not depend on $r$ and pressure is constant at larger radial distances from the jet centreline, in line with the literature (Guimarães et al. Reference Guimarães, Pinho and da Silva2023). Also, this is due to the dominance of inertial and viscous forces, as the gel’s yield stress and viscosity primarily govern penetration depth and morphology.

We now derive the Reynolds-averaged form of (4.9), by decomposing the velocity into its mean (bar notation) and fluctuating (prime notation) components in the form of:

(4.10) \begin{equation} \mathbf{{v}}\left ( {t,r,\theta ,y} \right ) = {\overline {\mathbf{v}}}\left ( {r,y} \right ) + \mathbf{{v}'}\left ( {t,r,\theta ,y} \right ), \end{equation}

where the overbar denotes circumferential and ensemble averaging, resulting in the following leading order motion equations:

(4.11) \begin{align} &{\overline v_r}\frac {{\partial {{\overline v}_y}}}{{\partial r}} + {\overline v_y}\frac {{\partial {{\overline v}_y}}}{{\partial y}} = \frac {1}{{Re^\dagger }}\left ( \frac {1}{r}\frac {\partial }{{\partial r}}\left ( {r\frac {\partial \overline {v}_y }{{\partial r}}} \right ) \right ) - \left [ {\frac {{\partial \overline {{{v'}_y}{{v'}_y}} }}{{\partial y}} + \frac {1}{r}\frac {\partial \big (r \overline {{{v'}_r}{{v'}_y}}\big )}{\partial r} } \right ], \end{align}
(4.12) \begin{align} &\qquad\qquad\qquad\qquad\qquad \frac {1}{r}\frac {{\partial \left (r{{\overline v}_r}\right )}}{{\partial r}} + \frac {{\partial {{\overline v}_y}}}{{\partial y}}{\textrm { }} = 0, \end{align}

where $\overline {{v'}_y {v'}_y}$ and $\overline { {v'}_r {v'}_y}$ are dominant Reynolds stresses (confirmed by PIV), and the flow is assumed to be statistically steady.

As shown in figure 3(bd), the profiles of ${\overline {v}_y}/{(\overline {v}_y)_c}$ , $ {\overline {v'_y v'_y}}/{(\overline {v'_y v'_y})_c}$ and ${\overline {v'_r v'_y}}/{(\overline {v'_y v'_y})_c}$ (with the subscript $c$ denoting the centreline) exhibit an approximate scaling behaviour, with the scaling variable defined as:

(4.13) \begin{equation} \eta = \frac {r}{r_{1/2}}, \end{equation}

in which $r_{1/2}(y)$ represents the radial distance from the centreline where the mean velocity drops to half of its centreline value, characterising the jet’s lateral spread (Pope Reference Pope2000; Kuhn et al. Reference Kuhn, Soria and Oberleithner2021). Note that, although dispersions and deviations from a universal scaling collapse are observed, particularly in the fingering regime (figure 3 e), where localised stress effects and flow confinement introduce variations in the velocity profile, our assumed approximate scaling captures the dominant trends and serves as a practical approximation for a simplified model. Thus, a degree of dependence on $\eta$ allows the mean velocity, $\overline {v}_y$ , to be approximately expressed via a functional decomposition:

(4.14) \begin{equation} \overline {v}_y = U(y) \Psi (\eta ), \end{equation}

where $\Psi (\eta )$ is found for a representative experiment in each regime via PIV data fitting (figure 3 b). Similarly, for the dominant Reynolds stress terms, we write:

(4.15) \begin{equation} \overline {{v}'_y{v}'_y} = {S}({y})\Phi ({\eta }), \,\,\,\,\, \overline {{v}'_r{v}'_y} = {S}({y})\chi ({\eta }), \end{equation}

where ${S}({y})$ , $\Phi ({\eta })$ , and $\chi (\eta )$ are found for each regime via PIV data fitting (figure 3 c–e).

Multiplying the Equation (4.11) by $\overline {v}_y$ and integrating it over volume gives:

(4.16) \begin{align} &\int _0^y {\int _0^\infty {\left ( {{{\overline v }_r}{{\overline v }_y}\frac {{\partial {{\overline v }_y}}}{{\partial r}} + \overline v _y^2\frac {{\partial {{\overline v }_y}}}{{\partial \mathsf{y}}}} \right )} } 2\pi r{\textrm d}r{\textrm d}\mathsf{y} = \nonumber \\ &\int _0^y {\int _0^\infty {\left [ {\frac {1}{{Re^\dagger }}\left ( {{{\overline v }_y}\frac {{{\partial ^2}{{\overline v }_y}}}{{\partial {r^2}}} + \frac {{{{\overline v }_y}}}{r}\frac {{\partial {{\overline v }_y}}}{{\partial r}}} \right ) - {{\overline v }_y}\frac {{\partial \overline {{{v}^{\prime}_y}{{v}^{\prime}_y}} }}{{\partial \mathsf{y}}}} - \frac {{\overline v }_y}{r}\frac {\partial \left (r \overline {{{v}^{\prime}_r}{{v}^{\prime}_y}}\right )}{\partial r}\right ]}} 2\pi r{\textrm d}r{\textrm d}\mathsf{y}, \end{align}

where $\mathsf{y}$ the dummy variable of integration. Using the continuity equation, the left-hand side of (4.16) is reformulated. The term $\overline {v}_y ({\partial ^2 \overline {v}_y}/{\partial r^2})$ is decomposed as ${1}/{2} ({\partial ^2 (\overline {v}_y^2)}/\def\luminalatbreak{} {\partial r^2}) - (({\partial \overline {v}_y}/{\partial r} )^2)$ ; the first component, contributing only $\sim$ 7 % across all regimes based on the PIV results, is neglected, yielding a simplified expression after some algebra:

(4.17) \begin{align} \frac {3}{2}\int _0^y \def\negativespace{}\def\negativespace{}{\int _0^\infty {\bar v_y^2} } \frac {{\partial {{\bar v}_y}}}{{\partial \mathsf{y}}}2\pi r\textrm{d}r\textrm{d}\mathsf{y} & = \frac {1}{{Re^{\dagger }}}\left [ {\int _0^y \def\negativespace{}\def\negativespace{}{\int _0^\infty {{{\bar v}_y}} } \frac {{\partial {{\bar v}_y}}}{{\partial r}}2\pi \textrm{d}r\textrm{d}\mathsf{y} - \int _0^y \def\negativespace{}\def\negativespace{}{\int _0^\infty \def\negativespace{}{{{\left (\def\negativespace{} {\frac {{\partial {{\bar v}_y}}}{{\partial r}}}\def\negativespace{} \right )}^2}} } 2\pi r\textrm{d}r\textrm{d}\mathsf{y}} \def\negativespace{}\right ] \notag \\&\quad - \int _0^y {\int _0^\infty {{{\bar v}_y} {\frac {{\partial \overline {{{v'}_y}{{v'}_y}} }}{{\partial \mathsf{y}}}}} } 2\pi r\textrm{d}r\textrm{d}\mathsf{y} \nonumber \\ & \quad - \int _0^y {\int _0^\infty \frac {{\overline v }_y}{r}\frac {\partial \left (r \overline {{{v'}_r}{{v'}_y}}\right )}{\partial r} } 2\pi r\textrm{d}r\textrm{d}\mathsf{y}. \end{align}

Subsequently, using the approximate functional decomposition approach ((4.13), (4.14), and (4.15)), along with integration by parts gives:

(4.18) \begin{align} &\frac {1}{2}\left [ {{U^3} {r_{1/2}^2} \int _0^\infty {{\Psi ^3}} 2\pi \eta {\textrm d}\eta -U^3(0) {r_{1/2}^2}(0) \int _0^\infty {{\Psi ^3}} 2\pi \eta {\textrm d}\eta } \right ] \notag \\ & = \frac {1}{{Re^{\dagger }}}\left [ {\int _0^y {\int _0^\infty {{U^2}} } \Psi \Psi '2\pi \textrm{d}\eta \textrm{d}\mathsf{y} - \int _0^y {\int _0^\infty {{U^2}} } {{\Psi '}^2}2\pi \eta {\textrm d}\eta {\textrm d}\mathsf{y}} \right ] \notag \\ &\quad - \int _0^y {\int _0^\infty U {r_{1/2}^2}} \Psi S' \Phi 2\pi \eta \textrm{d}\eta \textrm{d}\mathsf{y} - \int _0^y {\int _0^\infty U {r_{1/2}} } \Psi S \left (\chi + \chi ' \eta \right )2\pi \textrm{d}\eta \textrm{d}\mathsf{y}, \end{align}

Now, multiplying and dividing (4.18) by $\int _{0}^{\infty } \Psi 2 \pi {\eta } {\textrm d}{\eta }$ , isolating the jet flux ( $Q$ ), and applying the boundary condition ( $U(0)=1$ ), we eventually arrive at:

(4.19) \begin{equation} \frac {1}{2}\alpha Q \left ( {{U^2} - 1} \right ) = -\frac {1}{{Re^\dagger }}\int _0^y {{U^2}\beta \textrm{d}\mathsf{y}} - \int _0^y {QS'\gamma \textrm{d}\mathsf{y}}- \int _0^y {\sqrt {QU}S \zeta \textrm{d}\mathsf{y}}, \end{equation}

where the prime denotes the derivative and the model-derived parameters $\alpha$ , $\beta$ , $\gamma$ , and $\zeta$ are functions of the jet velocity and Reynolds stress profiles:

(4.20) \begin{equation} \left \{ {\begin{array}{*{20}{l}} {\alpha = \dfrac {{\int _0^\infty {{\Psi ^3}} \eta {\textrm d}\eta }}{{\int _0^\infty \Psi \eta {\textrm d}\eta }},}\\[6pt] {\beta = 2\pi \int _0^\infty {\left ( {\eta {{\Psi '}^2} - \Psi \Psi '} \right ){\textrm d}\eta } ,}\\[6pt] {\gamma = \dfrac {{\int _0^\infty \Psi \Phi \eta {\textrm d}\eta }}{{\int _0^\infty \Psi \eta {\textrm d}\eta }},}\\[6pt] {\zeta = \sqrt {2\pi } \dfrac {{\int _0^\infty \Psi \left (\chi + \chi ' \eta \right ) {\textrm d}\eta }}{\sqrt {\int _0^\infty \Psi \eta {\textrm d}\eta }}.} \end{array}} \right . \end{equation}

In (4.19), $Q= {U} r^2_{1/2} \int _{0}^{\infty } 2 \pi \eta \Psi {\textrm d}\eta = Q_0 + Q_e$ , where $Q_0$ is the injection flux and $Q_e$ accounts for entrainment, with PIV analysis showing $Q_e$ contributes 23–30 % across regimes. However, for simplicity and analytical tractability, we assume $Q \approx Q_0=\pi /4$ (constant), acknowledging a jet momentum underestimation. Also, since the last term in (4.19) (radial-axial Reynolds stress contribution) has a secondary effect on axial momentum transport, with PIV measurements showing its impact is 4–15 % of dominant terms across regimes, we omit it for simplicity.

Taking the derivative of (4.19) with respect to $y$ , integrating, and applying the boundary condition $U(0) = 1$ yield an analytical expression for the jet velocity:

(4.21) \begin{equation} U(y) = {\left ( {1 - \frac {{2\gamma }}{\alpha }\int _0^y {\exp \left ( {\frac {{2\beta \mathsf{y}}}{{\alpha Q_0Re^\dagger }}} \right )S'{\textrm d}\mathsf{y}} } \right )^{\frac {1}{2}}}\exp \left ( { - \frac {{\beta y}}{{\alpha Q_0Re^\dagger }}} \right ), \end{equation}

which is then integrated (via ${U} = ({{\textrm d}{L_p}}/{{\textrm d}{t}})$ ) to calculate the jet penetration depth ( $L_p$ ) as a function of time. We can, thus, derive the variation in the jet penetration depth over time based on the key parameters such as the jet velocity profiles, dominant Reynolds stress, injection velocity, and flow rate. Figure 3(b) show the jet velocity profiles used in our model, with $\alpha$ values of 0.25, 0.24, and 0.20, and $\beta$ values of 5.8, 5.7, and 5.4 for the mixing, jellyfish, and fingering regimes, respectively. The Reynolds stress profiles are detailed in figures 3(c) and 3(d), yielding $\gamma$ values of 0.32, 0.35, and 0.41, for the respective regimes. According to our PIV analysis for different cases, $\alpha$ , $\beta$ , and $\gamma$ vary by up to $\pm$ 8 %.

The profiles for $\Psi (\eta )$ , $\Phi (\eta )$ , $S(y)$ are extracted from PIV data and assumed constant within each flow regime (mixing, jellyfish, fingering). Model-derived parameters $\alpha$ , $\beta$ , and $\gamma$ encapsulate the axial velocity dynamics and interactions with Reynolds stresses, while $S(y)$ reflects axial decay, energy dissipation, and turbulence damping. These parameters, assumed applicable without re-fitting for each experiment, enable efficient estimation of $L_p$ over time. This provides a robust, validated approach, while simplifying analysis and enhancing scalability across experimental scenarios.

Figure 4. (a) Jet penetration depth over time for experiments (symbols) and model (lines) across three flow regimes: mixing ( $Re \approx 1300$ , $m\approx 1$ , blue), jellyfish ( $Re \approx 1300$ , $m\approx 4$ , red), and fingering ( $Re \approx 1300$ , $m\approx 11$ , green). (b) Model versus experimental results of $L_p$ at $t\approx O(10^3)$ , both multiplied by $Re$ to illustrate the data spread. The solid line shows $\hat {L}^{\textit{Model}}_p = \hat {L}^{\textit{Experiment}}_p$ . Data points’ face colour, size, and edge width indicate $Re$ , $m$ , and $Bn$ , with circles, squares, and triangles for mixing, jellyfish, and fingering regimes. Inset shows model outputs vs. experimental results of $\hat {L}_{p}$ (dimensional) from $\hat {t} = 0.4$ s to the experiment end, with black/red edges for start/end points and dashed lines for time variation.

Figure 4(a) compares the model’s outputs with experimental data across the three flow regimes, applying the obtained values of $\alpha$ , $\beta$ , and $\gamma$ to all cases with similar flow regimes. The model demonstrates reasonable estimations of the jet penetration depth over time, even in the challenging fingering regime with high yield stress. In the mixing regime, the model closely aligns with experimental values initially, with mid-time deviations converging later as energy dissipation is accounted for. In the jellyfish regime, the model initially overestimates the penetration depth but eventually aligns with experimental trends by the end of the injection period. The inset in figure 4(b) shows the variation of $\hat {L}_p$ between the model and experimental data, with initial overestimations likely due to unaccounted losses near the nozzle exit. This discrepancy is somewhat corrected over time, although occasional underestimations can also occur. The model’s accuracy generally improves over time across all flow regimes, although higher $Re$ values result in greater deviations, possibly due to underestimated Reynolds stresses. For a larger dataset, the main panel in figure 4(b) demonstrates the model’s overall estimative capability but also indicates increasing deviations at higher $Re$ .

5. Conclusions

A Newtonian jet penetrating a soft viscoplastic gel was studied across viscosity ratios from 1 to 22 and Reynolds numbers between 350 and 5000. Our experiments identified three distinct responses of the viscoplastic fluid: a mixing regime dominated by turbulence, a jellyfish regime with vortex formation and radial jet expansion, and a fingering regime where the jet becomes confined and forms localised fingers. To estimate the penetration depth over time, we developed an experimentally informed semi-analytical model incorporating key dimensionless parameters, such as effective viscosity and Reynolds stresses, and leveraging an approximate scaling approach. The model demonstrates reasonable estimations across all regimes, providing a robust framework for understanding jet interactions with soft viscoplastic gels. However, it does not account for long-term effects, such as dominant mixing or extensive finger formation, where nonlinearities and instabilities become significant. Future work should address these complexities to extend the model’s applicability to high-yield-stress environments and more intricate scenarios. Finally, we investigated the problem in $Re-m$ space, and future work incorporating other dimensionless groups, such as the Bingham number and power-law index, could enhance understanding of our jet flows.

Supplementary movies

Supplementary movies are available at https://doi.org/10.1017/jfm.2025.352.

Acknowledgements

SPM acknowledges the ESSOR PhD scholarship, and SMT thanks the Humboldt Research Fellowship for Experienced Researchers programme.

Funding

We acknowledge financial support from the NFRF (GF141041); CFI (GF130120, GQ130119, GF525075); CRC on Modelling Complex Flows (CG125810); NSERC Discovery (CG109154); NSERC RTI (CG132931); and NSERC Alliance International (CG141435).

Declaration of interests

The authors report no conflict of interest.

References

Bailey, S.J. & Appel, E.A. 2024 Sticky gels designed for tissue-healing therapies and diagnostics. Nature 625 (7995), 455457.CrossRefGoogle ScholarPubMed
Balmforth, N.J., Frigaard, I.A. & Ovarlez, G. 2014 Yielding to stress: recent developments in viscoplastic fluid mechanics. Annu. Rev. Fluid Mech. 46 (1), 121146.CrossRefGoogle Scholar
Bantawa, M., Keshavarz, B., Geri, M., Bouzid, M., Divoux, T., McKinley, G.H. & Del Gado, E. 2023 The hidden hierarchical nature of soft particulate gels. Nat. Phys. 19 (8), 11781184.CrossRefGoogle Scholar
Bazazi, P., Stone, H.A. & Hejazi, S.H. 2022 Spongy all-in-liquid materials by in-situ formation of emulsions at oil-water interfaces. Nat. Commun. 13 (1), 4162.CrossRefGoogle ScholarPubMed
Bonn, D., Denn, M.M., Berthier, L., Divoux, T. & Manneville, S. 2017 Yield stress materials in soft condensed matter. Rev. Mod. Phys. 89 (3), 035005.CrossRefGoogle Scholar
Buzzaccaro, S., Secchi, E. & Piazza, R. 2013 Ghost particle velocimetry: accurate 3D flow visualization using standard lab equipment. Phys. Rev. Lett. 111 (4), 048101.CrossRefGoogle ScholarPubMed
Cianchetti, M., Laschi, C., Menciassi, A. & Dario, P. 2018 Biomedical applications of soft robotics. Nat. Rev. Mater. 3 (6), 143153.CrossRefGoogle Scholar
Davies, P.O.A.L., Fisher, M.J. & Barratt, M.J. 1963 The characteristics of the turbulence in the mixing region of a round jet. J. Fluid Mech. 15 (3), 337367.CrossRefGoogle Scholar
Dombrowski, C., Lewellyn, B., Pesci, A.I., Restrepo, J.M., Kessler, J.O. & Goldstein, R.E. 2005 Coiling, entrainment, and hydrodynamic coupling of decelerated fluid jets. Phys. Rev. Lett. 95 (18), 184501.CrossRefGoogle ScholarPubMed
Gauding, M., Bode, M., Brahami, Y., Varea, É. & Danaila, L. 2021 Self-similarity of turbulent jet flows with internal and external intermittency. J. Fluid Mech. 919, A41.CrossRefGoogle Scholar
Guimarães, M.C., Pinho, F.T., da Silva, C.B. 2023 Viscoelastic jet instabilities studied by direct numerical simulations. Phys. Rev. Fluids 8 (10), 103301.CrossRefGoogle Scholar
Guyot, G., Cartellier, A. & Matas, J.P. 2020 Penetration depth of a plunging jet: from microjets to cascades. Phys. Rev. Lett. 124 (19), 194503.CrossRefGoogle ScholarPubMed
Hassanzadeh, H., Frigaard, I.A. & Taghavi, S.M. 2023 Neutrally buoyant miscible jets into viscoplastic ambient fluids. J. Non-Newtonian Fluid Mech. 320, 105107.CrossRefGoogle Scholar
Hosokawa, A., Kamamoto, K., Watanabe, H., Kusuno, H., Kobayashi, K.U. & Tagawa, Y. 2023 A phase diagram of the pinch-off-type behavior of impulsively-induced viscoelastic liquid jets. arXiv preprint arXiv: 2309.01364. (Last accessed: March 19, 2025).Google Scholar
Jalaal, M., Schaarsberg, M.K., Visser, C.W. & Lohse, D. 2019 Laser-induced forward transfer of viscoplastic fluids. J. Fluid Mech. 880, 497513.CrossRefGoogle Scholar
Jones, A.T., Shen, X., Walter, K.L., LaBranche, C.C., Wyatt, L.S., Tomaras, G.D., Montefiori, D.C., Moss, B., Barouch, D.H. & Clements, J.D. 2019 HIV-1 vaccination by needle-free oral injection induces strong mucosal immunity and protects against SHIV challenge. Nat. Commun. 10 (1), 798.CrossRefGoogle ScholarPubMed
Kuhn, P., Soria, J. & Oberleithner, K. 2021 Linear modelling of self-similar jet turbulence. J. Fluid Mech. 919, A7.CrossRefGoogle Scholar
Li, X. & Gong, J.P. 2024 Design principles for strong and tough hydrogels. Nat. Rev. Mater. 9 (6), 119.CrossRefGoogle Scholar
Philippe, P., Raufaste, C., Kurowski, P. & Petitjeans, P. 2005 Penetration of a negatively buoyant jet in a miscible liquid. Phys. Fluids 17 (5), 053601-1–053601-10.CrossRefGoogle Scholar
Pickering, E., Rigas, G., Schmidt, O.T., Sipp, D. & Colonius, T. 2021 Optimal eddy viscosity for resolvent-based models of coherent structures in turbulent jets. J. Fluid Mech. 917, A29.CrossRefGoogle Scholar
Pope, S.B. 2000 Turbulent Flows. Cambridge University Press.Google Scholar
Rayleigh, L. 1878 On the instability of jets. Proc. Lond. Math. Soc. 1 (1), 413.CrossRefGoogle Scholar
Reynolds, A.J. 1962 Observations of a liquid-into-liquid jet. J. Fluid Mech. 14 (4), 552556.CrossRefGoogle Scholar
Schoppink, J. & Rivas, D.F. 2022 Jet injectors: perspectives for small volume delivery with lasers. Adv. Drug Deliver. Rev. 182, 114109.CrossRefGoogle ScholarPubMed
Stokes, G.G. 1851 On the effect of the internal friction of fluids on the motion of pendulums. Trans. Cambridge Philos. Soc. 9, 8106.Google Scholar
Taheri, S., Bao, G., He, Z., Mohammadi, S., Ravanbakhsh, H., Lessard, L., Li, J. & Mongeau, L. 2022 Injectable, pore-forming, perfusable double-network hydrogels resilient to extreme biomechanical stimulations. Adv. Sci. 9 (2), 2102627.CrossRefGoogle ScholarPubMed
Thompson, R.L. & Soares, E.J. 2016 Viscoplastic dimensionless numbers. J. Non-Newtonian Fluid Mech. 238, 5764.CrossRefGoogle Scholar
Usta, M., Ahmad, M.R.C., Pathikonda, G., Khan, I., Gillis, P., Ranjan, D. & Aidun, C.K. 2023 Coaxial jets with disparate viscosity: mixing and laminarization characteristics. J. Fluid Mech. 955, A43.CrossRefGoogle Scholar
Vessaire, J., Varas, G., Joubaud, S., Volk, R., Bourgoin, M. & Vidal, V. 2020 Stability of a liquid jet impinging on confined saturated sand. Phys. Rev. Lett. 124 (22), 224502.CrossRefGoogle ScholarPubMed
da Vinci, L. 1510 Codex Leicester. Bill Gates Collection, historical document, Manuscript.Google Scholar
Xie, M., Shi, Y., Zhang, C., Ge, M., Zhang, J., Chen, Z., Fu, J., Xie, Z. & He, Y. 2022 In situ 3D bioprinting with bioconcrete bioink. Nat. Commun. 13 (1), 3597.CrossRefGoogle ScholarPubMed
Xie, X., Xu, Z., Yu, X., Jiang, H., Li, H. & Feng, W. 2023 Liquid-in-liquid printing of 3D and mechanically tunable conductive hydrogels. Nat. Commun. 14 (1), 4289.CrossRefGoogle ScholarPubMed
Figure 0

Figure 1. Schematic of experimental setup with camera imaging and PIV techniques.

Figure 1

Figure 2. (a) Sequence of experimental snapshots of mixing (Re$\approx$ 1250, m$\approx$ 3), jellyfish (Re$\approx$ 1600, m$\approx$ 4), and fingering (Re$\approx$ 1000, m$\approx$ 13) regimes. Snapshots at $\hat {t} = 0.29, 0.80, 1.42, 2.65$, and $3.18$ seconds (mixing regime, supplementary video 1); $\hat {t} = 0.63, 3.26, 5.04, 7.05$, and $9.03$ seconds (jellyfish regime, supplementary video 2); and $\hat {t} = 0.91, 3.46, 11.91, 18.85$, and $25.79$ seconds (fingering regime, supplementary video 3). (b) Regime classification in $Re- ({Re}/{m})$ plane, showing mixing, jellyfish, and fingering regimes, with dashed (3.1) and solid (3.2) line transition boundaries. Triangle-square and square-circle symbols mark transitions between fingering-jellyfish and jellyfish-mixing regimes, respectively. Pink dotted line represents an alternative relation using a third-order expansion of (3.1), given by $Re_c^{\textit{mixing} \to \textit{jellyfish}} = ({Re}/{m}) + ({1}/{120}) ( ({Re}/{m})^2) - ({4}/{10^6}) (({Re}/{m})^3)$.

Figure 2

Figure 3. Modelling framework and results: (a) Newtonian jet injection into a viscoplastic fluid, showing jet centreline (dashed-dot), coordinates, and penetration depth ($\hat {L}_p$) at $Re \approx 1000$ and $m \approx 4.5$. Normalised axial velocity (b), axial Reynolds stress (c), and radial-axial Reynolds stress (d) versus $\eta$ at different axial distances, with brighter symbols indicating greater distances ($30 \lesssim y \lesssim 100$). Each row corresponds to mixing, jellyfish, and fingering regimes (left to right). Fitted velocity curves (dashed-dotted) are $\cosh (\eta )^{-1.539}$ (mixing), $\cosh (\eta )^{-1.300}$ (jellyfish), and $\cosh (\eta )^{-0.864}$ (fingering), consistent with (Pope 2000). Fitted axial Reynolds stress curves (dashed) are $\cosh (\eta )^{-2.226}$, $\cosh (\eta )^{-1.594}$, and $\cosh (\eta )^{-0.623}$, and fitted radial-axial Reynolds stress curves (solid) are $0.22\sinh (\eta )\cosh (\eta )^{-2.5}$, $0.12\sinh (\eta )\cosh (\eta )^{-3}$, and $0.2\sinh (\eta )\cosh (\eta )^{-2.6}$ for the respective regimes. Mean squared errors between fitted and measured velocity profiles are $0.38\,\%$ (mixing), $0.08\,\%$ (jellyfish), $4.47\,\%$ (fingering), with Reynolds stress errors in a comparable range. (e) $S(y)$, versus $y$, with fitted curves $0.0039e^{-0.018y}$ (solid), $0.0044e^{-0.043y}$ (dashed), and $0.0031e^{-0.045y}$ (dotted).

Figure 3

Figure 4. (a) Jet penetration depth over time for experiments (symbols) and model (lines) across three flow regimes: mixing ($Re \approx 1300$, $m\approx 1$, blue), jellyfish ($Re \approx 1300$, $m\approx 4$, red), and fingering ($Re \approx 1300$, $m\approx 11$, green). (b) Model versus experimental results of $L_p$ at $t\approx O(10^3)$, both multiplied by $Re$ to illustrate the data spread. The solid line shows $\hat {L}^{\textit{Model}}_p = \hat {L}^{\textit{Experiment}}_p$. Data points’ face colour, size, and edge width indicate $Re$, $m$, and $Bn$, with circles, squares, and triangles for mixing, jellyfish, and fingering regimes. Inset shows model outputs vs. experimental results of $\hat {L}_{p}$ (dimensional) from $\hat {t} = 0.4$ s to the experiment end, with black/red edges for start/end points and dashed lines for time variation.

Supplementary material: File

Mousavi et al. supplementary material movie 1

Mixing flow regime example (Re ≈ 1250, m ≈ 3)
Download Mousavi et al. supplementary material movie 1(File)
File 4.5 MB
Supplementary material: File

Mousavi et al. supplementary material movie 2

Jellyfish flow regime example (Re ≈ 1600, m ≈ 4)
Download Mousavi et al. supplementary material movie 2(File)
File 25.6 MB
Supplementary material: File

Mousavi et al. supplementary material movie 3

Fingering flow regime example (Re ≈ 1000, m ≈ 13)
Download Mousavi et al. supplementary material movie 3(File)
File 18.5 MB