1. Introduction
In this paper we present an arbitrary Lagrangian–Eulerian (ALE) finite element method and open-source Julia code (Sahu Reference Sahu2024) to simulate the dynamics of curved and deforming lipid membranes. Our developments build on ALE theories of biological membranes where the surface parametrisation is independent of the in-plane flow of lipids (Hu, Zhang & W. Reference Hu, Zhang and W.2007; Torres-Sánchez et al. Reference Torres-Sánchez, Millán and Arroyo2019; Sahu et al. Reference Sahu, Omar and Mandadapu2020b ). In our simulations, we arbitrarily specify the dynamics of the discretised surface and avoid highly distorted mesh elements – all while not altering membrane dynamics. The utility of our implementation is demonstrated by analysing a biologically motivated scenario that is difficult to simulate with existing numerical methods: tether formation, elongation and subsequent translation.
Biological membranes are two-dimensional (2-D) materials, comprised of lipids and proteins, which make up the boundary of the cell and many of its internal organelles. The lipids and proteins flow in-plane as a 2-D fluid, while the membrane bends out-of-plane as an elastic shell. Such membranes play a fundamental role in many cellular processes, including endocytosis (Higgins & McMahon Reference Higgins and McMahon2002; McMahon & Boucrot Reference McMahon and Boucrot2011), cell migration (Lauffenburger & Horwitz Reference Lauffenburger and Horwitz1996) and tether network dynamics within the cell (Terasaki, Chen & Fujiwara Reference Terasaki, Chen and Fujiwara1986; Sciaky et al. Reference Sciaky, Presley, Smith, Zaal, Cole, Moreira, Terasaki, Siggia and Lippincott-Schwartz1997). Lipid membranes often undergo dramatic shape changes in which their in-plane and out-of-plane dynamics are coupled. Consequently, comprehensive models and advanced numerical techniques are needed to describe membrane behaviour. In the early 1970’s, the seminal contributions of Canham (Reference Canham1970), Helfrich (Reference Helfrich1973) and Evans (Reference Evans1974) – all of which can be viewed as extensions of Naghdi’s fundamental contributions to shell theory (Naghdi Reference Naghdi1973) – laid the foundation for theoretical developments (Evans & Hochmuth Reference Evans and Hochmuth1978; Waxman Reference Waxman1984; Zhong-can & Helfrich Reference Zhong-can and Helfrich1989; Steigmann Reference Steigmann1998, Reference Steigmann1999; Capovilla & Guven Reference Capovilla and Guven2002; Guven Reference Guven2004; Pollard, Al-Izzi & Morris Reference Pollard, Al-Izzi and Morris2024) and analysis (Seifert, Berndl & Lipowsky Reference Seifert, Berndl and Lipowsky1991; Seifert & Langer Reference Seifert and Langer1993; Bar-Ziv et al. Reference Bar-Ziv, Menes, Moses and Safran1995; Fournier Reference Fournier1996; Goldstein et al. Reference Goldstein, Nelson, Powers and Seifert1996; Seifert Reference Seifert1997; Powers, Huber & Goldstein Reference Powers, Huber and Goldstein2002; Du, Liu & Wang Reference Du, Liu and Wang2004; Vlahovska & Gracia Reference Vlahovska and Gracia2007; Agrawal & Steigmann Reference Agrawal and Steigmann2009; Stone Reference Stone2010; Agrawal & Steigmann Reference Agrawal and Steigmann2011; Rahimi, DeSimone & Arroyo Reference Rahimi, DeSimone and Arroyo2013; Maitra et al. Reference Maitra, Srivastava, Rao and Ramaswamy2014; Narsimhan, Spann & Shaqfeh Reference Narsimhan, Spann and Shaqfeh2015; Stone & Masoud Reference Stone and Masoud2015; Sabass & Stone Reference Sabass and Stone2016; Vlahovska Reference Vlahovska, Duprat and Stone2016; Al-Izzi, Sens & Turner Reference Al-Izzi, Sens and Turner2020; Fonda et al. Reference Fonda, Al-Izzi, Giomi and Turner2020; Lin et al. Reference Lin, Kumar, Richter, Wang, Schroeder and Narsimhan2021; Yu & Košmrlj Reference Yu and Košmrlj2023; Al-Izzi, Turner & Sens Reference Al-Izzi, Turner and Sens2024; Dharmavaram & Hanna Reference Dharmavaram and Hanna2024; Faizi, Granek & Vlahovska Reference Faizi, Granek and Vlahovska2024; Reboucas et al. Reference Reboucas, Faizi, Miksis and Vlahovska2024; Venkatesh & Narsimhan Reference Venkatesh and Narsimhan2024; Venkatesh, Bhargava & Narsimhan Reference Venkatesh, Bhargava and Narsimhan2025). However, the general, coupled nonlinear equations governing the in-plane and out-of-plane dynamics of a single-component membrane were not obtained until 2007 (Hu et al. Reference Hu, Zhang and W.2007). These governing equations were subsequently obtained via other techniques (Arroyo & DeSimone Reference Arroyo and DeSimone2009; Rangamani et al. Reference Rangamani, Agrawal, Mandadapu, Oster and Steigmann2012; Sahu, Sauer & Mandadapu Reference Sahu, Sauer and Mandadapu2017), and extended to describe the dynamics of multicomponent phase separation and chemical reactions with proteins in the surrounding fluid (Sahu et al. Reference Sahu, Sauer and Mandadapu2017). For our detailed perspective on the development of lipid membrane theories, see Chapter IV of Sahu (Reference Sahu2022).
The equations governing membrane dynamics are highly nonlinear partial differential equations written on a surface that is itself arbitrarily curved and deforming over time. One cannot in general analytically solve for the time evolution of lipid flows and membrane shape changes, which are intricately coupled. However, it is also difficult to solve the full membrane equations numerically, as standard solution methods from fluid and solid mechanics struggle to capture the membrane’s in-plane fluidity and out-of-plane elasticity. Many numerical studies accordingly investigated specific aspects of membrane behaviour. For example, several works captured the hydrodynamics of lipid flows on vesicles with a prescribed geometry, including the effects of the surrounding fluid as well as embedded proteins and other inclusions (Oppenheimer & Diamant Reference Oppenheimer and Diamant2010, Reference Oppenheimer and Diamant2011; Sigurdsson & Atzberger Reference Sigurdsson and Atzberger2016; Gross & Atzberger Reference Gross and Atzberger2018; Samanta & Oppenheimer Reference Samanta and Oppenheimer2021). These developments were recently extended with fluctuating hydrodynamics to incorporate phase separation and the discrete motion of proteins (Rower, Padidar & Atzberger Reference Rower, Padidar and Atzberger2022; Tran, Blanpied & Atzberger Reference Tran, Blanpied and Atzberger2022). Orthogonal efforts described elastic membrane deformations either (i) in the limit of no in-plane viscosity (Feng & Klug Reference Feng and Klug2006; Barrett et al. Reference Barrett, Garcke and Nürnberg2008a , Reference Barrett, Garcke and Nürnbergb ; Dziuk Reference Dziuk2008; Ma & Klug Reference Ma and Klug2008; Bonito, Nochetto & Pauletti Reference Bonito, Nochetto and Pauletti2010; Elliott & Stinner Reference Elliott and Stinner2010; Mercker et al. Reference Mercker, Marciniak-Czochra, Richter and Hartmann2013), possibly with the dynamics of the surrounding fluid (Narsimhan et al. Reference Narsimhan, Spann and Shaqfeh2015), or (ii) with the in-plane fluidity replaced by viscoelasticity (Rangarajan & Gao Reference Rangarajan and Gao2015; Zhu, Lee & Rangamani Reference Zhu, Lee and Rangamani2022). In both cases, the dynamic coupling between in-plane viscous flows and shape changes is not reflected (Sahu et al. Reference Sahu, Glisman, Tchoufag and Mandadapu2020a ). Still other works incorporated all of the aforementioned membrane complexities, but restricted their investigations to specific geometries (Rangamani et al. Reference Rangamani, Agrawal, Mandadapu, Oster and Steigmann2012, Reference Rangamani, Mandadapu and Oster2014; Walani, Torres & Agrawal Reference Walani, Torres and Agrawal2015; Hassinger et al. Reference Hassinger, Oster, Drubin and Rangamani2017) – for which the membrane equations are simplified.
We reiterate that many challenges arise when developing a general numerical method that truly captures in-plane viscous lipid flows, out-of-plane membrane bending and their coupling – all on an arbitrarily curved and deforming surface. Several studies (Barrett, Garcke & Nürnberg Reference Barrett, Garcke and Nürnberg2015; Rodrigues et al. Reference Rodrigues, Ausas, Mut and Buscaglia2015; Sauer et al. Reference Sauer, Duong, Mandadapu and Steigmann2017; Omar et al. Reference Omar, Sahu, Sauer and Mandadapu2020) took a Lagrangian approach, where the surface is discretised and the resulting mesh is convected with the physical, material velocity. Lagrangian implementations successfully capture membrane dynamics, but struggle to resolve in-plane flows as they lead to highly distorted elements. A remeshing procedure was used to maintain element aspect ratios, though it led to unphysical oscillations in the membrane curvature (Rodrigues et al. Reference Rodrigues, Ausas, Mut and Buscaglia2015). An alternative approach is for the mesh to only move normal to the surface, so that it is unaffected by in-plane flows. The mesh motion is then out-of-plane Lagrangian, as it tracks the material surface, and in-plane Eulerian. Such an approach, which we refer to as Eulerian, was implemented in Reuther, Nitschke & Voigt (Reference Reuther, Nitschke and Voigt2020) – though the membrane geometry was updated explicitly in a piecemeal manner. Given the importance of geometry to membrane dynamics (Sahu et al. Reference Sahu, Glisman, Tchoufag and Mandadapu2020a ), numerical methods relying on explicit mesh updates could suffer from issues known to affect explicit algorithms in the study of fluids (Zienkiewicz & Codina Reference Zienkiewicz and Codina1995; Zienkiewicz, Taylor & Zhu Reference Zienkiewicz, Taylor and Zhu2013). Moreover, even if a fully implicit Eulerian mesh motion was implemented as in our prior work (Sahu et al. Reference Sahu, Omar and Mandadapu2020b ), scenarios arise where the method will fail (Torres-Sánchez et al. Reference Torres-Sánchez, Millán and Arroyo2019). We show one such example in § 4.2.
Since neither Lagrangian nor Eulerian approaches can capture commonly observed membrane behaviours, a more general mesh motion is required. The ALE theory underlying a general mesh motion was independently derived by both others (Torres-Sánchez et al. Reference Torres-Sánchez, Millán and Arroyo2019) and ourselves (Sahu et al. Reference Sahu, Omar and Mandadapu2020b ). A new mesh motion was implemented in Torres-Sánchez et al. (Reference Torres-Sánchez, Millán and Arroyo2019), in which the mesh moved only in the direction normal to a known prior configuration of the membrane. In practice, such a choice was again not sufficiently general and required remeshing steps, which introduced errors in the numerical solution. Since the aforementioned approach resembles an Eulerian mesh motion, we hypothesise that it may not work in certain situations – such as the tether pulling scenario discussed in § 4.2. There is thus still a need for numerical implementations of more general mesh motions, which can be specified by the user when solving for membrane dynamics in a particular scenario.
The aforementioned limitations of numerical techniques motivate our development and implementation of a fully implicit ALE finite element method for lipid membranes. Rather than prescribing the mesh motion directly, we choose for the mesh velocity to satisfy a set of partial differential equations as if the mesh were itself another material. We then supply appropriate boundary conditions to the mesh velocity. In addition, the mesh and membrane are constrained to coincide via a Lagrange multiplier – which is understood as the mesh analogue of a normal force per unit area. The presence of a scalar Lagrange multiplier coupled to vector velocities leads to a numerical instability reminiscent of that identified by Ladyzhenskaya (Reference Ladyzhenskaya1969), Babuška (Reference Babuška1973) and Brezzi (Reference Brezzi1974), hereafter referred to as the LBB condition. We suppress the instability with the method of Dohrmann & Bochev (Reference Dohrmann and Bochev2004), where the Lagrange multiplier is projected onto a space of discontinuous, piecewise linear functions.
The remainder of this paper is organised as follows. In § 2, we present the strong and weak formulations of all equations governing the membrane and mesh. A summary of the corresponding finite element formulation is provided in § 3; additional details can be found in Appendix A. All numerical results from Lagrangian, Eulerian and ALE simulations are presented in § 4. We close with conclusions and pathways for future work in § 5. Our source code is provided in the Julia package MembraneAleFem.jl (Sahu Reference Sahu2024).
2. The governing equations: Strong and weak formulations
The membrane is treated as a single 2-D differentiable manifold embedded in the Euclidean space
$ \mathbb{R}^3$
, for which we implicitly assume no slip between the two bilayer leaflets. In our ALE formulation the membrane surface is parametrised by coordinates that need not follow material flows, as detailed in Sahu et al. (Reference Sahu, Omar and Mandadapu2020b
). Accordingly, the so-called mesh velocity
$ \boldsymbol{v}^{{{m}}}$
resulting from the parametrisation will in general not equal the membrane velocity
$ \boldsymbol{v}$
of material points. The dynamics of both the membrane and mesh are detailed in what follows; many of the results were derived in our prior work (Sahu et al. Reference Sahu, Omar and Mandadapu2020b
; Sahu Reference Sahu2022). The phospholipid bilayer is treated as an area-incompressible material – for which the areal density is constant and the surface tension,
$ \lambda$
, is a Lagrange multiplier field to be solved for. Our goal is to determine the fundamental unknowns
$ \boldsymbol{v}$
,
$ \boldsymbol{v}^{{{m}}}$
and
$ \lambda$
, as well as the membrane position
$ \boldsymbol{x}$
, over time.
2.1. Surface parametrisation, geometry and kinematics
We begin by reviewing the framework with which lipid membranes are described. Only the most relevant features are presented here, as Sahu (Reference Sahu2022) details our understanding of the membrane geometry and kinematics, while Sahu et al. (Reference Sahu, Omar and Mandadapu2020b ) provides the ALE theory.

Figure 1. Surface geometry. A schematic of the mapping
$ \boldsymbol{x} = \boldsymbol{x} (\zeta ^{\check {\alpha }}, t)$
, at a single instant in time, between the parametric domain
$ \varOmega$
and the membrane patch
$ \mathcal{P}$
. The in-plane basis vectors
$ \boldsymbol{a}_{\check {\alpha }}$
and unit normal vector
$ \boldsymbol{n}$
are shown at a point
$ \boldsymbol{x}$
on the patch, as are the in-plane unit normal
$ \boldsymbol{\nu }$
and unit tangent
$ \boldsymbol{\tau }$
at a point
$ \boldsymbol{x}^{}_{{{b}}}$
at the patch boundary
$ \partial \mkern -1mu \mathcal{P}$
.
Consider an arbitrarily curved and deforming membrane surface
$ \mathcal{S}$
, of which we examine a patch
$ \mathcal{P} \subset \mathcal{S}$
. At any time
$ t$
, the surface position
$ \boldsymbol{x}$
is parametrised by two coordinates:
$ \zeta ^{\check {1}}$
and
$ \zeta ^{\check {2}}$
. Here, as in Sahu et al. (Reference Sahu, Omar and Mandadapu2020b
), the ‘check’ accent
$ ( \, \check {\boldsymbol{\cdot }} \, )$
indicates the parametrisation can be arbitrarily specified. From now on, Greek indices span the set
$ \{ 1, 2 \}$
, such that
$ \boldsymbol{x} (\zeta ^{\check {\alpha }}, t) \in \mathbb{R}^3$
denotes the position of a point on the membrane surface. As shown in figure 1, there is a mapping from the parametric domain
$ \varOmega$
in the
$ \zeta ^{\check {1}}$
–
$ \zeta ^{\check {2}}$
plane to the membrane patch
$ \mathcal{P}$
. Area integrals over the patch are thus evaluated as

where
$ J^{}_{\! \varOmega }$
is the Jacobian of the mapping. Here
$ J^{}_{\! \varOmega }$
has dimensions of area and its functional form is provided below. At any point
$ \boldsymbol{x} (\zeta ^{\check {\alpha }}, t)$
, the vectors
$ \boldsymbol{a}^{}_{\check {\alpha }} := \partial \boldsymbol{x} / \partial \zeta ^{\check {\alpha }}$
form a basis of the tangent plane to the surface, which has unit normal
$ \boldsymbol{n} = (\boldsymbol{a}_{\check {1}} \times \boldsymbol{a}_{\check {2}}) / \lvert \boldsymbol{a}_{\check {1}} \times \boldsymbol{a}_{\check {2}} \rvert$
. Covariant components of the metric and curvature tensors are respectively
$ a_{{\check {\alpha }} {\check {\beta }}} := \boldsymbol{a}_{\check {\alpha }} \boldsymbol{\cdot } \boldsymbol{a}_{\check {\beta }}$
and
$ b_{{\check {\alpha }} {\check {\beta }}} := \boldsymbol{n} \boldsymbol{\cdot } \boldsymbol{x}_{, {\check {\alpha }} {\check {\beta }}} = \boldsymbol{n} \boldsymbol{\cdot } \boldsymbol{x}_{; {\check {\alpha }} {\check {\beta }}}$
. Here
$ (\, \boldsymbol{\cdot } \,)_{, {\check {\alpha }}} := \partial (\, \boldsymbol{\cdot } \,) / \partial \zeta ^{\check {\alpha }}$
and
$ (\, \boldsymbol{\cdot } \,)_{; {\check {\alpha }}}$
are respectively the partial and covariant derivatives with respect to
$ \zeta ^{\check {\alpha }}$
. The Jacobian
$ J^{}_{\! \varOmega }$
appearing in (2.1) is expressed as
$ J^{}_{\! \varOmega } = \sqrt { \rule {0mm}{1.50ex} } \overline { \det \, a_{{\check {\alpha }} {\check {\beta }}} \,}$
. The mean curvature
$ H$
is calculated as
$ H = ({1}/{2}) \mkern 1mu a^{\check {\alpha }\check {\beta }} b_{{\check {\alpha }} {\check {\beta }}}$
, where
$ a^{\check {\alpha }\check {\beta }}$
is the contravariant metric and is calculated as the matrix inverse of
$ a_{{\check {\alpha }} {\check {\beta }}}$
. In the present work, the Einstein summation convention is employed, in which repeated raised and lowered indices are summed over. The Gaussian curvature
$ K$
is computed as
$ K = \det (b_{{\check {\alpha }} {\check {\beta }}}) / \det (a_{{\check {\alpha }} {\check {\beta }}})$
. Finally, at points
$ \boldsymbol{x}^{}_{{{b}}}$
on the patch boundary
$ \partial \mkern -1mu \mathcal{P}$
, additional in-plane basis vectors are defined:
$ \boldsymbol{\nu } = \nu ^{\check {\alpha }} \mkern 1mu \boldsymbol{a}^{}_{\check {\alpha }} = \nu _{\check {\alpha }} \, \boldsymbol{a}^{\check {\alpha }}$
is the in-plane unit normal and
$ \boldsymbol{\tau } = \tau ^{\check {\alpha }} \mkern 1mu \boldsymbol{a}^{}_{\check {\alpha }} = \tau _{\check {\alpha }} \, \boldsymbol{a}^{\check {\alpha }}$
is the in-plane unit tangent (see figure 1). Line integrals over the patch boundary are calculated as

where
$ s$
is the arc length parametrisation of
$ \partial \mkern -1mu \mathcal{P}$
and
$ \varGamma := \partial \varOmega$
is the boundary of the parametric domain. In (2.2),
$ J^{}_{\varGamma } = [( \tau ^{\check {1}} )^2 + ( \tau ^{\check {2}} )^2]^{-1/2}$
is the Jacobian of the mapping from
$ \varGamma$
to
$ \partial \mkern -1mu \mathcal{P}$
.
When solving for the membrane state at a particular time, we seek the surface tension, material velocity and mesh velocity fields on the surface. To this end, it is useful to define relevant function spaces over the parametric domain
$ \varOmega$
. The space of square-integrable functions,
$ L^2 (\varOmega )$
, is expressed as

The Sobolev space of order
$ k$
,
$ H^k (\varOmega )$
, consists of functions that are square-integrable and also have up to
$ k$
partial derivatives that are square-integrable:

From the definitions in (2.3) and (2.4), it is evident that
$ H^0 (\varOmega ) = L^2 (\varOmega )$
. Moreover, since the material and mesh velocities are elements of
$ \mathbb{R}^3$
, we also define

as the space of functions in which each Cartesian component is an element of
$ H^k (\varOmega )$
. In (2.5) we denote
$ u^{}_{\!j} := \boldsymbol{u} \boldsymbol{\cdot } \boldsymbol{e}_{\!j}$
, where
$ \{ \boldsymbol{e}_{\!j} \}_{j = 1, 2, 3}$
is the canonical Cartesian basis in
$ \mathbb{R}^3$
.
We close with a discussion of membrane kinematics. The membrane velocity
$ \boldsymbol{v} = {\textrm{d}} \boldsymbol{x} / {\textrm{d}} t = \boldsymbol{\dot {x}}$
is the material time derivative of the position, and is expanded in the
$ \{ \boldsymbol{a}_{\check {\alpha }}, \boldsymbol{n} \}$
basis as
$ \boldsymbol{v} = v^{\check {\alpha }} \boldsymbol{a}_{\check {\alpha }} + v \boldsymbol{n}$
. The mesh velocity, which is treated as a fundamental unknown independent of the material velocity, is defined to be the rate of change of position when
$\zeta ^{\check {\alpha }}$
is held constant – expressed as
$ \boldsymbol{v}^{{{m}}} = ( \partial \check {\boldsymbol{x}} / \partial t ) \rvert _{\zeta ^{\check {\alpha }}}$
. In order for the material and mesh velocities to correspond to the same surface, kinematics require

such that the mesh motion is always out-of-plane Lagrangian (Sahu et al. Reference Sahu, Omar and Mandadapu2020b
). Importantly, the in-plane material velocity components
$v^{\check {\alpha }}$
and mesh velocity components
$ v^{\check {\alpha }}_{{{m}}} = \boldsymbol{v}^{{{m}}} \boldsymbol{\cdot } \boldsymbol{a}^{\check {\alpha }}$
need not coincide, such that within our ALE framework one can dictate how the mesh evolves within the membrane surface. In this work, (2.6) is referred to as the ALE kinematic constraint.
2.2. The balance of mass: Material incompressibility
As lipid membranes only stretch
$ 2\,\%{-}3 \,\%$
before tearing (Evans & Skalak Reference Evans and Skalak1980; Nichol & Hutter Reference Nichol and Hutter1996), we treat them as area-incompressible 2-D materials. The local form of the balance of mass, also referred to as the continuity equation and incompressibility constraint, is given by

where
$ t_{{\kern-1.5pt f}}$
is the end of the time interval under consideration. For notational simplicity, the functional dependence of all quantities will be suppressed, as will the domains of
$ \zeta ^{\mkern 3mu\check {\mkern -3mu\gamma }}$
and
$ t$
. The incompressibility constraint
$ \boldsymbol{a}^{\check {\alpha }} \boldsymbol{\cdot } \boldsymbol{v}_{, {\check {\alpha }}} = 0$
is equivalently expressed as
$ v^{\check {\alpha }}_{; {\check {\alpha }}} - 2 \mkern 1mu v H = 0$
, and is enforced by the Lagrange multiplier field
$ \lambda = \lambda (\zeta ^{\check {\alpha }}, t)$
– which has dimension of energy/area and acts as the membrane surface tension (Sahu Reference Sahu2022, Chapter V, § 6(a)–(c)).
The weak formulation of (2.7) is obtained by multiplying it with an arbitrary surface tension variation
$ \delta \lambda$
and integrating over the membrane area. The tension variation is assumed to be square-integrable, for which the weak form of the continuity equation is expressed as

Care must be taken when discretising (2.8) in the course of finite element analysis, as one could violate the LBB condition (Ladyzhenskaya Reference Ladyzhenskaya1969; Babuška Reference Babuška1973; Brezzi Reference Brezzi1974) and observe spurious surface tension oscillations in the numerical solution. A variety of techniques were developed to prevent such oscillations and numerically stabilise the system (Malkus & Hughes Reference Malkus and Hughes1978; Brezzi & Pitkäranta Reference Brezzi and Pitkäranta1984; Zienkiewicz & Nakazawa Reference Zienkiewicz and Nakazawa1984; Dortdivanlioglu et al. Reference Dortdivanlioglu, Krischok, Beirão da Veiga and Linder2018). We choose to employ the method of Dohrmann & Bochev (Reference Dohrmann and Bochev2004), as it is based on an underlying theory and is straightforward to implement numerically (Zienkiewicz et al. Reference Zienkiewicz, Taylor and Zhu2013).
2.2.1. The Dohrmann–Bochev method: Numerical stabilisation
The Dohrmann–Bochev method prevents spurious surface tension oscillations by projecting the surface tension onto a space of discontinuous, piecewise linear functions. The function space, denoted
$ \mkern 1.5mu\breve {\mkern -1.5mu L}$
, is defined in (3.7) following a discussion of the surface discretisation. The
$ L^2$
projection of a given surface tension field
$ \lambda \in L^2 (\varOmega )$
onto
$ \mkern 1.5mu\breve {\mkern -1.5mu L}$
is denoted
$ \breve {\lambda }$
and is defined through the relation (Dohrmann & Bochev Reference Dohrmann and Bochev2004)

In practice, the method of Dohrmann & Bochev (Reference Dohrmann and Bochev2004) is implemented by quadratically penalising surface tension deviations from the space
$ \mkern 1.5mu\breve {\mkern -1.5mu L}$
, for which the quantity

is subtracted from (2.8). Here,
$ \zeta$
is the 2-D intramembrane viscosity: a material property with dimensions of energy × time/area. In addition,
$ \alpha ^{\textit{DB}}$
is a user-chosen parameter with dimensions of area, such that (2.8) and (2.10) have the same dimensions – though the value of
$ \alpha ^{\textit{DB}}$
is not observed to affect simulation results. The numerically stabilised weak formulation of the incompressibility constraint is written as

where
$ \mathcal{G}_\lambda$
is the surface tension contribution to the direct Galerkin expression (Zienkiewicz & Taylor Reference Zienkiewicz and Taylor2014) given by

In (2.12),
$ \breve {\lambda }$
and
$ \delta \breve {\lambda }$
are understood to be the
$ L^2$
projections of their respective counterparts
$ \lambda$
and
$ \delta \lambda$
onto
$ \mkern 1.5mu\breve {\mkern -1.5mu L}$
, according to (2.9).
2.3. The balance of linear momentum: Membrane dynamics
Consider a general, arbitrarily curved and deforming 2-D material for which inertial effects are negligible. The local form of the balance of linear momentum for such a material is given by

where
$ \boldsymbol{T}^{\check {\alpha }}$
is the internal traction along a curve of constant
$\zeta ^{\check {\alpha }}$
on the surface and
$ \boldsymbol{f}$
is the net body force per unit area on the material by its surroundings. The balance of angular momentum additionally requires that the stress vectors be expressed as (Sahu Reference Sahu2022)

Here,
$ \sigma ^{{\check {\alpha }} {\check {\beta }}}$
contains the couple-free, in-plane stress components and
$ M^{{\check {\alpha }} {\check {\beta }}}$
contains the couple-stress components. If the constitutive form of
$ \sigma ^{{\check {\alpha }} {\check {\beta }}}$
and
$ M^{{\check {\alpha }} {\check {\beta }}}$
are known, then (2.13) and (2.14) govern the dynamics of a general 2-D material.
In lipid bilayers, the constitutive relations for
$ \sigma ^{{\check {\alpha }} {\check {\beta }}}$
and
$ M^{{\check {\alpha }} {\check {\beta }}}$
are well known, and depend on three material parameters. The first is the intramembrane viscosity
$ \zeta$
, which characterises the irreversibility of in-plane flows. The other parameters are the mean and Gaussian bending moduli, denoted by
$ k_{{b}}$
and
$ k_{{g}}$
, which have dimensions of energy and respectively penalise non-zero mean and Gaussian curvatures (Canham Reference Canham1970; Helfrich Reference Helfrich1973; Evans Reference Evans1974). We previously determined the membrane stresses and couple stresses within the framework of irreversible thermodynamics, and found that (Sahu Reference Sahu2022)

and

where

are the in-plane viscous stress components. The couple stresses (2.15) involve only
$ k_{{b}}$
and
$ k_{{g}}$
, and thus are purely elastic, while
$ \sigma ^{\check {\alpha }\check {\beta }}$
contains bending, tensile and viscous contributions (2.16).
The in-plane and out-of-plane equations governing lipid membrane dynamics are obtained by substituting (2.14)–(2.17) into (2.13) and contracting the result with
$ \boldsymbol{a}^{}_{\check {\alpha }}$
and
$ \boldsymbol{n}$
– which yields (Sahu Reference Sahu2022)

and

Equations (2.18) and (2.19) are respectively referred to as the in-plane and shape equations, and were independently obtained via different approaches (Hu et al. Reference Hu, Zhang and W.2007; Arroyo & DeSimone Reference Arroyo and DeSimone2009; Rangamani et al. Reference Rangamani, Agrawal, Mandadapu, Oster and Steigmann2012; Sahu et al. Reference Sahu, Sauer and Mandadapu2017). Here, the body force per unit area
$ \boldsymbol{f}$
is decomposed as
$ \boldsymbol{f} = \boldsymbol{a}^{\check {\alpha }} f_{\check {\alpha }} + f \boldsymbol{n}$
, and the operator
$ \varDelta _{{s}}$
acts on an arbitrary quantity
$ (\, \boldsymbol{\cdot } \,)$
as
$ \varDelta _{{s}} (\, \boldsymbol{\cdot } \,) := a^{{\check {\mu }} {\check {\nu }}} (\, \boldsymbol{\cdot } \,)_{; {\check {\mu }} {\check {\nu }}}$
. Note that while the membrane bends elastically, the in-plane viscosity
$ \zeta$
enters the shape equation (2.19) due to a coupling between in-plane stresses and curvature (Sahu et al. Reference Sahu, Glisman, Tchoufag and Mandadapu2020a
). The surface tension
$ \lambda$
and surface curvatures
$ H$
and
$ K$
also enter both the in-plane and shape equations, leading to non-trivial couplings between in-plane and out-of-plane dynamics.
2.3.1. The boundary conditions
One cannot determine a well-posed set of boundary conditions to (2.18) and (2.19) by inspection. A series of systematic developments for elastic shells (Green & Naghdi Reference Green and Naghdi1968; Steigmann Reference Steigmann1998, Reference Steigmann1999) underlie the formulation of the general lipid membrane boundary conditions (Capovilla, Guven & Santiago Reference Capovilla, Guven and Santiago2002; Arroyo & DeSimone Reference Arroyo and DeSimone2009; Rangamani et al. Reference Rangamani, Agrawal, Mandadapu, Oster and Steigmann2012; Sahu et al. Reference Sahu, Sauer and Mandadapu2017; Sauer & Duong Reference Sauer and Duong2017). In what follows, we highlight possible boundary conditions and provide their physical justification. Details of our own derivations, which reproduce earlier results, are provided in Chapter V, § 5(d) of Sahu (Reference Sahu2022).
The in-plane equations governing lipid flows (2.18) are identical to those governing a 2-D Newtonian fluid (Scriven Reference Scriven1960). We thus expect the boundary conditions to be similar to those of a fluid, in which one specifies either the velocity
$ \boldsymbol{v}$
or (for a surface) the force per length
$ \boldsymbol{F}$
on the boundary. For general 2-D materials, the force per length on the patch boundary is given by

which – for the case of lipid membranes – has bending, tensile and viscous contributions (recall
$ \boldsymbol{\nu }$
and
$ \boldsymbol{\tau }$
are boundary basis vectors; see figure 1). In our numerical implementation, on each edge of the membrane patch we specify a component of either the velocity
$ \boldsymbol{v}$
or the force per length
$ \boldsymbol{F}$
in each of the three Cartesian directions. We denote
$ \varGamma _{\!\! F}^{\mkern 1mu j}$
and
$ \varGamma _{\! v}^{\mkern 1mu j}$
as the respective portions of the boundary where
$ F_{\! j} := \boldsymbol{F} \boldsymbol{\cdot } \boldsymbol{e}_{\!j}$
and
$ v^{}_{\mkern -1mu j} := \boldsymbol{v} \boldsymbol{\cdot } \boldsymbol{e}_{\!j}$
are specified. For
$ j \in \{ 1, 2, 3 \}$
, the intersection
$ \varGamma _{\! v}^{\mkern 1mu j} \cap \varGamma _{\!\! F}^{\mkern 1mu j} = \varnothing$
and the closure of the union
$ \overline \varGamma _{\! v}^{\mkern 1mu j} \cup \varGamma _{\!\! F}^{\mkern 1mu j} = \varGamma$
. The boundary conditions are expressed as

where
$ \bar {v}^{}_{\mkern -1mu j}$
and
$ \mkern 4.5mu\overline {\mkern -4.5mu F \mkern -0.5mu}\mkern 0.5mu_{\! j}$
are known quantities that we prescribe.
The boundary conditions in (2.21) are necessary but not sufficient for a mathematically well-posed scenario. To see why, note that the membrane bending energy gives rise to the
$ \varDelta _{{s}} H = a^{{\check {\mu }} {\check {\nu }}} \mkern 1mu H_{; {\check {\mu }} {\check {\nu }}}$
term in the shape equation (2.19). Since the mean curvature contains two spatial derivatives of the surface position through the curvature components
$ b_{{\check {\alpha }} {\check {\beta }}} = \boldsymbol{n} \boldsymbol{\cdot } \boldsymbol{x}_{, {\check {\alpha }} {\check {\beta }}}$
, the
$ \varDelta _{{s}} H$
term contains four derivatives of the surface position. Following canonical developments in the theory of beam bending (Timoshenko Reference Timoshenko1921, Reference Timoshenko1922), we expect to specify two conditions along the entire boundary: either the out-of-plane velocity or force per length, as well as either the slope of the surface or the boundary moment

Since the velocity and force boundary conditions are already contained in (2.21), we need only specify one of the latter pair. More precisely, we partition the boundary into the disjoint sets
$ \varGamma _{\! n}$
and
$ \varGamma _{\! M}$
, with
$ \varGamma _{\! n} \cap \varGamma _{\! M} = \varnothing$
and
$ \overline { \mkern 1mu \varGamma _{\! n} \cup \varGamma _{\! M} } = \varGamma$
, and prescribe

where
$ \bar {v}_\nu$
and
$ \mkern 4.5mu\overline {\mkern -4.5mu M \mkern -1.5mu}\mkern 1.5mu$
are prescribed quantities. With (2.20)–(2.23), we have a mathematically well-posed set of boundary conditions for the governing equations.
2.3.2. The weak formulation
The weak formulation of the balance of linear momentum is obtained by first contracting (2.13) with an arbitrary velocity variation
$ \delta \boldsymbol{v}$
. At this point, we recognise the four spatial derivatives contained in the shape equation (2.19) through the
$ \varDelta _{{s}} H$
term will, in the subsequent weak form, yield two spatial derivatives of both the velocity variation and the surface position. Since
$ \boldsymbol{v}$
and
$ \boldsymbol{x}$
are assumed to lie in the same space of functions, both are elements of
$ \boldsymbol{H}^2 (\varOmega )$
– for which higher-order basis functions, with continuous first derivatives, are required in the numerical implementation. In accordance with the boundary conditions in (2.21) and (2.23), the space of admissible material velocity variations
$ \boldsymbol{\mathcal{V}}^{}_{\! 0}$
is expressed as

for
$ j \in \{ 1, 2, 3 \}$
. We integrate the result of the contraction over the membrane area, and substitute the general form of the stress vectors (2.14) to obtain

Starting with (2.25), a series of algebraic manipulations are required to determine the weak form of the linear momentum balance. The calculations rely on developments by Green & Naghdi (Reference Green and Naghdi1968) and Steigmann (Reference Steigmann1998), and can be found in Sauer et al. (Reference Sauer, Duong, Mandadapu and Steigmann2017), Sauer & Duong (Reference Sauer and Duong2017). Following the notation and development of our prior efforts (Sahu Reference Sahu2022, Chapter V,
$\mkern 1mu$
§ 5(d)), the weak formulation of the balance of linear momentum is expressed as

where

Note that jump forces can arise at corners of the patch and enter (2.27). They are assumed here to be zero. In (2.27),
$ \delta v_j := \delta \boldsymbol{v} \boldsymbol{\cdot } \boldsymbol{e}_{\!j}$
is the
$j$
th Cartesian component of the arbitrary velocity variation. It is useful to recognise that (2.27) is general to any material whose stress vectors can be expressed as in (2.14), with the corresponding boundary conditions in (2.21) and (2.23).
2.4. The dynamics of the mesh
In our ALE formulation, the material velocity
$ \boldsymbol{v}$
and the mesh velocity
$ \boldsymbol{v}^{{{m}}}$
are independent quantities. The evolution of the position
$ \boldsymbol{x} (\zeta ^{\check {\alpha }}, t)$
of the membrane surface is dictated by the mesh velocity via the relations

and

which both apply over all parametric coordinates
$ \zeta ^{\check {\alpha }} \mkern -1mu \in \varOmega$
and for all times
$ t \in [\mkern 1mu 0, t_{{\kern-1.5pt f}} \mkern 1mu]$
under consideration (see Sahu et al. (Reference Sahu, Omar and Mandadapu2020b
, § 2) for additional details). As we previously recognised
$ \boldsymbol{x} \in \boldsymbol{H}^2 (\varOmega )$
, (2.28) and (2.29) reveal
$ \boldsymbol{v}^{{{m}}} \in \boldsymbol{H}^2 (\varOmega )$
as well. In what follows, we discuss three mesh velocity schemes of increasing complexity, all of which are consistent with the kinematic constraint (2.6) such that the mesh and material always overlap. We begin with Lagrangian and Eulerian schemes, which were previously employed by both ourselves and others, and are known to suffer from limitations (Barrett et al. Reference Barrett, Garcke and Nürnberg2015; Rodrigues et al. Reference Rodrigues, Ausas, Mut and Buscaglia2015; Sauer et al. Reference Sauer, Duong, Mandadapu and Steigmann2017; Omar et al. Reference Omar, Sahu, Sauer and Mandadapu2020; Reuther et al. Reference Reuther, Nitschke and Voigt2020; Sahu et al. Reference Sahu, Omar and Mandadapu2020b
). We then discuss a new class of ALE schemes in which the mesh velocity itself satisfies dynamical equations similar to those that govern the material velocity.
2.5. The Lagrangian mesh motion
We begin with the simplest membrane motion, namely a Lagrangian scheme. In this case, the mesh velocity is prescribed to be equal to the material velocity:

In our numerical implementation, mesh velocity degrees of freedom are mapped to their material velocity counterparts such that (2.30) is satisfied identically over the entire surface. The direct Galerkin expression in the case of a Lagrangian mesh motion is then given by

In (2.31) the superscript ‘
${L}$
’ denotes a Lagrangian scheme, where
$ \boldsymbol{v}^{{{m}}} = \boldsymbol{v}$
throughout and mesh positions are updated according to (2.29). In addition,
$ \mathcal{G}_\lambda$
and
$ \mathcal{G}_{{v}}$
are respectively provided in (2.12) and (2.27).
2.6. The Eulerian mesh motion
The second scheme we implement is in-plane Eulerian, with
$ \boldsymbol{v}^{{{m}}} \boldsymbol{\cdot } \boldsymbol{a}^{}_{\check {\alpha }} = 0$
, and out-of-plane Lagrangian, for which
$ \boldsymbol{v}^{{{m}}} \boldsymbol{\cdot } \boldsymbol{n} = \boldsymbol{v} \boldsymbol{\cdot } \boldsymbol{n}$
. The in-plane and out-of-plane conditions are expressed as the single equation

where ‘
$ \otimes$
’ denotes the dyadic product. As there are no spatial derivatives in (2.32), no mesh velocity boundary conditions are to be prescribed for an Eulerian mesh motion.
The weak form of (2.32) is obtained by contracting it with an arbitrary mesh velocity variation
$ \delta \boldsymbol{v}^{{{m}}} \in \boldsymbol{H}^2 (\varOmega )$
, integrating over the membrane surface, and multiplying the result by a user-specified constant
$ \alpha _{{{m}}}^{\textit {E}}$
. The weak formulation of the Eulerian mesh velocity equation is then given by (Sahu et al. Reference Sahu, Omar and Mandadapu2020b
)

In (2.33) the superscript ‘
${E}$
’ signifies the scheme is Eulerian, and the constant
$ \alpha _{{{m}}}^{\textit {E}}$
has dimensions of energy × time/length
$ ^4$
such that
$ \mathcal{G}_{{m}}^{\textit {E}}$
has the same dimensions as
$ \mathcal{G}_\lambda$
and
$ \mathcal{G}_{{v}}$
. In this study,
$ \alpha _{{{m}}}^{\textit {E}} = \zeta / \ell ^2$
for all results presented, as varying
$ \alpha _{{{m}}}^{\textit {E}}$
was not observed to affect membrane behaviour (
$ \ell$
is a chosen characteristic length). The direct Galerkin expression for a scenario with an Eulerian mesh motion is expressed as

2.7. The ALE mesh motion
The final scheme we consider is neither in-plane Lagrangian nor in-plane Eulerian. Instead, the mesh is modelled as a separate 2-D material with its own constitutive relations and associated dynamical equations. The mesh analogues of the stress vectors are then expressed as (cf. (2.14))

where
$ \sigma ^{{\check {\alpha }} {\check {\beta }}}_{{m}}$
and
$ M^{{\check {\alpha }} {\check {\beta }}}_{{m}}$
are the mesh analogues of
$ \sigma ^{{\check {\alpha }} {\check {\beta }}}$
and
$ M^{{\check {\alpha }} {\check {\beta }}}$
. Here
$ \sigma ^{{\check {\alpha }} {\check {\beta }}}_{{m}}$
and
$ M^{{\check {\alpha }} {\check {\beta }}}_{{m}}$
can be specified arbitrarily without altering the dynamics of the membrane, and so the mesh dynamics can be that of an elastic, viscous or viscoelastic material. When the mesh velocity
$ \boldsymbol{v}^{{{m}}}$
is determined by solving an arbitrary set of governing equations, it will no longer satisfy the kinematic constraint (2.6) across the membrane surface. We enforce (2.6) with an additional Lagrange multiplier field, which physically acts on the mesh as an external body force per unit area in the normal direction – hereafter referred to as the mesh pressure
$ p^{{{m}}}$
. Following the developments of § 2.3, the dynamical equation governing the mesh dynamics is given by

Upon specification of appropriate boundary conditions, (2.6) and the three components of (2.36) uniquely determine the mesh pressure
$ p^{{{m}}}$
and the three components of the mesh velocity
$ \boldsymbol{v}^{{{m}}}$
. The ability to specify mesh boundary conditions can prevent undesirable mesh distortions, and is an advantage of our ALE method. Two choices of the mesh motion, for which
$ \sigma ^{{\check {\alpha }} {\check {\beta }}}_{{m}}$
and
$ M^{{\check {\alpha }} {\check {\beta }}}_{{m}}$
are prescribed, are discussed subsequently. In the first case, the mesh is area-compressible and purely viscous; in the second case it also resists bending. In both cases, the simplest boundary conditions for the mesh velocity are chosen. The investigation of more complex boundary conditions, as well as more involved mesh behaviours, is left to a future study.
2.7.1. The weak formulation of the kinematic constraint
The mesh pressure is an unknown Lagrange multiplier field enforcing the ALE kinematic constraint in (2.6). The weak form of (2.6) is obtained by multiplying it with an arbitrary variation
$ \delta p^{{{m}}} \in L^2 (\varOmega )$
, and integrating the result over the membrane surface to obtain

Equation (2.37) bears some resemblance to (2.8), which was obtained from the incompressibility constraint. In both equations, the variation of a scalar Lagrange multiplier interacts with a vector velocity. Moreover, when (2.37) is discretised, our numerical simulations exhibit an instability reminiscent of the checkerboarding that arises when the LBB condition is violated. We thus hypothesise that both numerical instabilities are similar, and attempt to stabilise the mesh pressure with the Dohrmann–Bochev method. Following the results of § 2.2.1, we express the weak formulation of the kinematic constraint (2.6) as

where the numerically stabilised mesh pressure contribution to the direct Galerkin expression is given by (cf. (2.12), (2.37))

In (2.39),
$ \breve {p}^{\mkern 1mu {{m}}}$
and
$ \delta \breve {p}^{\mkern 1mu {{m}}}$
are the
$ L^2$
projections of
$ p^{{{m}}}$
and
$ \delta p^{{{m}}}$
onto the space
$ \mkern 1.5mu\breve {\mkern -1.5mu L}$
, as defined through (2.9). The factor of
$ \ell ^2$
, for a characteristic length
$ \ell$
, is required for dimensional consistency. Our choice to employ the Dohrmann–Bochev method to stabilise
$ p^{{{m}}}$
does not yet sit on a firm theoretical footing, and currently can only be justified with empirical success. We hope to investigate the mathematical nature of the stabilisation of
$ p^{{{m}}}$
in a future paper.
2.7.2. The case where the mesh dynamics is purely viscous
For the case of purely viscous mesh dynamics, quantities are denoted with a superscript or subscript ‘
${v}$
’. The scheme is referred to as ‘ALE-viscous’, with the shorthand ‘Av’ or ‘ALE-v’. Since the mesh velocity field is area-compressible, the mesh motion resists both shearing and dilatation.
Strong formulation: we prescribe for the mesh analogue of the in-plane stresses and couple stresses to be (cf. (2.15)–(2.17))

where
$ \zeta ^{{{m}}}$
is a user-specified parameter that we refer to as the mesh viscosity. Importantly, since the mesh velocity is area-compressible, our choice in (2.40) resists both shear and dilatation of the mesh. By substituting (2.35) and (2.40) into (2.36), the dynamical equations governing the mesh are found to be (Sahu Reference Sahu2022, Chapter V,
$\mkern 1mu$
§ 5(c))

and

where
$ \boldsymbol{v}^{{{m}}} = v^{{{m}}}_{{\check {\alpha }}} \mkern 1mu \boldsymbol{a}^{\check {\alpha }} + v^{{{m}}} \mkern 1mu \boldsymbol{n}$
. The solution to (2.41) is independent of
$ \zeta ^{{{m}}}$
, and the mesh pressure in (2.42) will adjust so as to satisfy the kinematic constraint (2.6). The mesh dynamics accordingly do not depend on the choice of
$ \zeta ^{{{m}}}$
. In all numerical results presented, we choose
$ \zeta ^{{{m}}} = \zeta$
for simplicity.
Boundary conditions: since the mesh velocity satisfies differential equations (see (2.41) and (2.42)) rather than an algebraic equation (cf. (2.30), (2.32)), boundary conditions for the mesh velocity are required. To this end, we introduce the mesh analogue of the force per length on the boundary:

Recalling
$ M^{{\check {\alpha }} {\check {\beta }}}_{{{m}}, \mkern 1mu {v}} = 0$
according to (2.40), the purely viscous mesh dynamics cannot sustain moments – as is the case for a fluid film (Sahu et al. Reference Sahu, Omar and Mandadapu2020b
). Consequently, neither slope nor moment boundary conditions are required on the boundary, since the mesh analogue of the moment
$ M^{{m}} = 0$
identically throughout. With this understanding, the parametric boundary
$ \varGamma$
is partitioned into the portions
$ \varGamma _{\! v}^{{{m}} j}$
and
$ \varGamma _{\!\! F}^{{{m}} j}$
, on which
$ v^{{{m}}}_{\mkern -1mu j} := \boldsymbol{v}^{{{m}}} \boldsymbol{\cdot } \boldsymbol{e}_{\!j}$
and
$ F^{{m}}_{\! j} := \boldsymbol{F}^{{m}} \boldsymbol{\cdot } \boldsymbol{e}_{\!j}$
are respectively specified. For simplicity in our numerical formulation, the mesh is assumed to be either static or force-free on its entire boundary, for which (cf. (2.21))

Weak formulation: the weak formulation of the purely viscous ALE mesh behaviour is obtained by drawing analogy to the developments in § 2.3.2. First, the space of arbitrary mesh velocity variations is defined as

By contracting (2.36) with an arbitrary mesh velocity variation
$ \delta \boldsymbol{v}^{{{m}}} \in \boldsymbol{\mathcal{V}}^{{{m}}}_{\! 0, {v}}$
and repeating the steps of § 2.3.2, the weak formulation of the ALE-viscous mesh motion is found to be (cf. (2.26) and (2.27))

where

In (2.46) and (2.47), the superscript ‘Av’ refers to the choice of a purely viscous ALE mesh motion. The direct Galerkin expression for this motion is then written as

2.7.3. The case where the mesh dynamics is viscous and resist bending
We now consider the case where the mesh dynamics is viscous in response to in-plane shears and area dilatations, and additionally resist bending. The mesh motion is referred to as ‘ALE-viscous-bending’, with shorthand ‘Avb’ or ‘ALE-vb’. Corresponding quantities are denoted with a subscript or superscript ‘vb’.
Strong formulation: we choose for the bending resistance of the mesh to arise in the same manner as the membrane itself, with the viscous resistance identical to that of the ALE-viscous scenario. To this end, we choose (cf. (2.15)– (2.17) and (2.40))

and

where
$ \pi ^{{\check {\alpha }} {\check {\beta }}}_{{m}}$
is defined in (2.40). In (2.49) and (2.50),
$ k^{{m}}_{{b}}$
and
$ k^{{m}}_{{g}}$
are user-specified parameters; we always choose
$ k^{{m}}_{{b}} = k_{{b}}$
and
$ k^{{m}}_{{g}} = k_{{g}}$
for simplicity. It is well known that bending terms from the Helfrich free energy do not enter the in-plane equations (Rangamani et al. Reference Rangamani, Agrawal, Mandadapu, Oster and Steigmann2012), and so the in-plane dynamical mesh equations in the ALE-vb case are given by (2.41). The out-of-plane dynamical equation is expressed as (cf. (2.19) and (2.42))

Contrary to the ALE-v scenario (2.42), here the relative magnitude of
$ \zeta ^{{{m}}}$
and
$ k^{{m}}_{{b}}$
do affect the resultant mesh pressure
$ p^{{{m}}}$
. Investigating the relationship between
$ p^{{{m}}}$
,
$ \zeta ^{{{m}}}$
and
$ k^{{m}}_{{b}}$
– as well as the ensuing mesh dynamics – is left to a later study.
Boundary conditions: we once again introduce the mesh analogue of the force per length on the boundary, which in this case is written as (cf. (2.20) and (2.43))

In principle, one could specify boundary forces for the mesh dynamics. However, as in the ALE-viscous case, we choose simple boundary conditions where the mesh is either static or force-free on the entire boundary (2.44). In addition, since the mesh couple stresses are non-zero, either slope or moment boundary conditions are required for
$ \boldsymbol{v}^{{{m}}}$
– where the analogue of the boundary moment is calculated as
$ M^{{{m}}} = M^{{\check {\alpha }} {\check {\mu }}}_{{m}} \mkern 1mu \nu ^{}_{\check {\alpha }} \mkern 1mu \nu ^{}_{\check {\mu }}$
(cf. (2.22)). At present, we take the simpler approach and specify zero-slope conditions on the entire boundary (cf. (2.23)):

The ability to specify the slope of the mesh at a boundary will be relevant in the tether pulling scenario of § 4.2.
Weak formulation: at this point, it is straightforward to write the weak formulation of the mesh dynamics. The space of arbitrary mesh variations is given by

The weak formulation is then expressed as

where we define (cf. (2.27))

The direct Galerkin expression for the scenario with the ALE-vb mesh motion is written as

3. The finite element formulation
With the Lagrangian, Eulerian and ALE weak formulations, our goal is to solve for the state of the membrane over time. We seek to determine (i) the membrane velocity
$ \boldsymbol{v}$
, (ii) the mesh velocity
$ \boldsymbol{v}^{{{m}}}$
, and (iii) the membrane tension
$ \lambda$
– for all parametric points
$ \zeta ^{\check {\alpha }} \in \varOmega$
and times
$ t \in [\mkern 1mu 0, t_{{\kern-1.5pt f}} \mkern 1mu]$
. In doing so, the surface position is obtained from the mesh velocity through (2.29). We cannot solve for the highly nonlinear membrane behaviour exactly, and turn to the finite element method to numerically calculate the approximate solutions
$ \boldsymbol{v}^{}_h$
,
$ \boldsymbol{v}^{{{m}}}_h$
and
$ \lambda _h$
. An overview of our solution method is presented in what follows. Further specifics can be found in Appendix A and in the MembraneAleFem.jl documentation (Sahu Reference Sahu2024).
3.1. The surface discretisation and space of solutions
Let us begin by assuming there is a discretisation
$ \mathcal{T}_h$
of the parametric domain
$ \varOmega$
into nel (number of elements) non-overlapping finite elements
$ \varOmega ^e$
of characteristic length
$ h$
:

Our task is to choose the set of basis functions over each element
$ \varOmega ^e \in \mathcal{T}_h$
– denoted
$ \{ N^e_k (\zeta ^{\check {\alpha }}) \}$
, where the index
$ k$
ranges from
$ 1$
to
$ \textit {nen}$
(number of elemental nodes) – such that the resultant solution spaces are consistent with their infinite-dimensional counterparts. A complexity arises because mesh and membrane velocities belong to the space
$ \boldsymbol{H}^2 (\varOmega )$
. Accordingly,
$ \boldsymbol{v}^{}_h$
and
$ \boldsymbol{v}^{{{m}}}_h$
are required to be continuous and have continuous first derivatives across elements. We satisfy the continuity criteria with a so-called tensor product of quadratic B-spline basis functions in each of the
$ \zeta ^{\check {1}}$
and
$ \zeta ^{\check {2}}$
directions (Piegl & Tiller Reference Piegl and Tiller1997; Cottrell, Hughes & Bazilevs Reference Cottrell, Hughes and Bazilevs2009). For our purposes, the parametric domain
$ \varOmega$
is partitioned into a rectangular grid of finite elements, though more advanced discretisations are now established (Toshniwal, Speleers & Hughes Reference Toshniwal, Speleers and Hughes2017; Wei et al. Reference Wei, Zhang, Toshniwal, Speleers, Li, Manni, Evans and Hughes2018; Paul et al. Reference Paul, Zimmermann, Mandadapu, Hughes, Landis and Sauer2020; Koh, Toshniwal & Cirak Reference Koh, Toshniwal and Cirak2022). The resultant scalar and vector finite-dimensional spaces
$ \mathcal{U}_h$
and
$ \boldsymbol{\mathcal{U}}^{}_{\! h}$
are respectively given by

and

for which

by construction. In (3.2),
$ C^m (\varOmega )$
denotes the space of scalar functions on
$ \varOmega$
with
$ m$
continuous derivatives, and
$ \mathbb{Q}_{n} (\varOmega ^e)$
denotes the tensor-product space of
$ n$
th-order polynomials in the
$ \zeta ^{\check {1}}$
and
$ \zeta ^{\check {2}}$
directions on the finite element
$ \varOmega ^e$
. The arbitrary variations
$ \delta \boldsymbol{v}^{}_h$
and
$ \delta \boldsymbol{v}^{{{m}}}_h$
are also assumed to lie in
$\boldsymbol{\mathcal{U}}^{}_{\! h}$
, and are respectively elements of the spaces

3.1.1. The membrane tension, mesh pressure and Dohrmann–Bochev projection
In general, the Lagrange multipliers
$ \lambda$
and
$ p^{{{m}}}$
are elements of
$ L^2 (\varOmega )$
, and their finite-dimensional counterparts
$ \lambda _h$
and
$ p^{{{m}}}_h$
need not be restricted to lie in
$ \mathcal{U}_h$
. However, for convenience in the numerical implementation, we do in fact choose

Since the same basis functions are used for the velocity and surface tension, the well-known inf–sup condition is violated and the LBB instability arises (Ladyzhenskaya Reference Ladyzhenskaya1969; Babuška Reference Babuška1973; Brezzi Reference Brezzi1974). We apply the method of Dohrmann & Bochev (Reference Dohrmann and Bochev2004) to stabilise unphysical oscillations in both the surface tension and mesh pressure. To do so,
$ \lambda _h$
and
$ p^{{{m}}}_h$
are projected onto the space

where
$ \mathbb{P}_n (\varOmega ^e)$
denotes the space of polynomials of order
$ n$
on
$ \varOmega ^e$
. Since functions in
$ \mkern 1.5mu\breve {\mkern -1.5mu L}$
form a plane over each rectangular element
$ \varOmega ^e$
, they are discontinuous across finite elements.
3.2. The method of numerical solution
An overview of our numerical implementation is presented below. Additional details are provided in Appendix A and the MembraneAleFem.jl package repository (Sahu Reference Sahu2024).
Once the parametric domain is discretised into finite elements, all membrane unknowns are expressed as the sum of B-spline basis functions multiplied by membrane degrees of freedom. The degrees of freedom are collected into a column vector and are generically denoted
$ [ \boldsymbol{u} ]$
; they are also referred to as nodal values and their dependence on time is implied. Variations of membrane unknowns are similarly decomposed, with degree-of-freedom variations compatible with all boundary conditions collected into the column vector
$ [ \delta \boldsymbol{u} ]$
. Since the direct Galerkin expression
$ \mathcal{G}$
is linear in the arbitrary variations, its discretised counterpart
$ \mathcal{G}_h$
is expressed as

In (3.8) the residual vector
$ [ \boldsymbol{r} ]$
is defined to be

Since the variations
$ [ \delta \boldsymbol{u} ]$
in (3.9) are arbitrary, the membrane dynamics at any time
$ t \in [\mkern 1mu 0, t_{{\kern-1.5pt f}} \mkern 1mu]$
satisfy the nonlinear equation

Equation (3.10) is solved with the Newton–Raphson method, where a sequence of progressively better estimates of the membrane state
$ [ \boldsymbol{u} ]$
is generated. Given the
$ j$
th estimate
$ [ \boldsymbol{u} ]_j$
, the
$ (j + 1)$
th estimate is calculated according to

where the global tangent diffusion matrix at the
$ j$
th iteration is defined as

In prior studies (Sauer et al. Reference Sauer, Duong, Mandadapu and Steigmann2017; Sahu et al. Reference Sahu, Omar and Mandadapu2020b
), the tangent diffusion matrix was calculated analytically: an involved task. In the present work, however,
$ [ \boldsymbol{K} ]$
is calculated by numerically differentiating
$ [ \boldsymbol{r} ]$
with respect to each entry of
$ [ \boldsymbol{u} ]$
. In particular, the tangent matrix (3.12) is calculated to machine precision by extending
$ [ \boldsymbol{r} ]$
and
$ [ \boldsymbol{u} ]$
into the complex plane, following the general developments of Lyness & Moler (Reference Lyness and Moler1967), Lyness (Reference Lyness1968) (see also, e.g. Tanaka et al. Reference Tanaka, Fujikawa, Balzani and Schröder2014). We note that when calculating derivatives with respect to mesh velocity degrees of freedom, one must also perturb the surface positions, as the two are related through (2.28) and (2.29).
4. The results of numerical simulations
We now present results from numerical simulations, where our finite element implementation – including the various mesh motions discussed previously – was used to simulate the dynamics of lipid membranes. We first validate our code against a standard numerical benchmark, namely pure bending of a lipid membrane patch. We then simulate the drawing of a tube from a membrane sheet: a process known to be important in various biological phenomena, including dynamic rearrangement of the endoplasmic reticulum (Terasaki et al. Reference Terasaki, Chen and Fujiwara1986, Reference Terasaki2013) and tether formation by proteins travelling along microtubules (Leduc et al. Reference Leduc, Campàs, Zeldovich, Roux, Jolimaitre, Bourel-Bonnet, Goud, Joanny, Bassereau and Prost2004). The well-known quasi-static tether pulling behaviour (Powers et al. Reference Powers, Huber and Goldstein2002) is confirmed, and the effect of pull speed on pull force is quantified. We close by laterally translating the tether across the membrane surface with our ALE scheme – a result which cannot be obtained with established Lagrangian or Eulerian methods.
4.1. The pure bending of a flat patch
We begin with a simple scenario: starting with a flat membrane patch, boundary moments are applied to the left and right edges to bend the membrane, as shown schematically in figure 2(a). For a given applied moment, the equilibrium configuration is known analytically to be a portion of a cylinder (Sauer et al. Reference Sauer, Duong, Mandadapu and Steigmann2017; Sauer & Duong Reference Sauer and Duong2017). The case of pure bending is thus used to validate our numerical implementation, including the three mesh motions. For a given observable, errors in the numerical result
$ u_h$
with respect to the known analytical solution
$ u$
are calculated as

where
$ h$
denotes the characteristic length of a finite element in the parametric domain.

Figure 2. Pure bending scenario. (a) Boundary moments are applied to an initially flat patch in the
$x$
–
$y$
plane (top). The right edge is free to move in the
$x$
and
$y$
directions, but is constrained in the
$z$
direction. At equilibrium, the membrane forms a portion of a cylinder (bottom). (b) Time dependence of the prescribed boundary moment, according to (4.4). Errors are calculated at the final time
$ t_{{\kern-1.5pt f}}$
, as indicated by the blue arrow and circle. (c) Result of a numerical simulation on a
$ 128 \times 128$
mesh. The calculated surface tension differs from the analytical solution by less than three parts in 10 000, as shown by the colour bar.
4.1.1. The problem set-up
In our code, the membrane patch is initially a square of side length
$ \ell$
in the
$x$
–
$y$
plane. The initial velocity and position are respectively given by

where
$ \zeta ^{\check {\alpha }} \in \varOmega := [\mkern 1mu 0, 1] \times [\mkern 1mu 0, 1]$
and the real-space coordinates
$x$
and
$y$
are parametrised as

The magnitude of the applied moment, denoted
$ \mkern 4.5mu\overline {\mkern -4.5mu M \mkern -1.5mu}\mkern 1.5mu (t)$
, is linearly ramped up over time until
$ t = t_{{M}}$
– after which the boundary moment is held constant at the final value
$\mkern 4.5mu\overline {\mkern -4.5mu M \mkern -1.5mu}\mkern 1.5mu_{\! {{f}}}$
(see figure 2
b):

For times
$ t \gt t_{{M}}$
, the applied boundary moment is constant in time and the membrane is bent into a portion of a cylinder, as shown in figure 2(c). The boundary forces and moments of this stationary solution are calculated in Appendix B and summarised in table 1. Here,
$ r_{\mkern -1mu {{c}}}$
and
$ \lambda _0$
are respectively the cylinder radius and surface tension, the latter of which is constant in the equilibrium configuration. We choose the Gaussian bending modulus
$ k_{{g}} = - k_{{b}} / 2$
such that
$ M = 0$
on the top and bottom edges, and no body force (
$ \boldsymbol{f} = \boldsymbol{0}$
) such that
$ \lambda _0 = k_{{b}} / (4 \mkern 1mu r_{\mkern -1mu {{c}}}^{\, 2})$
and
$ \boldsymbol{F} = \boldsymbol{0}$
on the left and right edges (see table 1). In this case, the cylinder radius
$ r_{\mkern -1mu {{c}}}$
is related to the final boundary moment
$ \mkern 4.5mu\overline {\mkern -4.5mu M \mkern -1.5mu}\mkern 1.5mu_{\! {{f}}}$
according to

All of the boundary conditions prescribed in our numerical implementation for Lagrangian and Eulerian mesh motions are provided in table 2. For the ALE simulations, we choose the ALE-v mesh motion discussed in § 2.7.2. The first three rows of table 2 are then repeated for the mesh counterparts
$ \bar {v}^{{{m}}}_{\mkern -1mu j}$
and
$ \mkern 4.5mu\overline {\mkern -4.5mu F \mkern -0.5mu}\mkern 0.5mu^{{m}}_{\! j}$
. Recall that
$ M^{{{m}}} = 0$
by construction, so there is no mesh analogue of the moment boundary condition. As a result, the membrane deforms only due to the physically applied moment
$ \mkern 4.5mu\overline {\mkern -4.5mu M \mkern -1.5mu}\mkern 1.5mu (t)$
– which is also the case for the Lagrangian and Eulerian simulations.
Table 1. Moments and forces on the boundary of a portion of a cylindrical membrane with radius
$ r_{\mkern -1mu {{c}}}$
, as shown in figure 2 and calculated in Appendix B. The constant surface tension
$ \lambda _0$
is set by the net body force per area in the normal direction,
$ f := \boldsymbol{f} \boldsymbol{\cdot } \boldsymbol{n}$
. In figure 2(a) the cylindrical basis vector
$ \boldsymbol{e}^{}_\theta$
is tangent to the black curve and points to the left.

Table 2. Boundary conditions prescribed in the numerical implementation. We choose to set
$ k_{{g}} = - k_{{b}} / 2$
and
$ \boldsymbol{f} = \boldsymbol{0}$
; the latter yields
$ \lambda _0 = k_{{b}} / (4 \mkern 1mu r_{\mkern -1mu {{c}}}^{\, 2})$
. The prescribed moment
$ \mkern 4.5mu\overline {\mkern -4.5mu M \mkern -1.5mu}\mkern 1.5mu(t)$
is given by (4.4). The first three rows are repeated for
$ \bar {v}^{{{m}}}_{\mkern -1mu j}$
and
$ \mkern 4.5mu\overline {\mkern -4.5mu F \mkern -0.5mu}\mkern 0.5mu^{{m}}_{\! j}$
in the case of a viscous ALE mesh motion.

4.1.2. Non-dimensionalisation
For the case of pure bending, there are five membrane parameters: the bending modulus
$ k_{{b}}$
, final bending moment
$ \mkern 4.5mu\overline {\mkern -4.5mu M \mkern -1.5mu}\mkern 1.5mu_{\! {{f}}}$
, ramp-up time
$ t_{{M}}$
, patch length
$ \ell$
and intramembrane viscosity
$ \zeta$
. These parameters all dimensionally consist of mass, length and time. The pure bending scenario is thus completely described by two dimensionless quantities. The first is the Föppl–von Kármán number, which compares surface tension to bending forces (Sahu et al. Reference Sahu, Glisman, Tchoufag and Mandadapu2020a
). For a cylinder with
$ \boldsymbol{f} = \boldsymbol{0}$
, the membrane surface tension is constant:
$ \lambda = \lambda _0 := k_{{b}} / (4 \mkern 1mu r_{\mkern -1mu {{c}}}^{\, 2}) = \mkern 4.5mu\overline {\mkern -4.5mu M \mkern -1.5mu}\mkern 1.5mu_{\! {{f}}}^2 / k_{{b}}$
, where (4.5) was substituted in the last equality (Evans & Yeung Reference Evans and Yeung1994). The Föppl–von Kármán number
$ {\varGamma }$
is then given by

The second dimensionless quantity is the Scriven–Love number
$ {{S\scriptstyle {L}}}$
, which captures dynamical effects and in this scenario is set by
$ t_{{M}}$
(Sahu et al. Reference Sahu, Glisman, Tchoufag and Mandadapu2020a
). As the characteristic velocity scale of membrane deformations is given by
$ \ell / t_{{M}}$
, the Scriven–Love number is expressed as

which can be understood as a ratio between the fundamental membrane time scale
$ \ell ^2 \zeta / k_{{b}}$
and the ramp-up time
$ t_{{M}}$
. For the results presented, we run simulations with
$ \zeta = 1.0$
,
$ k_{{b}} = 1.0$
and
$ \ell = 1.0$
. The Föppl–von Kármán and Scriven–Love numbers are then set by choosing
$ \mkern 4.5mu\overline {\mkern -4.5mu M \mkern -1.5mu}\mkern 1.5mu_{\! {{f}}}$
and
$ t_{{M}}$
, respectively.

Figure 3. Errors in the (a) surface tension and (b) mean curvature at the final time
$ t_{{\kern-1.5pt f}}$
, according to (4.1), when the membrane patch is subjected to pure bending boundary conditions. The mesh consists of
$ \textit {nel} = 1 / h^2$
parametric area elements, each of which is a square with side length
$ h$
. Here,
$ 1 / h$
ranges from
$ 2^1$
to
$ 2^7$
in powers of two. The labels ‘E’, ‘Av’ and ‘L’ refer to Eulerian, ALE-viscous and Lagrangian mesh motions. In all cases, the convergence of the error confirms our numerical implementation is working as expected. Relevant parameters are specified in § 4.1.3; we also choose
$ \zeta = 1.0$
,
$ k_{{b}} = 1.0$
and
$ \ell = 1.0$
.
4.1.3. Results
We simulate pure bending scenarios where
$ {\varGamma } = 0.25$
and
$ {{S\scriptstyle {L}}} = 0.5$
, for which
$ \mkern 4.5mu\overline {\mkern -4.5mu M \mkern -1.5mu}\mkern 1.5mu_{\! {{f}}} = k_{{b}} / (2 \mkern 1mu \ell )$
and
$ t_{{M}} = 2 \mkern 1mu \ell ^2 \zeta / k_{{b}}$
(see (4.6) and (4.7)). Lagrangian, Eulerian and ALE-viscous simulations are carried out on meshes ranging from
$ 2 \times 2$
to
$ 128 \times 128$
. Other relevant parameters are
$ \alpha ^{\textit{DB}} = \ell ^2$
,
$ \alpha _{{{m}}}^{\textit {E}} = \zeta / \ell ^2$
,
$ \Delta t = 0.1 \mkern 1mu \ell ^2 \zeta / k_{{b}}$
and
$ t_{{\kern-1.5pt f}} = 4 \mkern 1mu t_{{M}}$
. The final configuration from one such simulation, at time
$ t_{{\kern-1.5pt f}}$
, is shown in figure 2(c). The
$L^2$
error in surface tension and mean curvature, relative to the analytical solution and calculated at time
$ t_{{\kern-1.5pt f}}$
according to (4.1), is plotted as a function of mesh size in figure 3. Note that the analytical solution is valid only for a static patch; we thus deform the membrane slowly relative to the intrinsic time scale
$ \ell ^2 \zeta / k_{{b}}$
and calculate errors after it has relaxed sufficiently. All three mesh motions converge towards the analytical solution upon mesh refinement, indicating our numerical implementation is working as expected. We believe the error from the Eulerian mesh motion is larger than that of the other two schemes because boundary conditions are not prescribed for the mesh velocity. Instead, material velocity boundary conditions are prescribed and weakly communicated to the mesh velocity through (2.33); edges are also the location where errors are the largest (see figure 2
c).
4.2. The pulling of a tether from a flat sheet
When a single point on the membrane is displaced in the direction
$ \boldsymbol{n}$
normal to the surface, a tent-like shape forms when deformations are small. As the point continues to be displaced, the bilayer undergoes a non-trivial morphological transition and forms a cylindrical tether (Evans & Yeung Reference Evans and Yeung1994; Dai & Sheetz Reference Dai and Sheetz1995; Fygenson, Marko & Libchaber Reference Fygenson, Marko and Libchaber1997; Raucher & Sheetz Reference Raucher and Sheetz1999). Tethers are known to arise in biological settings, including in the endoplasmic reticulum (Terasaki et al. Reference Terasaki, Chen and Fujiwara1986; Scott et al. Reference Scott, Steen, Huber, Westrate and Koslover2024) and at the junction between cells (Sowinski et al. Reference Sowinski2008; Dubey et al. Reference Dubey2016; Imachi et al. Reference Imachi2020). In addition, tethers form in in vitro settings with optical tweezers (Evans & Yeung Reference Evans and Yeung1994; Cuvelier et al. Reference Cuvelier, Derényi, Bassereau and Nassoy2005) and through the polymerisation of microtubules (Fygenson et al. Reference Fygenson, Marko and Libchaber1997) – possibly with the use of molecular motors (Roux et al. Reference Roux, Cappello, Cartaud, Prost, Goud and Bassereau2002; Koster et al. Reference Koster, VanDuijn, Hofs and Dogterom2003). Tether pulling is also commonly used in assays to probe both static and dynamic membrane properties, as one can measure the force
$ \boldsymbol{\mathcal{F}}_{\! {{p}}}$
required to pull the tether (Cuvelier et al. Reference Cuvelier, Derényi, Bassereau and Nassoy2005; Koster et al. Reference Koster, Cacciuto, Derényi, Frenkel and Dogterom2005; Shi et al. Reference Shi, Graber, Baumgart, Stone and Cohen2018). An analysis of membrane tether pulling is thus relevant to both biological and in vitro scenarios. In what follows, we focus on the simplest case, in which a tether is pulled from an initially flat sheet.
4.2.1. The problem set-up
In numerical simulations the membrane patch is initially a square of side length
$ \ell$
in the
$x$
–
$y$
plane, centred at the origin. We set
$ v_z = 0$
on the entire patch boundary, and also pin the centre node of each edge (
$ \boldsymbol{v} = \boldsymbol{0}$
) to remove rigid body rotations and translations. In addition, zero-slope boundary conditions are enforced by constraining
$ v_z = 0$
for nodes adjacent to the boundary – hereafter referred to as inner boundary nodes. Finally, an in-plane force/length
$ \boldsymbol{\bar {F}}_\parallel = \lambda _0 \, \boldsymbol{\nu }$
is applied on the boundary, where
$ \boldsymbol{\nu }$
is the in-plane unit normal at the patch boundary and
$ \lambda _0$
is the static surface tension – which we choose. All boundary conditions are summarised in table 3, including those for the Eulerian and ALE schemes where
$ \boldsymbol{v}^{{{m}}}$
is also a fundamental unknown.
Table 3. Boundary conditions prescribed in our numerical implementation to pull a tether, with Lagrangian (LAG), Eulerian (EUL) and ALE-viscous-bending (ALE-vb) mesh motions. Here, ‘Bdry’ refers to nodes on the boundary, ‘Inner’ refers to inner nodes adjacent to the boundary, ‘Pull’ corresponds to nodes associated with the central element
$ \varOmega ^e_{{p}}$
and ‘Pin’ refers to nodes at the centre of each edge. The latter conditions are required to prevent rigid body rotations and translations. The ‘Inner’ column enforces zero-slope boundary conditions.

The initial membrane state is given by

where
$ \zeta ^{\check {\alpha }} \in \varOmega := [\mkern 1mu 0, 1] \times [\mkern 1mu 0, 1]$
. The corresponding membrane position is expressed as

with

At time
$ t = 0$
, we seek to vertically displace the centre of the membrane patch, where
$ x = 0$
and
$ y = 0$
. However, as shown in Appendix C, there is no unique way to specify the velocity at a single point given our use of B-spline basis functions, as they are not interpolatory. Instead, we vertically displace the portion of the membrane corresponding to the entire parametric element
$ \varOmega ^e_{{p}}$
containing the point
$ \zeta _{{p}} \hspace {-3.7pt} {}^{{\check {\alpha }}} := (0.5, 0.5)$
at the centre of the parametric domain:

In (4.11),
$ \boldsymbol{v}^{}_{{p}} (t)$
is a known function of time. It is given by
$ v^{}_{\!{p}} \mkern 1mu \boldsymbol{e}_z$
, for constant
$ v^{}_{\!{p}}$
, unless otherwise specified. Following the derivations in Appendix C, (4.11) is enforced by setting all nodal velocity degrees of freedom associated with
$ \varOmega ^e_{{p}}$
to
$ \boldsymbol{v}^{}_{{p}} (t)$
. The resultant pull force
$ \boldsymbol{\mathcal{F}}_{\! {{p}}} (t)$
is calculated as the sum of the corresponding components of the residual vector.
4.2.2. Non-dimensionalisation
In the scenario under consideration, there are five membrane parameters: the bending modulus
$ k_{{b}}$
, equilibrium surface tension
$ \lambda _0$
, 2-D intramembrane viscosity
$ \zeta$
, patch length
$ \ell$
and speed of tube drawing
$ v^{}_{\!{p}}$
. We hope to include the hydrodynamics of the surrounding fluid, including membrane permeability (Alkadri & Mandadapu Reference Alkadri and Mandadapu2024), and the effects of cytoskeletal contacts (Shi et al. Reference Shi, Graber, Baumgart, Stone and Cohen2018) in a future effort – and better compare our results with experiments (Cuvelier et al. Reference Cuvelier, Derényi, Bassereau and Nassoy2005; Brochard-Wyart et al. Reference Brochard-Wyart, Borghi, Cuvelier and Nassoy2006). Since the five quantities dimensionally involve only mass, length and time, two dimensionless numbers once again determine the evolution of the system. The Föppl–von Kármán number
$ {\varGamma }$
is given by (Sahu et al. Reference Sahu, Glisman, Tchoufag and Mandadapu2020a
)

and can be interpreted in two ways given the morphological changes that occur when pulling a tether. First, when the membrane is nearly planar and height deflections are small, gradients in membrane shape occur over the length scale
$ \ell$
;
$ {\varGamma }$
then quantifies the relative magnitude of tensile and bending forces. In our simulations,
$ {\varGamma }^{-1}$
is small, and consequently a boundary layer of characteristic width
$ \sqrt { k_{{b}} / \lambda _0 } \ll \ell$
develops at the point of application of the pull force (Powers et al. Reference Powers, Huber and Goldstein2002). As the membrane is pulled further, a tether grows from the boundary layer region. The tether is close in shape to a cylinder, and the radius of a cylindrical membrane is well known to be (Zhong-can & Helfrich Reference Zhong-can and Helfrich1989)

Equation (4.13) can be obtained from energetic arguments alone, or equivalently, from a balance of bending and tensile forces. Note that in situations where there is a pressure jump across the membrane due to the surrounding fluid and
$ f = \boldsymbol{f} \boldsymbol{\cdot } \boldsymbol{n} \ne 0$
, (4.13) and (4.17) are not valid. See the discussion in Chapter IX, § 1(a) of Sahu (Reference Sahu2022). With (4.13), the Föppl–von Kármán number (4.12) is understood as the ratio

Equations (4.12)–(4.14) confirm that the Föppl–von Kármán number describes membrane energetics, as only lengths and the parameters
$ k_{{b}}$
and
$ \lambda _0$
– which respectively have dimensions of energy and energy per area – are involved.
The second dimensionless quantity is the Scriven–Love number
$ {{S\scriptstyle {L}}}$
, which compares the magnitude of viscous and bending forces in shaping the membrane (Sahu et al. Reference Sahu, Glisman, Tchoufag and Mandadapu2020a
). In this scenario, we define the Scriven–Love number as

We chose to include
$ r_{\mkern -1mu {{c}}}$
– rather than
$ \ell$
– in (4.15) because we are primarily concerned with the behaviour of the tether, rather than the entire patch. Upon substituting (4.13) into (4.15), rearranging terms and recognising
$ \zeta / \lambda _0$
is the fundamental time scale of lipid flows (Sahu et al. Reference Sahu, Glisman, Tchoufag and Mandadapu2020a
), we find the Scriven–Love number can be equivalently expressed as

i.e. a ratio of the speed at which the tether is pulled to the natural velocity scale of lipid flows in the tube (Sahu et al. Reference Sahu, Glisman, Tchoufag and Mandadapu2020a
). Thus, as
$ {{S\scriptstyle {L}}}$
tends to zero, we approach the quasi-static limit.
In our source code, when running tether pulling simulations, we choose
$ \zeta = 1.0$
,
$ k_{{b}} = 1.0$
and
$ \lambda _0 = 0.25$
. Our choice yields a tube radius
$ r_{\mkern -1mu {{c}}} = 1.0$
according to (4.13). The Föppl–von Kármán number
$ {\varGamma }$
is altered by varying the patch size
$ \ell$
, while the Scriven–Love number
$ {{S\scriptstyle {L}}}$
is modified by changing the pull speed
$ v^{}_{\!{p}}$
.
4.2.3. The comparison of different mesh motions
We begin by pulling a tether in a scenario with Föppl–von Kármán number
$ {\varGamma } = 1024$
and Scriven–Love number
$ {{S\scriptstyle {L}}} = 0.1$
. Given these dimensionless numbers,
$ \ell = 64 \mkern 1mu r_{\mkern -1mu {{c}}}$
and
$ v^{}_{\!{p}} = 0.1 \mkern 1mu k_{{b}} / (\zeta \mkern 1mu r_{\mkern -1mu {{c}}})$
according to (4.14) and (4.15). Snapshots from tether pulling simulations with each mesh motion are shown in figure 4, with corresponding videos in the MembraneAleFem.jl package repository (Sahu Reference Sahu2024). A zoomed-in view of the late-time snapshots, with and without the underlying mesh, are shown in figure 5 to emphasise the advantages of the ALE scheme. In addition, figure 6 shows how the
$ z$
component of the pull force,
$ \mathcal{F}_{\! {{p}}, \mkern 1mu z} := \boldsymbol{\mathcal{F}}_{\! {{p}}} \boldsymbol{\cdot } \boldsymbol{e}_z$
, varies as a function of the vertical displacement
$ z^{}_{{p}}$
. We note that in the quasi-static limit (
$ {{S\scriptstyle {L}}} \rightarrow 0$
), the steady-state pull force of a perfect cylinder is given by Evans & Yeung (Reference Evans and Yeung1994)

Note that some studies define
$ \kappa := k_{{b}} / 2$
to be the bending modulus, in which case
$ \mathcal{F}_{\! {eq}} = 2 \pi \sqrt { 2 \mkern 1mu \kappa \mkern 1mu \lambda _0 }$
. Since a tether pulled from a flat patch deviates slightly from a cylinder (Powers et al. Reference Powers, Huber and Goldstein2002), (4.17) serves as an excellent approximation for
$ \mathcal{F}_{\! {{p}}, \mkern 1mu z}$
when
$ {{S\scriptstyle {L}}} = 0$
.
In what follows, we comment on the efficacy of the three mesh motions: Lagrangian (LAG), Eulerian (EUL) and ALE-viscous-bending (ALE-vb). Note that additional variables are used to solve for membrane behaviour in the latter two schemes: the fundamental variables are
$ \boldsymbol{v}$
and
$ \lambda$
(LAG);
$ \boldsymbol{v}$
,
$ \boldsymbol{v}^{{{m}}}$
and
$ \lambda$
(EUL); and
$ \boldsymbol{v}$
,
$ \boldsymbol{v}^{{{m}}}$
,
$ \lambda$
and
$ p^{{{m}}}$
(ALE-vb). We present the number of degrees of freedom for each motion, as well as the wall clock run time, in table 4. When appropriate, our results are compared to those of prior theoretical and numerical developments.
Table 4. Number of degrees of freedom (DOFs) and wall clock run time for the three different mesh motions, corresponding to the results presented in figure 4. All computations were carried out on a single node of the Perlmutter system at the National Energy Research Scientific Computing Center, with area element calculations distributed across 32 cores.

$^{\textit{a}}$
Scaled to the total number of time steps, as the simulation failed.

Figure 4. Snapshots from tether pulling simulations with a
$ 65 \times 65$
mesh, in which three different mesh motions are employed. The colour bar indicates the surface tension, in units of
$ k_{{b}} / r_{\mkern -1mu {{c}}}^2$
, for all snapshots. Times are measured in units of
$ \zeta r_{\mkern -1mu {{c}}}^2 / k_{{b}}$
. (Left) Lagrangian simulations (LAG) successfully generate a tether. The morphological shape change from a tent to a tube occurs around time
$ t = 120$
. Since the membrane is area incompressible, the patch boundary must be pulled inwards to accommodate the tether surface area. (Centre) The Eulerian mesh motion (EUL) is not able to form a tube. Rather, the tent morphology persists and bulges outward until the method fails around time
$ t = 143$
. The material velocity degrees of freedom of the central element are constrained to move upwards; no such constraint is placed on the mesh velocity degrees of freedom. (Right) An ALE mesh motion that is viscous and resists bending (ALE-vb) successfully forms a tether. Both material and mesh velocities of the central element are constrained to move upwards. Since the mesh is area compressible, the patch boundary can be constrained to remain stationary as the tether develops. For all three mesh motions,
$ \Delta t = 0.5$
,
$ {{S\scriptstyle {L}}} = 0.1$
and
$ {\varGamma } = 1024$
, for which
$ \ell = 64 \, r_{\mkern -1mu {{c}}}$
and
$ v^{}_{\!{p}} = 0.1 \, k_{{b}} / (\zeta r_{\mkern -1mu {{c}}} )$
according to (4.14) and (4.15). See also supplementary movies 1–3.

Figure 5. Zoomed-in views of the
$ t = 240 \, \zeta r_{\mkern -1mu {{c}}}^2 / k_{{b}}$
snapshots from figure 4, with the underlying mesh shown (top) and hidden (bottom). In the Lagrangian simulations (left), the mesh is drawn into the tube along with the lipids due to the areal incompressibility of the membrane. Mesh elements close to the diagonal of the square patch, where
$ y = \pm \mkern 1mu x$
, become highly distorted – which leads to an artificially low surface tension in the transition region between the tether and surrounding membrane. Numerical artefacts are visible in the striation of the tension. The ALE-viscous-bending result (right), in contrast, shows a less distorted mesh because the choice of mesh constitution resists both shear and dilatation. No surface tension artefacts are visible, and a smooth tension gradient is observed from the flat patch into the tether.

Figure 6. The
$ z$
component of the pull force (
$ \mathcal{F}_{\! {{p}}, \mkern 1mu z}$
) as a function of the
$ z$
displacement. The dashed cyan line is the result from the linear theory, as presented in (4.20). The Eulerian simulation (blue circles, dotted line) is unable to form a tether, and is unphysical. The Lagrangian (red triangles, dashed line) and ALE (black squares, solid line) simulations capture the tent-to-tube transition, after which the Lagrangian steady-state pull force
$ \mathcal{F}_{\! {\textit{ss}}}$
is slightly larger. Both overshoot the equilibrium pull force
$ \mathcal{F}_{\! {eq}}$
(4.17) due to dynamical effects from tether pulling (see § 4.2.4). Numerical parameters are those specified in figure 4.
Lagrangian scheme: we first consider results from simulations with a Lagrangian mesh motion, as shown in the left column of figures 4 and 5 (see supplementary movie 1 available at https://doi.org/10.1017/jfm.2025.10553). A tent-like structure develops and grows until the vertical displacement
$ z^{}_{{p}} := v^{}_{\!{p}} \mkern 1mu t \approx 8 \mkern 1mu r_{\mkern -1mu {{c}}}$
, for which
$ t \approx 80 \, \zeta \mkern 1mu r_{\mkern -1mu {{c}}}^{\, 2} / k_{{b}}$
. Around this point, the linear relationship between
$ \mathcal{F}_{\! {{p}}, \mkern 1mu z}$
and
$ z^{}_{{p}}$
breaks down, as shown in figure 6. The membrane undergoes a morphological transition between
$ t \approx 80 \, \zeta \mkern 1mu r_{\mkern -1mu {{c}}}^{\, 2} / k_{{b}}$
and
$ t \approx 120 \, \zeta \mkern 1mu r_{\mkern -1mu {{c}}}^{\, 2} / k_{{b}}$
, during which
$ \mathcal{F}_{\! {{p}}, \mkern 1mu z}$
reaches its largest value. After the shape transition, a tether continues to be drawn at an approximately constant force, which we refer to as the steady-state force
$ \mathcal{F}_{\! {\textit{ss}}}$
(see figure 6). However, a pronounced area of low surface tension develops in the region where the tether meets the flat patch (see figures 4 and 5). We comment on the unphysical nature of this development following a presentation of our ALE results.
The majority of our findings from Lagrangian tether pulling simulations, as well as the Lagrangian implementation itself, are not new. Powers et al. (Reference Powers, Huber and Goldstein2002) provided a detailed account of the axisymmetric, quasi-static membrane shape in response to vertical displacements, and along with Derényi et al. (Reference Derényi, Jülicher and Prost2002) numerically calculated the force versus displacement curve in the quasi-static limit. More recently, rate effects were included in axisymmetric simulations that focused primarily on the slip between two monolayer leaflets (Rahimi & Arroyo Reference Rahimi and Arroyo2012) – an effect not considered in the present work. General non-axisymmetric Lagrangian formulations were developed since then. The implementation in Rodrigues et al. (Reference Rodrigues, Ausas, Mut and Buscaglia2015) included rate effects when pulling a tether, but the force was not provided as a function of displacement and it is unclear if the membrane underwent the morphological transition from a tent to a tube in simulations. In contrast, Sauer et al. (Reference Sauer, Duong, Mandadapu and Steigmann2017) pulled a tether from the centre of an initially axisymmetric mesh and illustrated the tent-to-tube transition. While this implementation seems to be capable of capturing rate effects, the reported tether pulling results employed a numerical stabilisation scheme that did not capture the coupling between in-plane lipid flows and out-of-plane forces. Thus, to the best of our knowledge, we present the first non-axisymmetric Lagrangian simulation capturing the dynamics of a tether pulled from a membrane patch – though we do not consider the results to be novel, as Rodrigues et al. (Reference Rodrigues, Ausas, Mut and Buscaglia2015) and Sauer et al. (Reference Sauer, Duong, Mandadapu and Steigmann2017) may have been able to do the same.
Eulerian scheme: let us next examine results from an Eulerian mesh motion. As shown in supplementary movie 2 and the centre column of figure 4, the numerical implementation fails to form a tether. In addition, unphysical gradients develop in the surface tension, and the code fails at time
$ t = 143 \, \zeta \mkern 1mu r_{\mkern -1mu {{c}}}^{\, 2} / k_{{b}}$
. In what follows, we discuss why such a failure is to be expected, as a tether cannot form when the mesh motion is Eulerian.
We begin by introducing
$ J$
and
$ J^{{{m}}}$
as the relative area dilatations of the membrane and mesh, respectively. These dilatations are related to the corresponding flow fields according to (Sahu Reference Sahu2022, Chapter V, § 1(c))

Since the membrane is incompressible,
$ \dot {J} / J = 0$
and
$ v^{\check {\alpha }}_{; {\check {\alpha }}} = 2 \mkern 1mu v H$
. To understand how the mesh dilates or compresses, we recognise that
$ v^{{{m}}} = v$
and
$ v^{\check {\alpha }}_{{{m}}} = 0$
according to (2.32). We thus find that

At this point, we make three observations regarding the geometry and dynamics of the initial tent formation: (i)
$ v \ge 0$
everywhere, (ii)
$ H \lt 0$
in a central region where the surface is concave down, and (iii)
$ H \ge 0$
elsewhere. Accordingly,
$ \dot {J}^{{m}} / J^{{{m}}} \gt 0$
in the centre of the tent, and the mesh continuously dilates. In this manner, the mesh expands laterally and under-resolves the region where the morphological transition would occur.
Mesh dilatation at the patch centre, where
$ \zeta ^{\check {1}} = 0.5$
and
$ \zeta ^{\check {2}} = 0.5$
(see (4.10)), is quantified in figure 7. Here, the relative area change of the mesh
$ J^{{{m}}}$
is plotted as a function of the membrane displacement
$ z^{}_{{p}}$
. The dilatation by two orders of magnitude suggests an Eulerian simulation with a similar increase in the number of elements. Such a patch, even after the large dilatation, may be sufficiently resolved to undergo the morphological transition to a tube. However, even if a tether was formed, it likely could not be extended with an Eulerian mesh motion: lipid flows are primarily in-plane during tether elongation, so the Eulerian mesh would remain stationary and the tether would be poorly resolved. We thus believe the inability to pull a tether is a general failure of Eulerian methods, including those implemented previously (Reuther et al. Reference Reuther, Nitschke and Voigt2020; Sahu et al. Reference Sahu, Omar and Mandadapu2020b
). Moreover, since the ALE implementation of Torres-Sánchez et al. (Reference Torres-Sánchez, Millán and Arroyo2019) employed a mesh motion that was close to Eulerian, we are unsure if it would be able to successfully pull a tether.

Figure 7. Mesh dilatation of the Eulerian simulation shown in figure 4. (a) The relative area change at the patch centre,
$ J^{{{m}}}$
, is plotted as a function of the displacement
$ z^{}_{{p}}$
. The dashed vertical line is the displacement at which a tent is expected to transition to a tether, according to Lagrangian and ALE data in figure 6. (Inset) At small displacements, simulation data agrees with an approximate solution to (4.19). Here
$ v = \dot {z}_{{{p}}}$
; if the mean curvature is linear in
$ z^{}_{{p}}$
then
$ H \sim - z^{}_{{p}} / r_{\mkern -1mu {{c}}}^{\, 2}$
and integrating (4.19) yields
$ \ln J^{{{m}}} \sim (z^{}_{{p}} / r_{\mkern -1mu {{c}}})^2$
. The number 16 in the analytical expression is a fitting parameter. (b) The same data in a log–log plot suggests a power-law growth of the dilatation at large displacements, prior to the expected morphological transition (dotted vertical line) – though the data does not span even a single decade.
ALE scheme: we close by discussing the results of our ALE mesh motion. As shown in figures 4–6 and supplementary movie 3, a tether was successfully pulled with the ALE-vb scheme, and the surface tension in the region where the tether meets the flat patch was approximately constant. This latter point is quantified by considering the range of surface tension values at time
$ t = 240 \, \zeta \mkern 1mu r_{\mkern -1mu {{c}}}^{\, 2} / k_{{b}}$
. In the ALE simulation, the lowest (dimensionless) value of the surface tension is
$ 0.249$
– approximately equal to the magnitude of the in-plane force per length
$ \lvert \boldsymbol{\bar {F}}_\parallel \rvert = \lambda _0 = 0.25$
maintained on the boundary. Larger tension values, in this case up to
$ 0.345$
, arise in the tether to draw in lipids during the dynamic process of pulling. In contrast, the minimum value of the tension in Lagrangian simulations is
$ 0.191$
: well below
$ \lvert \boldsymbol{\bar {F}}_\parallel \rvert$
and also less than the larger values (up to
$ 0.446$
) attained on the tether. Since lipids flow from regions of low to high tension, there does not seem to be a smooth flow of lipids from the flat patch into the tether. We thus find the ALE-viscous-bending mesh motion to be superior to its Lagrangian counterpart, and report only ALE-vb results for the remainder of our tether pulling analysis.
It is important to note that while the ALE-vb scheme successfully pulls a tether, the purely viscous ALE mesh motion – employed in § 4.1 – does not: as the centre of the patch is translated upwards, the tether tapers to a point. We believe an angular shape arises in the ALE-v simulation because of incompatible boundary conditions for the material and mesh. More specifically, since all nodes over a single element are translated upwards, a zero-slope condition is implicitly prescribed at the element boundary. Both the membrane and ALE-vb mesh can sustain such a requirement due to their inherent bending stiffness. In contrast, the ALE-v mesh has no bending rigidity and so cannot maintain zero slope on the element boundary. The tether resulting from the ALE-v scheme is consequently unphysical.
4.2.4. The geometry and dynamics of tether pulling
In comparing different mesh motions in § 4.2.3, all simulations were carried out at a single choice of
$ {\varGamma }$
and
$ {{S\scriptstyle {L}}}$
. We now investigate the effects of altering the patch size relative to the tether radius, as well as changing the speed of tether pulling – which respectively modify the Föppl–von Kármán and Scriven–Love numbers. We confirm that
$ {\varGamma }$
dictates the initial slope of the force versus displacement curve, as previously observed (Derényi et al. Reference Derényi, Jülicher and Prost2002; Powers et al. Reference Powers, Huber and Goldstein2002; Sauer et al. Reference Sauer, Duong, Mandadapu and Steigmann2017), while
$ {{S\scriptstyle {L}}}$
captures the overshoot of the tether pull force relative to the equilibrium (or quasi-static) result.
We first investigate the dependence of the pull force on the patch geometry. Figure 8(a) plots
$ \mathcal{F}_{\! {{p}}, \mkern 1mu z}$
as a function of
$ z^{}_{{p}}$
for different values of
$ {\varGamma }$
, at fixed
$ {{S\scriptstyle {L}}} = 0.1$
. We observe that
$ {\varGamma }$
alters the initial slope of the force versus displacement curve, but does not affect the steady-state pull force after the tether is formed – the latter of which is expected to be independent of the patch size (see (4.17) and Powers et al. Reference Powers, Huber and Goldstein2002; Derényi et al. Reference Derényi, Jülicher and Prost2002). To calculate the initial dependence of
$ \mathcal{F}_{\! {{p}}, \mkern 1mu z}$
on
$ z^{}_{{p}}$
, we analyse the tent-like membrane shape when deformations are small. In this limit, the shape equation (2.19) is decoupled from in-plane lipid flows, and identical to its quasi-static counterpart (Sahu et al. Reference Sahu, Glisman, Tchoufag and Mandadapu2020a
). We thus take the small-deformation, quasi-static membrane tent result from Powers et al. (Reference Powers, Huber and Goldstein2002) and calculate the pull force to be given by (cf. (4.17))

which is shown as the dashed cyan line in figure 6. Our calculation of (4.20) is provided in Appendix D, and the collapse of force versus displacement curves is shown in figure 8(b).
The dependence of the pull force on the speed of tether pulling is investigated next. Results from simulations with variable
$ {{S\scriptstyle {L}}}$
and fixed
$ {\varGamma } = 1024$
are plotted in figure 8(c). The observed increase in pull force with increasing Scriven–Love number can be justified as follows. During tether pulling, lipids in the surrounding region flow inwards towards the tether. A mass balance, with the assumption of axisymmetry, indicates that the flow speed is approximately
$ v^{}_{\!{p}} \mkern 1mu r_{\mkern -1mu {{c}}} / r$
, where
$ r$
is the distance from the
$ z$
axis. Since the in-plane velocity grows as we approach the tether from its surroundings, a surface tension gradient is required to sustain the flow of lipids. As the tether is pulled more quickly, larger velocities and, thus, larger tension gradients ensue. A greater surface tension in the tether results, and leads to the larger pull force observed in simulations.
At present, we are unable to determine a general functional form for the long-time, steady-state pull force as a function of
$ {{S\scriptstyle {L}}}$
due to the high degree of nonlinearity in the governing equations. The main difficulty is that the surface tension and membrane geometry enter both the in-plane and shape equations. Moreover, the viscous–curvature coupling forces – which arise due to the flow of lipids – lead to
$ O({{S\scriptstyle {L}}})$
changes of the membrane shape in the tent-like region. Since we are unable to approximate how the steady-state pull force depends on the Scriven–Love number, we choose not to collapse the data contained in figure 8(c). Instead, we plot the steady-state pull force
$ \mathcal{F}_{\! {\textit{ss}}}$
as a function of the Scriven–Love number in figure 8(d). We expect
$ \mathcal{F}_{\! {\textit{ss}}} \approx \mathcal{F}_{\! {eq}}$
when
$ {{S\scriptstyle {L}}} \rightarrow 0$
. In the limit where
$ {{S\scriptstyle {L}}} \ll 1$
, we approximate dynamical effects by calculating the change in surface tension under the assumption of no membrane shape changes. We find that

which is shown as the dotted grey line in figure 8(d). Evidently, (4.21) is a reasonable approximation when
$ {{S\scriptstyle {L}}} \lt 0.1$
.

Figure 8. The
$ z$
component of the pull force (
$ \mathcal{F}_{\! {{p}}, \mkern 1mu z}$
) as a function of
$ z^{}_{{p}}$
for different values of
$ {\varGamma }$
and
$ {{S\scriptstyle {L}}}$
. (a) When
$ {{S\scriptstyle {L}}} = 0.1$
and
$ {\varGamma }$
is varied, the initial slope of the force vs displacement curve is altered according to the linear theory (see (4.20)). (b) By scaling the
$ z$
displacement appropriately, the data collapses – with the steady-state pull force independent of
$ {\varGamma }$
after tether formation. The cyan line has slope one. (c) When
$ {\varGamma } = 1024$
and
$ {{S\scriptstyle {L}}}$
is varied, the initial slope of the force vs displacement curve is unchanged. The tent-to-tube transition occurs at larger displacements, and the long-time pull force increases with
$ {{S\scriptstyle {L}}}$
. (d) Long-time pull force,
$ \mathcal{F}_{\! {\textit{ss}}}$
, as a function of
$ {{S\scriptstyle {L}}}$
for
$ {\varGamma } = 1024$
. The dotted grey line is the linear prediction from (4.21), which agrees with simulation results when
$ {{S\scriptstyle {L}}} \ll 1$
– as highlighted by the zoomed-in inset. The nonlinear dependence of
$ \mathcal{F}_{\! {\textit{ss}}}$
on
$ {{S\scriptstyle {L}}}$
arises from the coupling between in-plane flows and out-of-plane shape deformations.

Figure 9. Tether translation in the
$ + \boldsymbol{e}_x$
(left column) and then
$ - \boldsymbol{e}_x$
(right column) directions. Times are measured in units of
$ \zeta \mkern 1mu r_{\mkern -1mu {{c}}}^{\, 2} / k_{{b}}$
, the pull velocity is specified in (4.22), and all other parameters are identical to those in figure 4. The vertical red lines are a visual aid to highlight the lateral translation of the tether. For a video of the simulation results, see supplementary movie 4.
4.2.5. The lateral translation of a membrane tether
Once a tube is pulled from a patch of membrane, a lateral force applied at the end of the tether causes it to translate relative to the surrounding region. Lipids quickly flow and readjust to accommodate the translating tether due to the in-plane fluidity of the membrane. While tether translation via a lateral force is observed in in vitro studies (Datar et al. Reference Datar, Bornschlögl, Bassereau, Prost and Pullarkat2015; Gomis Perez et al. Reference Gomis Perez, Dudzinski, Rouches, Landajuela, Machta, Zenisek and Karatekin2022; Shi, Innes-Gold & Cohen Reference Shi, Innes-Gold and Cohen2022), the physics of tether translation remains poorly understood. A theoretical description of tether translation is difficult because the system is no longer axisymmetric, and the greatly simplified axisymmetric equations (Derényi et al. Reference Derényi, Jülicher and Prost2002; Powers et al. Reference Powers, Huber and Goldstein2002; Agrawal & Steigmann Reference Agrawal and Steigmann2009,Reference Agrawal and Steigmann2011; Omar et al. Reference Omar, Sahu, Sauer and Mandadapu2020) no longer apply. Tether translation also cannot be captured with general, non-axisymmetric Lagrangian numerical methods – as laterally translating the tether induces a rigid body translation of the entire patch (see Appendix E). Arbitrary Lagrangian–Eulerian finite element methods are thus required to simulate tether translation.
In figure 9 we present snapshots from a simulation of tether translation, which employed the ALE-vb mesh motion. Starting with the
$ t = 240 \, \zeta \mkern 1mu r_{\mkern -1mu {{c}}}^{\, 2} / k_{{b}}$
configuration shown in figures 4 and 5, we prescribe a lateral velocity in the
$ x$
direction given by

where in both cases
$ v^{}_{\!{p}} = 0.1 \mkern 1mu k_{{b}} / (\zeta \mkern 1mu r_{\mkern -1mu {{c}}})$
to be consistent with our choice
$ {{S\scriptstyle {L}}} = 0.1$
. In this simulation,
$ {\varGamma } = 1024$
is unchanged. We observe that the surface tension does not appreciably change, as the tether is translated slowly relative to the fundamental time scale of lipid rearrangements. Our results demonstrate that ALE methods can be used in scenarios where Lagrangian methods fail, and set the stage for future investigations of the forces, geometry and dynamics of tether translation. A video of the laterally pulled tether simulation is provided in the software repository (Sahu Reference Sahu2024).
5. Conclusions and future work
In the present work we (i) developed a robust ALE numerical method for lipid membranes and (ii) applied the method to a scenario where established Lagrangian and Eulerian schemes fail. In our development, the mesh is treated as a material that is independent from the membrane – with the mesh equations of motion and corresponding boundary conditions arbitrarily prescribed by the user. Mesh and material surfaces are constrained to overlap with a Lagrange multiplier field, which enters the mesh dynamics as a force per area in the normal direction. By choosing for the mesh to resist shear and dilatation (through the mesh viscosity
$ \zeta ^{{{m}}}$
) and bending (through the mesh bending moduli
$ k^{{m}}_{{b}}$
and
$ k^{{m}}_{{g}}$
), we successfully pulled a tether from a flat patch and then translated it laterally across the membrane surface. In contrast, Lagrangian and Eulerian simulations are respectively unable to translate and pull a tether. Our results thus mark the first numerical demonstration of lateral tether translation. We also analysed the dynamics of tether pulling by determining how the pull force increases monotonically with increasing pull speed. Our findings were presented in terms of the Föppl–von Kármán number
$ {\varGamma }$
and Scriven–Love number
$ {{S\scriptstyle {L}}}$
, which define the tether pulling scenario.
We close by highlighting that our numerical implementation (Sahu Reference Sahu2024) is structured such that one can easily choose different constitutive relations and boundary conditions for the mesh motion. In this manner, the mesh itself can resist shear, dilatation or their rates of change – irrespective of the material behaviour. Only purely viscous and viscous-with-bending mesh motions were considered in the present work; we aim to investigate different choices of mesh motion in a future study. As a consequence of the modular structure of the numerical implementation, it is straightforward to adapt the code to simulate 2-D materials with different constitution. We hope to support the open-source community in doing so. We also intend to extend our method to describe additional phenomena governing biological membranes – including the coupling between lipid flows and the hydrodynamics of the surrounding fluid (Narsimhan et al. Reference Narsimhan, Spann and Shaqfeh2015; Gross & Atzberger Reference Gross and Atzberger2018; Faizi et al. Reference Faizi, Granek and Vlahovska2024), membrane permeability (Alkadri & Mandadapu Reference Alkadri and Mandadapu2024), the effects of embedded particles (Stone Reference Stone2010; Stone & Masoud Reference Stone and Masoud2015; Sabass & Stone Reference Sabass and Stone2016) and in-plane diffusion and phase transitions in multicomponent membrane systems (Subramaniam et al. Reference Subramaniam, Guidotti, Manoharan and Stone2013; Yu & Košmrlj Reference Yu and Košmrlj2023; Venkatesh & Narsimhan Reference Venkatesh and Narsimhan2024; Venkatesh et al. Reference Venkatesh, Bhargava and Narsimhan2025).
Supplementary movies
Supplementary movies are available at https://doi.org/10.1017/jfm.2025.10553.
Acknowledgements
We are indebted to Professor Kranthi Mandadapu for many scientific discussions on membrane dynamics and finite element methods within the setting of differential geometry. We are also grateful to Professor Panos Papadopoulos and Dr. Yannick Omar for bringing important details of finite element analysis to our attention, including the method of Dohrmann & Bochev (Reference Dohrmann and Bochev2004) and numerical differentiation of the residual vector (Tanaka et al. Reference Tanaka, Fujikawa, Balzani and Schröder2014). It is a pleasure to thank Dr. Joël Tchoufag regarding conversations on fluid mechanics and membrane dynamics in different geometries. Simulations were carried out on the Perlmutter high-performance computing system at the National Energy Research Scientific Computing Center. Our open-source package is written in the Julia programming language (Bezanson et al. Reference Bezanson, Edelman, Karpinski and Shah2017). Figures and videos of the simulation results were generated with the open-source Julia package Makie.jl (Danisch & Krumbiegel Reference Danisch and Krumbiegel2021). This work was partially supported by the Welch Foundation via Grant No. F-2208-20240404.
Declaration of interests
The authors report no conflict of interest.
Appendix A. The finite element implementation
The main novelty of the present work is the treatment of the mesh as an independent material with its own constitution, as discussed in § 2.7. However, none of the finite element techniques used here are new. All fundamental unknowns, their arbitrary variations and the surface position itself are discretised with the same basis functions – as is standard in isoparametric finite element methods. Over any element
$ \varOmega ^e \in \mathcal{T}_h$
, we have

In (A1) the matrix
$ [ \boldsymbol{N}^e (\zeta ^{\check {\alpha }}) ]$
contains the non-zero elemental basis functions (see § 3.1) and the column vector
$ [ \boldsymbol{x}^e (t) ]$
collects the corresponding local degrees of freedom (or nodal positions) in the usual manner of finite element analysis:

Here
$ [ \boldsymbol{1} ]$
is the
$ 3 \times 3$
identity matrix in Cartesian coordinates and

The velocity and mesh velocity are similarly decomposed as

and

for all
$ \zeta ^{\check {\alpha }} \in \varOmega ^e$
. Since the membrane tension and mesh pressure are scalar quantities, they are respectively discretised as

and

for all
$ \zeta ^{\check {\alpha }} \in \varOmega ^e$
, where
$ [ \mkern -9mu\boldsymbol{\mkern 9mu\boldsymbol{N}^e} (\zeta ^{\check {\alpha }}) ]$
is the row vector of non-zero elemental basis functions given by (cf. (A2))

Assuming a known solution at time
$ t$
, the temporal evolution of the membrane is obtained via the backward Euler method. In particular, the membrane surface is updated according to (cf. (2.28), (2.29))

or equivalently,

As discussed in § 3.2, the mesh velocity and all other unknowns satisfy the residual vector equation

which is solved via the Newton–Raphson method – in which
$ [ \boldsymbol{u} (t) ]$
is used as the initial guess. We thus close our discussion of the finite element implementation by presenting the contributions to the residual vector; all details can be found in the software documentation (Sahu Reference Sahu2024).
A.1. The contributions to the residual vector
Our finite element implementation allows one to choose whether the mesh motion is Lagrangian, Eulerian or ALE – and, in the latter case, whether the mesh dynamics is purely viscous or viscous with a bending resistance. The direct Galerkin expressions in (2.31), (2.34), (2.48) and (2.57) dictate the residual vector for each mesh motion. In what follows, we present the local residual vector contributions – corresponding to a single finite element
$ \varOmega ^e \in \mathcal{T}_h$
– for each fundamental unknown.
A.1.1. The surface tension contribution
As detailed in Sahu et al. (Reference Sahu, Omar and Mandadapu2020b ), the residual vector corresponding to (2.12) is given by

where all quantities in curly braces arise from the method of Dohrmann & Bochev (Reference Dohrmann and Bochev2004). The matrices
$ [ \boldsymbol{G}^e ]$
and
$ [ \boldsymbol{H}^e ]$
in (A12) are constructed from the basis functions
$ \mkern -4.5mu\breve {\mkern 4.5mu N^e_{k}} (\zeta ^{\check {\alpha }})$
to the space
$ \mkern 1.5mu\breve {\mkern -1.5mu L}$
onto which surface tensions are projected. Since functions in
$ \mkern 1.5mu\breve {\mkern -1.5mu L}$
are linear and form a plane over
$ \varOmega ^e$
(see (3.7)), they can be expressed as
$ a + b \mkern 1mu \zeta ^{\check {1}} + c \mkern 1mu \zeta ^{\check {2}} = 0$
for some constants
$ a, b, c \in \mathbb{R}$
. In our numerical implementation, the planar basis functions over a single element are chosen to be
$ \, \mkern -4.5mu\breve {\mkern 4.5mu N^e_{1}} (\zeta ^{\check {\alpha }}) = 1$
,
$ \mkern -4.5mu\breve {\mkern 4.5mu N^e_{2}} (\zeta ^{\check {\alpha }}) = \xi (\zeta ^{\check {1}})$
and
$ \mkern -4.5mu\breve {\mkern 4.5mu N^e_{3}} (\zeta ^{\check {\alpha }}) = \eta (\zeta ^{\check {2}})$
, where
$ \xi$
and
$ \eta$
parametrise the reference square
$ [-1, 1] \times [-1, 1]$
onto which
$ \varOmega ^e$
is mapped. Accordingly, over any element
$ \varOmega ^e \in \mathcal{T}_h$
we define the row vector

with which the matrices
$ [ \boldsymbol{G}^e ]$
and
$ [ \boldsymbol{H}^e ]$
in (A12) are expressed as

A.1.2. The material velocity contribution
It is straightforward to determine the contributions to the residual vector from (2.27). We take advantage of the symmetry of
$ \sigma ^{{\check {\alpha }} {\check {\beta }}}$
and
$ M^{{\check {\alpha }} {\check {\beta }}}$
to obtain

In (A15), it is understood that the
$ x$
,
$ y$
and
$ z$
components of the force terms are written to appropriate entries of the residual vector.
A.1.3. The Eulerian mesh velocity contribution
For the choice of an Eulerian mesh motion, the residual vector contribution corresponding to (2.33) is given by

A.1.4. The ALE mesh pressure contribution
When an ALE mesh motion is employed, the mesh pressure ensures the kinematic constraint (2.6) is satisfied. With the Dohrmann–Bochev method applied once again, the residual vector resulting from (2.39) is given by (cf. (A12))

A.1.5. The ALE mesh velocity contribution
When an ALE mesh motion is employed, the residual vector contribution from the mesh velocity looks similar to that from the material velocity (cf. (A15)):

In (A18),
$ \sigma ^{{\check {\alpha }} {\check {\beta }}}_{{{m}}, \mkern 1mu {v}}$
and
$ M^{{\check {\alpha }} {\check {\beta }}}_{{{m}}, \mkern 1mu {v}}$
(respectively
$ \sigma ^{{\check {\alpha }} {\check {\beta }}}_{{{m}}, \mkern 1mu {vb}}$
and
$ M^{{\check {\alpha }} {\check {\beta }}}_{{{m}}, \mkern 1mu {vb}}$
) are substituted when the motion is ALE-viscous (respectively ALE-viscous-bending).
Appendix B. The static portion of a cylinder
Here we consider a static membrane patch that is a portion of a cylinder – for which

In (B1),
$ \theta$
and
$ z$
are the canonical cylindrical coordinates and
$ r_{\mkern -1mu {{c}}}$
is the cylinder radius. With our differential geometric formulation, we arbitrarily choose to parametrise the surface as

such that

It is then straightforward to determine (Sahu Reference Sahu2022, Chapter IX, § 1)

for which the mean and Gaussian curvatures are respectively given by

The couple-stress components
$ M^{{\check {\alpha }} {\check {\beta }}}$
and couple-free in-plane stress components
$ \sigma ^{{\check {\alpha }} {\check {\beta }}}$
are calculated via (2.15) and (2.16) as

and

where, for a static patch, there are no viscous stresses – for which
$ \pi ^{{\check {\alpha }} {\check {\beta }}} = 0$
. Here
$ \lambda _0$
denotes the constant surface tension, which is set by the balance of forces in the out-of-plane direction. With some additional calculations, we find the stress vectors
$ \boldsymbol{T}^{\check {\alpha }}$
to be given by

We now separately consider the top and bottom edges, where
$ \zeta ^{\check {2}}$
is fixed, and the left and right edges, where
$ \zeta ^{\check {1}}$
is fixed.
The top and bottom surfaces: at the top (
$ \zeta ^{\check {2}} = 1$
) and bottom (
$ \zeta ^{\check {2}} = 0$
) edges,
$ \boldsymbol{\nu } = \pm \boldsymbol{e}_z$
,
$ \nu ^{}_{\check {1}} = 0$
,
$ \nu ^{}_{\check {2}} = \pm \ell$
,
$ \boldsymbol{\tau } = \mp \boldsymbol{e}^{}_\theta$
,
$ \tau ^{}_{\check {1}} = \mp r_{\mkern -1mu {{c}}}$
and
$ \tau ^{}_{\check {2}} = 0$
. We thus determine (see § 2.3.1)

and

The sign of the moment in table 1 is opposite that of (B9) due to the difference in orientation of the unit normal with respect to the surface.
The left and right surfaces: at the left (
$ \zeta ^{\check {1}} = \alpha$
) and right (
$ \zeta ^{\check {1}} = 0$
) edges,
$ \boldsymbol{\nu } = \pm \boldsymbol{e}^{}_\theta$
,
$ \nu ^{}_{\check {1}} = \pm r_{\mkern -1mu {{c}}}$
,
$ \nu ^{}_{\check {2}} = 0$
,
$ \boldsymbol{\tau } = \pm \boldsymbol{e}_z$
,
$ \tau ^{}_{\check {1}} = 0$
and
$ \tau ^{}_{\check {2}} = \pm \ell$
. In the same manner, we calculate

and

The moment in (B11) again differs from the moment reported in table 1 due to the choice of normal vector.
Appendix C. The numerical calculation of the pull force
In this section we determine how to set the membrane velocity
$ \boldsymbol{v}$
to a desired value
$ \boldsymbol{v}^{}_{{p}}$
at the centre of the membrane patch. Given the use of non-interpolatory basis functions, there is no unique way to do so. We thus choose to set the membrane velocity over the entire finite element containing the point of interest. The pull force
$ \boldsymbol{\mathcal{F}}_{\! {{p}}} (t)$
resulting from the imposed displacement is calculated via variational arguments.
C.1. The inability to uniquely displace a single point on the membrane
To begin, at a chosen point
$ (\zeta _{{p}} \hspace {-3.3pt} {}^{{\check {1}}}, \zeta _{{p}} \hspace {-3.3pt} {}^{{\check {2}}})$
we intend for

where
$ \boldsymbol{v}^{}_{{p}} (t)$
is a known function. With the velocity discretisation in (A4), (C1) can be expressed as

where
$ \{ N^{e}_{\mkern -1mu k} (\zeta _{{p}} \hspace {-3.7pt} {}^{{\check {\alpha }}}) \}$
and
$ \{ \boldsymbol{v}^{e}_{\mkern -1mu k} \}$
are respectively the local basis functions and nodal velocities of the finite element
$ \varOmega ^e_{{p}}$
containing
$ \zeta _{{p}} \hspace {-3.7pt} {}^{{\check {\alpha }}}$
. Importantly, the basis function values are set by the choice of
$ \zeta _{{p}} \hspace {-3.7pt} {}^{{\check {\alpha }}}$
, and they are non-interpolatory (Piegl & Tiller Reference Piegl and Tiller1997) – and thus, all
$ \textit {nen}$
nodal velocities contribute to (C2). Since
$ \boldsymbol{v}^{}_{{p}} \in \mathbb{R}^3$
and
$ \boldsymbol{v}^{e}_{\mkern -1mu k} \in \mathbb{R}^3$
for all
$ k$
, (C2) can be understood as a system of three scalar equations involving
$ 3 \boldsymbol{\cdot }\textit {nen}$
scalar unknowns. As
$ \textit {nen} = ( \textit {poly} + 1)^2$
for B-splines of polynomial order poly (Piegl & Tiller Reference Piegl and Tiller1997), there is no unique way to specify the
$ \{ \boldsymbol{v}^{e}_{\mkern -1mu k} \}$
in (C2). In practice, we set

Given the properties of B-spline functions (Piegl & Tiller Reference Piegl and Tiller1997), (C3) results in a uniform prescribed velocity over the entire parametric element
$ \varOmega ^e_{{p}}$
– expressed as

C.2. The pull force when a finite element is uniformly displaced
In this section we calculate the force
$ \boldsymbol{\mathcal{F}}_{\! {{p}}}$
required to pull an entire finite element at a prescribed velocity as in (C4). From the strong form of the governing membrane equations (2.18), (2.19) we recognise the pull force is the area integral of the net body force per area
$ \boldsymbol{f}$
on the membrane – expressed as

Since the prescribed velocity in (C4) can be viewed as a constraint,
$ \boldsymbol{f}$
is then understood as the associated Lagrange multiplier field over the element
$ \varOmega ^e_{{p}}$
.
Our task now is to determine how to calculate
$ \boldsymbol{\mathcal{F}}_{\! {{p}}}$
numerically, which is accomplished by considering how the weak form would be modified if (C4) was not satisfied directly. The principle of virtual power would then necessitate the quantity

be subtracted from the direct Galerkin expression, where
$ \delta J^{}_{\! \varOmega } = \varDelta t \mkern 1mu J^{}_{\! \varOmega } \, \boldsymbol{a}^{{\check {\alpha }}} \boldsymbol{\cdot } \delta \boldsymbol{v}^{{{m}}}_{, {\check {\alpha }}}$
(see Appendix C.2.1 of Sahu et al. Reference Sahu, Omar and Mandadapu2020b
). At this point, the fundamental unknowns and arbitrary variations are discretised as

In (C7) we introduced the shorthand
$ [ \boldsymbol{N}^e_{{p}} ] := [ \boldsymbol{N}^e (\zeta ^{\check {\alpha }}) ]$
for all
$ \zeta ^{\check {\alpha }} \in \varOmega ^e_{{p}}$
. Upon substituting (C7) into the second line of (C6) and defining the elemental mass matrix
$ [ \boldsymbol{M}^e_{{\! p}} ]$
as

we find the quantity

is to be subtracted from the discretised weak Galerkin expression
$ \mathcal{G}_h$
. We now separately consider the portions of (C9) arising from variations in the pull force, material velocity and mesh velocity.
C.2.1. The pull force contribution
The portion of the direct Galerkin expression associated with the net force per unit area
$ \boldsymbol{f}$
can be expressed as

Importantly, the mass matrix
$ [ \boldsymbol{M}^e_{{\! p}} ]$
is invertible and the variation
$ [ \delta \boldsymbol{f}^e_{{\! p}}]$
is arbitrary, so the nodal velocity degrees of freedom are found to be given by

It is a well-known property of B-splines (Piegl & Tiller Reference Piegl and Tiller1997) that if the constraint
$ \boldsymbol{v} (\zeta ^{\check {\alpha }}, t) = \boldsymbol{v}^{}_{{p}} (t)$
is enforced over the entire element, then the unique solution for the nodal degrees of freedom is given by

By comparing (C11) and (C12), we recognise that

which will be useful in our subsequent developments.
C.2.2. The material velocity contribution
The material velocity portion of the discretised direct Galerkin expression
$ \mathcal{G}_h$
is given by

where
$ [ \delta \boldsymbol{v} ]$
is the global variation of the nodal velocities,
$ [ \delta \boldsymbol{v}^e_{{\! p}} ]$
is the velocity variation of the nen nodes associated with
$\varOmega ^e_{{p}}$
and
$ [ \boldsymbol{r}^{{v}} ]$
is the velocity portion of the global residual vector in the absence of a pull force. Let us imagine reordering the global velocity degrees of freedom such that those associated with
$ \varOmega ^e_{{p}}$
appear last, and those not associated with the pull force appear first. The global velocity unknowns, their arbitrary variation and residual vector in the absence of a pull force are respectively expressed as

Here, the subscript ‘
$\bar {{{p}}}$
’ is used to signify ‘not p,’ i.e. degrees of freedom not associated with the pulled element
$ \varOmega ^e_{{p}}$
. Substituting (C15) into (C14) yields

Since the arbitrary variations
$ [ \delta \boldsymbol{v}_{\bar {{{p}}}} ]$
and
$ [ \delta \boldsymbol{v}^e_{{\! p}} ]$
are independent of one another, (C16) requires that
$ [ \mkern 1mu \boldsymbol{r}^{{v}}_{\bar {{{p}}}} \mkern 1mu ] = [ \boldsymbol{0} ]$
– which is the equation one obtains when the nodal velocities over
$ \varOmega ^e_{{p}}$
are set directly. Additionally, the nodal body force degrees of freedom are found to be given by

By substituting (C7) and then (C17) into (C5), we calculate the pull force as

Finally, recognising the mass matrix is symmetric and substituting the transpose of (C13) into (C18) yields

which is straightforward to calculate within finite element subroutines. Equation (C19) is the main result of this section.
C.2.3. The mesh velocity contribution
The mesh velocity portion of
$ \mathcal{G}_h$
can be written as

for any arbitrary variation
$ [ \delta \boldsymbol{v}^{{m}} ]$
. By reordering global velocity degrees of freedom in the same manner as (C15), we find that

for all independent variations
$ [ \delta \boldsymbol{v}_{\bar {{{p}}}} ]$
and
$ [ \delta \boldsymbol{v}^e_{{\! p}} ]$
. Equation (C21) requires that
$ [ \mkern 1mu \boldsymbol{r}^{{m}}_{\bar {{{p}}}} \mkern 1mu ] = [ \boldsymbol{0} ]$
, which solves for all mesh velocity degrees of freedom not associated with the finite element
$ \varOmega ^e_{{p}}$
. In addition, we find that

where
$ [ \boldsymbol{r}^{{{m}}, e}_{{p}} ]$
is the mesh velocity portion of the residual vector corresponding to
$ \varOmega ^e_{{p}}$
in the absence of a pull force. The following discussion explains why (C22) is not used in our code.
C.3. The numerical implementation
In our numerical implementation we calculate the pull force
$ \boldsymbol{\mathcal{F}}_{\! {{p}}} (t)$
directly – rather than with a Lagrange multiplier field. To do so, we set the
$ \textit {nen}$
nodal material velocity values
$ \{ \boldsymbol{v}^{e}_{\mkern -1mu k} (t) \}$
associated with
$ \varOmega ^e_{{p}}$
directly according to (C12). Though these nodes are removed from the degree-of-freedom list, the residual
$ [ \boldsymbol{r}^{{{v}}, e}_{{p}} ]$
is still calculated, from which the pull force is determined according to (C19). We also directly set the
$ \textit {nen}$
mesh velocity degrees of freedom
$ \{ \boldsymbol{v}^{{{m}}, e}_{\mkern -1mu k} (t) \}$
associated with
$ \varOmega ^e_{{p}}$
to be
$ \boldsymbol{v}^{}_{{p}} (t)$
in a similar fashion. Since the nodal mesh velocities on
$ \varOmega ^e_{{p}}$
are known, they are removed from the degree-of-freedom list, and so the integral term in (C22) is not evaluated in practice.
Appendix D. The pull force at small deformations
Consider a membrane patch which, prior to any deformation, is at the constant surface tension
$ \lambda _0$
and spans the region between two concentric circles in the
$ x$
–
$ y$
plane. We denote
$ \ell$
as the diameter of the outer circle and
$ 2 \mkern 1mu r_{\!p}$
as the diameter of the inner circle. Eventually, we will take the limit as
$ r_{\!p} \rightarrow 0$
. When a deformation is applied quasi-statically and the membrane height
$ h = h (r, \theta )$
above the
$ x$
–
$ y$
plane is small (i.e.
$ h \ll \ell$
), the membrane shape is known to satisfy (Sahu Reference Sahu2022, Chapter VII)

where

is the 2-D Laplacian expressed in terms of the canonical cylindrical coordinates
$ r$
and
$ \theta$
. In what follows, all lengths are non-dimensionalised by
$ \ell / 2$
, for which

With (4.12) and (D3), the shape equation (D1) is presented in dimensionless form as

In comparing (D4) with the description in Powers et al. (Reference Powers, Huber and Goldstein2002), we recognise the small parameter
$ \epsilon$
in the latter is given by
$ 2 / {\varGamma }$
in our notation. Thus, for an axisymmetric membrane with (i) prescribed displacement
$ z^{}_{{p}}$
at
$ r = r_{\!p}$
, (ii) zero slope at
$ r = r_{\!p}$
, (iii) no displacement at
$ r = \ell / 2$
, and (iv) no moment at
$ r = \ell / 2$
, we reproduce the Powers et al. (Reference Powers, Huber and Goldstein2002) solution as

where

is a constant parameter defined for notational convenience. Note that in our numerics, we require zero slope rather than zero moment on the outer boundary. However, when membrane deformations are small, we do not expect this choice of boundary condition to affect the pull force calculation. In (D5) and (D6),
$ K_0$
and
$ K_1$
are modified Bessel functions of the second kind, and
$ z^{*}_{{{p}}} := 2 \mkern 1mu z^{}_{{p}} / \ell$
is the dimensionless displacement at the inner membrane boundary – which is located at
$ r^* = r^{*}_{\mkern -1mu {{p}}} := 2 \mkern 1mu r_{\!p} / \ell$
.
With the solution for the membrane shape at a given displacement in (D5), we seek to determine the magnitude of the pull force in the
$ \boldsymbol{e}_z$
direction:
$ \mathcal{F}_{\! {{p}}, \mkern 1mu z}$
. The pull force is related to
$ F_z$
, the force per length in the
$ \boldsymbol{e}_z$
direction at the inner boundary, via

Since the membrane patch is parallel to the
$ x$
–
$ y$
plane at the inner boundary, we recognise that (Sahu Reference Sahu2022, Chapter V, § 6(e))

where the mean curvature
$ H = ({1}/{2}) \mkern 1mu {\nabla} ^2 h$
in the limit of small deformations. Moreover, since the membrane height is axisymmetric and has zero slope at the inner boundary, we apply (D2) and find that

With the height solution in (D5) and properties of Bessel functions (Abramowitz & Stegun Reference Abramowitz and Stegun1964), the terms in parenthesis in (D9) are found to be

and

Substituting (D10) and (D11) into (D9) yields

At this point, we evaluate the limit by recognising
$ \alpha \rightarrow 0$
as
$ r^{*}_{\mkern -1mu {{p}}} \rightarrow 0$
. In addition, we use the Bessel function relations (Abramowitz & Stegun Reference Abramowitz and Stegun1964)

By substituting (4.14), (D6) and (D13) into (D12) and rearranging terms, we obtain the small-deformation pull force expression in (4.20) – presented here as


Figure 10. Tether translation in the
$ \boldsymbol{e}_x$
direction with a Lagrangian mesh motion. In the left column, simulation boundary conditions are those specified in the first two rows of table 3. Twisting and tilting of the tether, along with further striation of the surface tension, ensue. In the right column, an attempt is made to remove unphysical constraints on the membrane by only pinning the centre of the bottom boundary. Upon lateral pulling, the mesh rotates about the centre of the bottom edge; twisting and tension striation are once again observed. Both simulations fail to capture the expected behaviour shown in the left column of figure 9. As in the main text, times are measured in units of
$ \zeta \mkern 1mu r_{\mkern -1mu {{c}}}^{\, 2} / k_{{b}}$
, the pull velocity is specified in the first row of (4.22), and all other parameters are identical to those in figure 4. The vertical red lines are a visual aid to highlight the motion of the tether.
Appendix E. The translation of a tether with a Lagrangian mesh motion
When a pulled tether is translated laterally across a membrane surface, lipids in the surrounding patch flow in-plane to accommodate the large out-of-plane shape deformations. For a given prescribed lateral tether velocity as in (4.22), we do not know the corresponding velocity field
$ \boldsymbol{v} (\zeta ^{\check {\alpha }}, t)$
over the membrane patch a priori. Thus, any boundary conditions that pin certain nodes yield unphysical results, as shown in figure 10. If no nodes are pinned, however, a rigid body translation results – and the dynamics of lateral tether motion are not obtained. Lagrangian simulations are thus unable to capture the physics of tether translation.