Zum Inhalt
Erschienen in:

Open Access 07.08.2024 | Research

A variational model for finger-driven cell diffusion in the extracellular matrix

verfasst von: Antonino Favata, Andrea Rodella, Stefano Vidoli

Erschienen in: Meccanica | Ausgabe 8/2024

Aktivieren Sie unsere intelligente Suche, um passende Fachinhalte oder Patente zu finden.

search-config
loading …

Abstract

Der Artikel geht auf das komplizierte Zusammenspiel zwischen Zellen und der extrazellulären Matrix (ECM) ein und betont die Rolle der mechanischen Eigenschaften der ECM bei der Zellmigration. Es wird ein Variationsmodell eingeführt, das die Entstehung von Anisotropie und Mikroknickungen im ECM erfasst, was zu lokalisierten Feldmustern führt, die an experimentelle Ergebnisse erinnern. Das Modell berücksichtigt auch die Zelldiffusion und zeigt, wie sich Zellen als Reaktion auf diese mechanischen Veränderungen im Gewebe ausbreiten. Die Studie beleuchtet die Dynamik faserhaltiger Materialien und ihre zentrale Rolle bei zellulären Prozessen und bietet Einblicke in die Gewebetechnik, die regenerative Medizin und die Krebsforschung.
Hinweise
Antonino Favata, Andrea Rodella and Stefano Vidoli have contributed equally to this work.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

1 Introduction

Cells spreading within tissues, such as cancer metastases or fibroblasts, engage in interactions with both the cellular matrix, and other cells. In recent times, considerable attention has been focused on elucidating the mechanics of cell movement resulting from their interactions with the extracellular matrix (ECM). In [1012, 33] an in-depth description of the physiological mechanisms underlying these motions is provided.
The mobility of cells is influenced by a multifaceted interplay of factors, encompassing not only biochemical signaling within the cell and subsequent cytoskeletal rearrangements, but also the biophysical attributes of the surrounding environment. Central to these environmental factors is the ECM, made of intricate network of insoluble structural fibrous proteins like collagen type I, glycoproteins, and soluble glycosaminoglycans. Collectively, these components serve as microstructural guidance cues and biochemical stimuli, influencing the movement of individual entities.
Cell migration occurs through two main mechanisms. The first mode is characterized as mesenchymal migration, where cells penetrate the matrix by generating a pathway around them, relying on proteolysis and actin polymerization. The second mode involves amoeboid migration, where cells move through any available pathway by squeezing through, utilizing membrane protrusions known as blebs. We mainly focus on the former mechanism, although in a very simplified framework. In this kind of migration, cells degrade the surrounding matrix by breaking it down after creating protrusions, subsequently remodeling it along their trajectory; in particular cells exhibit the ability to alter the orientation of collagen fibers, aligning them in the direction of their motion, through a mechanism called proteolysis [18].
Significantly, the ECM undergoes considerable remodeling, profoundly impacting cell spreading, motility, and differentiation. Fiber reorientation within the ECM plays a pivotal role in guiding the migration of both normal and cancer cells. Understanding the mechanical behavior of the ECM is paramount for deciphering the intricate interplay between fibrous materials and cellular responses.
The mechanical remodeling of the ECM is marked by noteworthy spatial patterns of densification and fiber alignment, localized within finger-like bands connecting distant cell clusters. Remarkably, cancer cells display a preference for invading regions of the ECM with higher density and greater fiber alignment, with specific ECM density and alignment patterns identified as biomarkers for breast cancer [6].
Cell-induced forces bring about significant changes in ECM structure. Along the axis of each tether, the ECM undergoes substantial straining, with deformations reaching up to \(40\%\) of its original length [14]. Conversely, in the transverse direction, the ECM undergoes compression, resulting in a reduction of thickness to as little as half or even a quarter of its initial value. The unexpected localization of densification within the tethers highlights the intricate interplay between cellular forces and ECM mechanics [14].
Fig. 1
(a) Illustration of a cell cluster surrounded by randomly oriented collagen fibers, representing the extracellular matrix. (b) The cluster, depicted by the circle, exerts inward pressure p on the surroundings leading to compression (and eventual buckling) of fibers aligned in the hoop direction
Fig. 1 visually represents these processes, displaying a cartoon of a cell cluster surrounded by ECM, illustrating the contraction and deformation of the ECM, summarizing the experimental findings from [14]. In this work, it is posited that tether formation is induced by the buckling instability of network fibers under cell-induced compression.
The association with instability, as discussed in [8, 14], was initially established through experiments and modeling on fiber networks, open-cell foams, and fibrin [19, 20]. Individual fibers offer support against tensile forces and may become stiffer as tension increases [17, 32]. However, when subjected to compression, these fibers buckle, leading to a loss of stiffness. These microscopic buckling instabilities give rise to the formation of macroscopic bands characterized by intense compressive deformation and high density: within these bands, the fibers are predominantly buckled and compacted. These dense regions alternate with areas of standard density and minimal compressive strain, where the fibers are generally straight and loosely organized [14].
The observed mechanical behavior of the ECM during cellular remodeling underscores the dynamic nature of fibrous materials and their pivotal role in cellular processes. Understanding the mechanisms underpinning these intricate deformation patterns and their implications for cell behavior holds significant promise across various fields, including tissue engineering, regenerative medicine, and cancer research.
In this work, we present a simple variational formulation of a chemo-mechanical model within a phase-field framework, elucidating fundamental aspects of the phenomena mentioned previously. This includes (i) the emergence of anisotropy through a macroscopic measure of fiber stiffness in the direction of transverse anisotropy, (ii) the manifestation of micro-buckling leading to material softening, (iii) the emergence of localized field patterns reminiscent of experimental findings on finger-like structures formation, and (iv) a diffusion of a phase representing cells that spread around the tissue. To maintain clarity in fundamental concepts without introducing unnecessary complexities, we deliberately refrain from incorporating large deformations. Upon reaching a critical strain, an abrupt jump in the response results in the emergence of a sudden anisotropy. Physically, we interpret this as microscopic-scale buckling. In terms of modeling, this categorizes our model alongside materials with a domain of reversibility, resembling those with an elastic range.
In our prior work [8], we addressed a comparable issue, emphasizing solely mechanical aspects while overlooking cell diffusion and its interplay with the matrix. Specifically, we examined the manifestation of localized anisotropy in processes involving loading and unloading, and introduced a memory effect to depict cyclic behavior. In this study, we solely focus on the loading phase, yet integrate chemical diffusion into the analysis. We finally notice that, despite out previous model, which was rate-independent, in the present contribution rate-independent processes are coupled to the rate-dependent one of cell diffusion; this requires a different framework, although governing equations of the dissipative behavior are also proposed in a variational form, using an energetic formulation of rate-dependent materials [2326].
The paper is organized as follows. In Sec. 2 we describe the elastic energy and the material parameters of a transversely isotropic material, with the aim to elucidate the transition from an initially isotropic material to one with transverse isotropy. In Sec. 3 we introduce the relevant fields entering in our formulation, lay down the general balance equations and the thermodynamic restrictions, propose a variational formulation of the evolution equations; therefore, we suggest specific constitutive choices and derive the governing equations. In Sec. 4 some numerical results are shown and discussed.

2 From isotropy to buckling-induced anisotropy in the elastic response

As detailed in the introduction, due to micro-buckling, the ECM suddenly undergoes a transition from isotropy to anisotropy. In this section, we delve into understanding how the stiffness of a plane transversely isotropic material depends on its four key material constants: E (Young modulus), \(\nu \) (Poisson coefficient), \(c_1\), and \(c_{2}\) (transversely isotropic constants), as discussed in [7, 30]. Our focus is on highlighting a material that not only exhibits a softer response in a specific direction but also maintains consistent stiffness in the orthogonal direction. The primary objective is to elucidate the transformation from an initially isotropic material to one with transverse isotropy, offering insights into admissible anisotropic material parameters within the space ensuring the positivity of the elastic energy. In this realm, we opt for a trajectory that highlights a decrease in stiffness aligned with transverse isotropy. For more in-depth information, readers are referred to [7].
The elastic energy density is assumed to be
$$\psi _{\textrm{e}}=\widehat{\psi }_{\textrm{e}}({\textbf {E}}({\textbf {u}}), \vartheta )=\frac{1}{2} \mathbb {C}({\textbf {n}}(\vartheta )){\textbf {E}}({\textbf {u}})\cdot {\textbf {E}}({\textbf {u}}),$$
where \({\textbf {E}}({\textbf {u}})=\text {sym\,}\nabla {{\textbf {u}}}\) is the linearized strain measure expressed in terms of the displacement field \({\textbf {u}}\) and \(\mathbb {C}\big ({\textbf {n}}(\vartheta )\big )\) is the elasticity tensor of a linearly elastic transversely isotropic material, with respect to the unit direction \({\textbf {n}}(\vartheta )\), parametrized by the angle \(\vartheta \). This elasticity tensor, in a two-dimensional setting, can be represented through
$$\begin{aligned}\widehat{\psi }_{\textrm{e}}\Big ({\textbf {E}}({\textbf {u}}),\vartheta \Big )&=\dfrac{E}{2\, (1-\nu ^2)}\,\left[ (1-\nu )\, |{\textbf {E}}({\textbf {u}})|^2 +\nu (\text{ tr }\, {\textbf {E}}({\textbf {u}}))^2\right] +\nonumber \\& \quad +c_1 \Big (\text{ tr }\, {\textbf {E}}({\textbf {u}})\Big )\,{\textbf {E}}({\textbf {u}}){\textbf {n}}(\vartheta ) \cdot {\textbf {n}}(\vartheta ) +c_2 \,\Big ({\textbf {E}}({\textbf {u}})\,{\textbf {n}}(\vartheta )\cdot {\textbf {n}}(\vartheta )\Big )^2 \end{aligned}$$
(1)
where the first term on the right-hand-side represents the elastic energy of an isotropic material, while the elastic moduli \(c_1\) and \(c_2\) modulate the anisotropic response. We define a softening material when the ratio between the Young modulus tested along the parallel and orthogonal directions with respect to the fiber \({\textbf {n}}(\vartheta )\) is less than 1:
$$\begin{aligned} \frac{E_\vartheta ^\parallel }{E_\vartheta ^\perp } = 1+2(c_1+c_2)\,\frac{1-\nu ^2}{E}<1 \end{aligned}$$
(2)
Fig. 2
Polar plot of the Young modulus for \(0\le c_2 \le -c_{2m}\). In yellow the admissibility region in the plane \((c_1/E_0,c_2/E_0)\)
To match the goal of this section, we can consider a material with \(c_1=0\). Indeed, the mechanical effect of the constant \(c_2\) is visualized in polar plots of the Young modulus (see Fig. 2) as a function of the angle between the testing and fiber directions. To ensure the positiveness of the energy density (1) the constants \(c_1, c_2\) must belong to the yellow region represented in Fig. 2 where
$$\begin{aligned} &\\& \quad E_0 :=E/(1-\nu ^2)>0, &\\& -c_{2m}:=-E_0(1-\nu ^2)\le c_2, &\\& -1<\nu <1/2.\end{aligned}$$
One can nondimensionalize the constant \(c_2\) by defining \(\alpha :=-c_2/c_{2m}\) with \(0\le \alpha \le 1\), consequently the elastic energy density takes the form:
$$\begin{aligned} \widehat{\psi }_\textrm{e}\Big ({\textbf {E}}({\textbf {u}}),\vartheta \Big )=\dfrac{E_0}{2}\,\left[ (1-\nu )\, |{\textbf {E}}({\textbf {u}})|^2 +\nu (\text{ tr }\, {\textbf {E}}({\textbf {u}}))^2 - \alpha \left( {\textbf {E}}{\textbf {n}}\cdot {\textbf {n}}\right) ^2 \right] \end{aligned}$$
(3)
It is worth noticing, with the aid of Fig. 2, that while \(\alpha \) tends to 1 the stiffness in the fiber direction \({\textbf {n}}(\vartheta )\) tends to 0.
Looking at the bigger picture, the decrease in stiffness along the direction of buckled fibers reveals micro-buckling at the macroscopic level. Consequently, following our previous contribution [8], we suggest interpreting \(\alpha \) in (3) not as a constant material parameter but as a field undergoing evolution.

3 Energy functional for finger-driven cell diffusion

We suppose that cells are able to diffuse within the matrix, and denote by c the cell density field, so that \( \int _\Pi c \) represents the net mass of the diffusing cells in an arbitrary part \(\Pi \subset \Omega \) (at time t). We further introduce the vectorial species flux \({\textbf {j}}\). The cell mass balance is the requirement that
$$\begin{aligned} \frac{\text {d}}{\text {d}t}\left( \int _\Pi c \right) = -\int _{\partial \Pi }{\textbf {j}}\cdot \varvec{\nu }, \end{aligned}$$
(4)
for every region \(\Pi \subset \Omega \). The first term in the right-hand side of (4) gives the rate at which the cells are transported to \(\Pi \) by diffusion across \(\partial \Pi \), being \(\varvec{\nu }\) the outward unit normal; the minus sign renders this term non-negative when the flux \({\textbf {j}}\) points into \(\Pi \). Eq. (4) yields the local cells mass-balance
$$\begin{aligned} \dot{c} = - \text {div}{\textbf {j}}. \end{aligned}$$
(5)
Here and henceforth a superimposed dot denotes a time derivative.
Drawing on [5, 13, 16], our proposed model relies on internal state variables that characterize the evolution of the system, which depends on two fundamental energy aspects: an internal energy density and a dissipation potential. Let \(\mathbb {S}=\{{\textbf {E}}({\textbf {u}}), \vartheta , \alpha , c, \nabla {\alpha }\}\) be the list of the state variables:
  • \({\textbf {E}}({\textbf {u}})\) is the infinitesimal strain tensor associated to the displacement field \({\textbf {u}}\), as already defined.
  • \(\vartheta \) is the fiber direction field. The unit vector \({\textbf {n}}(\vartheta )\), as detailed in Sect. 2, denotes the direction of fibers that can rearrange themselves, leading to remodeling. More precisely, it indicates the direction in which anisotropy manifests, or rather better, the direction of micro-buckling occurrence.
  • \(\alpha \) acts as the anisotropy gauge field, with its magnitude capable of punctually increasing from 0, indicating a locally isotropic material, to 1, signifying a transversely isotropic material with virtually zero stiffness in the fiber direction \({\textbf {n}}(\vartheta )\). Its gradient, \(\nabla {\alpha }\), influences the behavior of the anisotropy gauge field as regularizing field, an essential ingredient for capturing the formation of finger-like bands within the material.
We consider the second law of thermodynamics as an a priori constraint, expressed for isothermal processes through the Clausius-Duhem inequality:
$$\begin{aligned} \delta :={\textbf {S}}\cdot \dot{{\textbf {E}}}+\mu \dot{c}-{\textbf {j}}\cdot \nabla \mu -\dot{\psi }\ge 0. \end{aligned}$$
(6)
Here \(\psi \) is the Helmholtz free energy density, \({\textbf {S}}\) is the Cauchy stress tensor and \(\mu \) represents the chemical potential, allowing cells to migrate from the boundary into the domain. In this context, we assume chemical equilibrium between cells and the substrate.
Exploiting the dissipation inequality within the framework of materials with internal variables needs some caution, because they are a priori constrained to obey a constitutive evolution law, as detailed in the seminal work by Coleman and Gurtin [5]. More specifically, the material, at each point, is characterized by the response functions
$$\begin{aligned} \begin{aligned} \psi =\widehat{\psi }(\mathbb {S}), \quad {\textbf {S}}=\widehat{{\textbf {S}}}(\mathbb {S}), \quad \mu =\widehat{\mu }(\mathbb {S}), \quad \dot{\vartheta }=\mathfrak {g}(\mathbb {S}), \quad \dot{\alpha }=\mathfrak {f}(\mathbb {S}). \end{aligned} \end{aligned}$$
(7)
Since \(\dot{\psi }=\partial _\mathbb {S}\widehat{\psi }\cdot \dot{\mathbb {S}}\), on requiring that (6) be satisfied whatever the local continuation of any conceivable process, one easily finds the following standard constitutive relations
$$\begin{aligned} \begin{aligned} {\textbf {S}}=\widehat{{\textbf {S}}}(\mathbb {S})=\partial _{{\textbf {E}}}\widehat{\psi }(\mathbb {S}), \qquad \mu =\widehat{\mu }(\mathbb {S})=\partial _c \widehat{\psi }(\mathbb {S}), \end{aligned} \end{aligned}$$
(8)
for the stress and the chemical potential. The presence of \(\nabla \alpha \) in the list of state variables produces some difficulties, as detailed in [22]. Indeed, the presence of a non-local term would likely affect the entropy production, and then the dissipation; more specifically, according to [22], an extra entropy flux is in order, which would result in the absorption of some terms in the reduced dissipation inequality. In particular, the term \(\partial _{\nabla \alpha }\widehat{\psi }(\mathbb {S})\cdot \nabla \dot{\alpha }\) appearing when evaluating \(\dot{\psi }\) in (6) kind of evaporate according to this reasoning (see [22] for more details). Now it remains to discuss the residual dissipation related to the variation of \(\vartheta \), \(\alpha \), and due to \({\textbf {j}}\). As to \(\vartheta \) we choose not to associate a dissipation. As to the latter, it is customarily accepted that the dissipation rate \(\delta \) reveals different entropic effects having distinct origins, so that the positiveness of the dissipation rate is assumed separately for the local and diffusive contributions (see also [22]). For this reason, we assume that the following relations hold true, which are stronger than (6) and not its direct consequence:
$$\begin{aligned} \delta _{\textrm{loc}}\ge 0, \quad \delta _{\textrm{diff}}\ge 0, \qquad \delta _{\textrm{diff}}:=-{\textbf {j}}\cdot \nabla \mu , \quad \delta _{\textrm{loc}}:=\delta -\delta _{\textrm{diff}}. \end{aligned}$$
(9)
or rather
$$\begin{aligned} -\partial _{\alpha }\widehat{\psi }(\mathbb {S})\dot{\alpha }\ge 0, \qquad -{\textbf {j}}\cdot \nabla \mu \ge 0, \end{aligned}$$
(10)
where the flux \({\textbf {j}}\) is in need of further constitutive prescriptions.
In order to ensure the inequality (10)\(_2\), we introduce a dissipation potential \(\varphi _{\textrm{diff}}\), a 2-homogeneous convex function of the flux \({\textbf {j}}\), such that
$$\begin{aligned} \delta _{\textrm{diff}}=\partial _{{\textbf {j}}}\varphi _{\textrm{diff}}(\mathbb {S}; {\textbf {j}})\cdot {\textbf {j}}\ge 0. \end{aligned}$$
(11)
For the following purposes, we also introduce the dual diffusive dissipation potential, a 2-convex homogeneous function of \(\nabla \mu \), as a suitable Legendre transform, such that
$$\varphi _{\textrm{diff}}( \mathbb {S}; {\textbf {j}})=\sup _{\nabla \mu } \{{\textbf {j}}\cdot \nabla \mu -\varphi ^\star _{\textrm{diff}}( \mathbb {S}; \nabla \mu ) \}.$$
With this, we establish the connection between the cell flux and the gradient of the chemical potential, expressed as:
$$\begin{aligned} {\textbf {j}}=\partial _{\nabla \mu }\varphi ^\star _{\textrm{diff}}( \mathbb {S}, \nabla \mu )=-{\textbf {M}}(\mathbb {S})\nabla \mu , \end{aligned}$$
(12)
where the positive semidefinite second order \({\textbf {M}}(\mathbb {S})\) represents the mobility tensor that, for the time being, we admit being dependent on all the state list \(\mathbb {S}\). Eq. (12) is a generalized version of the classical Fick’s law. A specific dependence of the mobility tensor on the state variable list \(\mathbb {S}\) will be presented in the following.

3.1 Variational formulation of the evolution equations

In this Section, we aim to introduce a variational formulation of the problem at hand. For this purpose, we use an energetic formulation building on [2326]. Let us introduce the stored energy functional
$$\begin{aligned} \mathscr {E}({\textbf {u}}, \vartheta , c, \alpha ):=\int _{\Omega }\psi (\mathbb {S}), \end{aligned}$$
(13)
and the total external power, given by the sum of two contributions:
$$\begin{aligned} \mathscr {P}(\dot{{\textbf {u}}})=\mathscr {P}_{\textrm{m}}(\dot{{\textbf {u}}})+\mathscr {P}_{\textrm{c}}({\textbf {j}}), \end{aligned}$$
(14)
the mechanical power of external loads as
$$\begin{aligned} \mathscr {P}_{\textrm{m}}(\dot{{\textbf {u}}}):=\int _{\partial \Omega }{\textbf {t}}\cdot \dot{{\textbf {u}}}, \end{aligned}$$
(15)
and the chemical external power:
$$\begin{aligned} \mathscr {P}_{\textrm{c}}( {\textbf {j}}):=-\int _{\partial \Omega }\mu \,{\textbf {j}}\cdot \varvec{\nu }. \end{aligned}$$
(16)
We notice that the chemical external power can be considered as a functional of \(\dot{c}\), if the cells-mass balance (5) considered.1 With a slight abuse of notation we therefore denote the chemical external power functional as \(\mathscr {P}_{\textrm{c}}( \dot{c})\) and the total external power functional as \(\mathscr {P}( \dot{{\textbf {u}}}, \dot{c})\).
We acknowledge that dissipative processes can be characterized by the dissipation power functional, which divides into two parts, in accordance with the splitting given in eq. (9). Specifically, we introduce the local dissipation power functional
$$\begin{aligned} \mathscr {D}_{\textrm{loc}}({\textbf {u}}, \vartheta , c, \alpha ; \dot{\alpha })=\int _{\Omega }\varphi _{\textrm{loc}}(\mathbb {S}; \dot{\alpha },\nabla \dot{\alpha }), \end{aligned}$$
(17)
and a diffusive dissipation power functional, playing a pivotal role in formulating rate-type variational principles for the coupled deformation-diffusion problem. We assume that the diffusive dissipation potential functional is defined through the following variational principle [25]:
$$\begin{aligned} \mathscr {D}_{\textrm{diff}}( {\textbf {u}}, \vartheta , c, \alpha ; \dot{c})=\sup _{\mu \in H^1(\Omega )}\left\{ \int _\Omega -\mu \dot{c}-\varphi ^\star _{\textrm{diff}}(\mathbb {S}; \nabla \mu ), \right\} . \end{aligned}$$
(18)
The Euler-Lagrange equation of the optimization problem (18) determines the evolution of the cell mass concentration field c:
$$\begin{aligned} \dot{c}=\text {div}\big ( {\textbf {M}}(\mathbb {S})\nabla \mu \big ). \end{aligned}$$
(19)
We introduce the following functional spaces for the state variable fields and the corresponding variations:
$$\begin{aligned}&\mathscr {U}=\left\{ {\textbf {u}}\in H^1(\Omega ,\mathbb {R}^3)\,:\,{\textbf {u}}={\textbf {0}}\;\;\textrm{on}\;\;\partial \Omega \right\} ,\quad \mathscr {T}= L^2(\Omega ,[-\pi /2,\pi /2]), \quad \mathscr {A}=H^1(\Omega ,[0,1]),\\&\mathscr {C}=\left\{ c \in H^1(\Omega )\,:\,c=c_0 \;\;\textrm{on}\;\;\partial \Omega \right\} \\&\widetilde{\mathscr {U}}=\left\{ {\textbf {v}}\in H^1(\Omega ,\mathbb {R}^3)\,:\,{\textbf {v}}={\textbf {0}} \;\;\textrm{on}\;\;\partial \Omega \right\} ,\quad \widetilde{\mathscr {T}}=L^2(\Omega ,\mathbb {R}), \quad \widetilde{\mathscr {A}}=H^1(\Omega ,\mathbb {R}),\\&\widetilde{\mathscr {C}}=\left\{ \tilde{c} \in H^1(\Omega )\,:\,\tilde{c}=0 \;\;\textrm{on}\;\;\partial \Omega \right\} . \end{aligned}$$
Based on the energy and the dissipation power functionals \(\mathscr {E}\), \(\mathscr {D}_{\textrm{loc}}\), and \(\mathscr {D}_{\textrm{diff}}\), at a given state \(\mathbb {S}\) at time t, we postulate the existence of the rate-type potential
$$\begin{aligned}\underbrace{\mathscr {L}({\textbf {u}}, \vartheta , c, \alpha ; \dot{{\textbf {u}}}, \dot{\vartheta }, \dot{\alpha }, \dot{c})}_{\text {rate potential}}&:=\underbrace{\frac{\text {d}}{\text {d}t} \mathscr {E}({\textbf {u}}, \vartheta , c, \alpha )}_{\text {rate of energy}} \\ +\underbrace{\mathscr {D}_{\textrm{loc}}({\textbf {u}}, \vartheta , c, \alpha ; \dot{\alpha })}_{\text {local dissipation power}}\nonumber & +\underbrace{\mathscr {D}_{\textrm{diff}}( {\textbf {u}}, \vartheta , c, \alpha ; \dot{c})}_{\text {diffusive dissipation power}} \\& -\underbrace{\mathscr {P}(\dot{{\textbf {u}}},\dot{c})}_{\text {external power}} \end{aligned}$$
(20)
The evolution problem can be determined as a rate-type stationarity problem [3, 2325]:
$$\begin{aligned} \inf _{\dot{{\textbf {u}}}\in \widetilde{\mathscr {U}}}\inf _{\dot{\vartheta }\in \widetilde{\mathscr {T}}}\inf _{\dot{\alpha }\in \widetilde{\mathscr {A}}}\inf _{\dot{c}\in \widetilde{\mathscr {C}}}\,\mathscr {L}({\textbf {u}}, \vartheta , c, \alpha ; \dot{{{\textbf {u}}}}, \dot{\vartheta },\dot{\alpha }, \dot{c}). \end{aligned}$$
(21)
This principle yields the force balance equation, and the evolution laws for the cell mass, the fiber orientation and the anisotropy gauge field.
Instead of c, we find it convenient to work with the chemical potential \(\mu \). The main reason is that we consider the cell cluster and the ECM in chemical equilibrium, and therefore is more natural using this field as primary variable. Moreover, a major challenge associated with the minimization principle (21) lies in the fact that \(\mathscr {D}_{\textrm{diff}}( {\textbf {u}}, \vartheta , c, \alpha ; \dot{c})\) is defined through the Legendre transform (18). An explicit representation of \(\mathscr {D}_{\textrm{diff}}\), which would be necessary for obtaining the governing laws, is achievable, but it is not straightforward.2 We therefore adopt a different formulation, and introduce the grand-canonical potential [15]
$$\begin{aligned} \omega =\widehat{\omega }(\mathbb {S}^\star )=\widehat{\psi }(\mathbb {S})-\mu \,\widehat{c}(\mathbb {S}^\star ), \end{aligned}$$
(22)
with \(\mathbb {S}^\star =\{{\textbf {E}}({\textbf {u}}), \vartheta ,\mu , \alpha , \nabla \alpha \}\) and \(\widehat{c}(\mathbb {S}^\star )=c\) the constitutive response map for the concentration, written in terms of \(\mathbb {S}^\star \). This leads to the following constitutive equations for the stress and the cell mass concentration:
$$\begin{aligned} {\textbf {S}}=\widetilde{{\textbf {S}}}(\mathbb {S}^\star )=\partial _{{\textbf {E}}}\widehat{\omega }(\mathbb {S}^\star ), \quad c=\widehat{c}(\mathbb {S}^\star )=-\partial _\mu \widehat{\omega }(\mathbb {S}^\star ). \end{aligned}$$
(23)
Moreover, we introduce the extended3 dissipation potential functional
$$\begin{aligned} \mathscr {D}^\star _{\textrm{diff}}({\textbf {u}}, \vartheta , c, \alpha ; \dot{c}, \mu ):=\int _\Omega -\mu \dot{c}-\varphi ^\star _{\textrm{diff}}(\mathbb {S}; \nabla \mu ). \end{aligned}$$
(24)
We notice that this functional encompasses the constitutive dissipation potential function \(\varphi ^\star \), responsible for regulating the diffusion process.
The principle (21) can be therefore replaced by the following rate-type saddle problem (see [24]):
$$\begin{aligned} \inf _{\dot{{\textbf {u}}}\in \widetilde{\mathscr {U}}}\inf _{\dot{\vartheta }\in \widetilde{\mathscr {T}}}\inf _{\dot{\alpha }\in \widetilde{\mathscr {A}}}\inf _{\dot{c}\in \widetilde{\mathscr {C}}}\sup _{\mu \in \mathscr {M}}\,\mathscr {L}^\star ({\textbf {u}}, \vartheta , c, \alpha , \mu ; \dot{{\textbf {u}}}, \dot{\vartheta }, \dot{\alpha }, \dot{c}), \end{aligned}$$
(25)
where
$$\begin{aligned}\mathscr {L}^\star ({\textbf {u}}, \vartheta , c, \alpha , \mu ; \dot{{\textbf {u}}}, \dot{\alpha }, \dot{c})&:=\frac{\text {d}}{\text {d}t} \mathscr {E}^\star ({\textbf {u}}, \vartheta , \mu , \alpha )\nonumber \\&\quad+\mathscr {D}_\textrm{loc}( {\textbf {u}}, \vartheta , c, \alpha ; \dot{\alpha })\\&\quad+\mathscr {D}^\star _\textrm{diff}({\textbf {u}}, \vartheta , c, \alpha ; \dot{c}, \mu )\\&\quad -\mathscr {P}^\star (\dot{{\textbf {u}}}, \mu ) \end{aligned}$$
(26)
where
$$\begin{aligned} \mathscr {E}^\star ({\textbf {u}}, \vartheta , \mu , \alpha )=\int _{\Omega }\omega (\mathbb {S}^\star ), \qquad \mathscr {P}^\star (\dot{{\textbf {u}}}, \mu )=\int _{\partial \Omega }\mu \,{\textbf {j}}\cdot \varvec{\nu } \end{aligned}$$
(27)
are the stored grand-canonical energy and the external power, respectively. Moreover, the following functional space is introduced:
$$\begin{aligned} \mathscr {M}=\{ \mu \in H^1(\Omega ) \,:\, \mu =\mu _0 \;\;\textrm{on}\;\;\partial \Omega \}, \end{aligned}$$
(28)
where \(\mu _{0}\) is the assigned chemical potential on the boundary. In the next section we will provide specific choices of \(\mathscr {L}^\star \) and the consequent governing equations.

3.2 Constitutive choices

In order to properly take into account the interplay between elastic and chemical effects, we choose the grand-canonical potential and the dissipation potentials as follows:
$$\begin{aligned} \widehat{\omega }(\mathbb {S}^\star )=\frac{1}{2}\mathbb {C}\big (\vartheta , \alpha \big ){\textbf {E}}({\textbf {u}})\cdot {\textbf {E}}({\textbf {u}})+\mu {\textbf {B}}\cdot {\textbf {E}}({\textbf {u}})-\frac{1}{2} \kappa \mu ^2, \end{aligned}$$
(29)
with the elasticity tensor given by
$$\begin{aligned} &\\& \mathbb {C}\big (\vartheta , \alpha \big )= \underbrace{E_0 \big ((1-\nu )\hbox {{1}}+\nu {\textbf {1}}\otimes {\textbf {1}}\big )}_{\mathbb {C}_\textrm{is}} &\\& \qquad \qquad \underbrace{-E_0\alpha \,{\textbf {n}}(\vartheta )\otimes {\textbf {n}}(\vartheta ) \otimes {\textbf {n}}(\vartheta )\otimes {\textbf {n}}(\vartheta )}_{\mathbb {C}_\textrm{an} (\vartheta , \alpha )}\end{aligned}$$
being \( \mathbb{1} \) and \({\textbf {1}}\) the fourth and second-order identity tensors, and stress-chemical modulus assigned as
$$\begin{aligned} {\textbf {B}}=\beta {\textbf {1}}. \end{aligned}$$
(30)
The scalar \(\kappa >0\) is the chemistry modulus. The constitutive equations (23) therefore yield the stress and the concentration response:
$$\begin{aligned} \begin{aligned}&{\textbf {S}}&=\mathbb {C}\big (\vartheta , \alpha \big ){\textbf {E}}({\textbf {u}})+\beta \mu {\textbf {1}}\\&=E_0\big ((1-\nu ){\textbf {E}}({\textbf {u}}) +\nu \,\text{ tr }\, {\textbf {E}}({\textbf {u}}){\textbf {1}}\big )\\&\quad-E_0\alpha {\textbf {E}}({\textbf {u}}){\textbf {n}}\cdot {\textbf {n}}+\beta \mu {\textbf {1}},\\&c=\kappa \mu -\beta \text {div}{\textbf {u}}. \end{aligned} \end{aligned}$$
(31)
Therefore, the stress and the concentration response maps are coupled through the stress-chemical modulus \(\beta \). The stress response is isotropic until \(\alpha =0\) and incorporates a spherical part proportional to the chemical potential. As to the dual diffusive dissipation, we set
$$\begin{aligned} \varphi ^\star _\textrm{diff}(\mathbb {S}^\star ; \nabla \mu )=-\frac{1}{2}{\textbf {M}}(\vartheta , \alpha )\nabla \mu \cdot \nabla \mu . \end{aligned}$$
(32)
More specifically, we prescribe the mobility tensor to be linearly dependent on the anisotropic gauge field as follows:
$${\textbf {M}}(\vartheta , \alpha )=\textsf{m}_\textrm{is} {\textbf {1}}+\alpha \textsf{m}_\textrm{an}{\textbf {n}}(\vartheta )\otimes {\textbf {n}}(\vartheta ),$$
where the two non-negative parameters \(\textsf{m}_\textrm{is}\) and \(\textsf{m}_\textrm{an}\) account for the cell mobility; the greater their magnitude, the more facilitated the cells are to spread into the medium. Again, until \(\alpha =0\), the diffusion keeps isotropic.
Finally the local dissipation is chosen as
$$\begin{aligned} \varphi _\textrm{loc}(\mathbb {S}; \dot{\alpha },\nabla \dot{\alpha })= \textsf{w}\big ( \dot{\alpha }+\ell ^2\nabla \alpha \cdot \nabla \dot{\alpha } \big ) +I^{+}(\dot{\alpha }), \end{aligned}$$
(33)
with \(\textsf{w}>0\) and where we have introduced the indicator function \(I^{+}(\dot{\alpha })\) of the set of positive real numbers defined as
$$\begin{aligned} I^{+}(\dot{\alpha }) & = {\left\{ \begin{array}{ll} 0 &{} \text {if } \dot{\alpha } \ge 0, \\ \infty &{} \text {otherwise}, \end{array}\right. } \quad \text {with} \\ \quad \partial I^{+}(\dot{\alpha }) & = {\left\{ \begin{array}{ll} 0 &{} \text {if } \dot{\alpha } > 0, \\ \mathbb {R}^{-} &{} \text {if } \dot{\alpha } = 0, \\ \emptyset &{} \text {otherwise}. \end{array}\right. } \end{aligned}$$
(34)
where we have denoted by \(\partial I^{+}(\dot{\alpha })\) the subdifferential of the indicator function [9, 28]. The first term in (33) is responsible for a threshold effect, which is understandable because the field \(\alpha \) begins to evolve with the onset of micro-buckling, occurring when a critical load is reached. The second is a regularizing term, ruled by an internal length \(\ell \), on which the width of localization fingers depend. The presence of the indicator function in \(\varphi _\textrm{loc}\) is due to the fact that we deal with a material having a reversibility domain with a threshold, that will be determined later on. The assumed expression for \(\varphi _\textrm{loc}\) aims to model a rate-independent material, for what concerns this dissipative contribution, suggesting that the response to anisotropy is influenced solely by the ordered sequence of states experienced during a process, rather than the duration of the deformation process itself.
We therefore end up with the following rate-type saddle problem (see [24]):
$$\begin{aligned} \\& \inf _{\dot{{\textbf {u}}}\in \widetilde{\mathscr {U}}}\inf _{\dot{\vartheta }\in \widetilde{\mathscr {T}}}\inf _{\dot{\alpha }\in \widetilde{\mathscr {A}}} \inf _{\dot{c}\in \widetilde{\mathscr {C}}}\sup _{\mu \in \mathscr {M}} \left\{ \int _{\Omega } \dot{\omega }(\mathbb {S}^\star ) \right. \nonumber \\& \quad \left. -\frac{1}{2} {\textbf {M}}(\alpha ,\vartheta )\nabla \mu \cdot \nabla \mu +\textsf{w} \big (\dot{\alpha }+\ell ^2\nabla \alpha \cdot \nabla \dot{\alpha } \big )\right. \nonumber \\& \quad \left. +I^+(\dot{\alpha }) -\int _{\partial \Omega }{\textbf {t}}\cdot \dot{{\textbf {u}}} \right\} . \end{aligned}$$
(35)
With standard arguments, from (35), we obtain the force balance equation
$$\begin{aligned} \text {div}\Big (E_0\big ((1-\nu ){\textbf {E}}({\textbf {u}}) +\nu \,\text{ tr }\, {\textbf {E}}({\textbf {u}}){\textbf {1}}\big )-E_0\alpha {\textbf {E}}({\textbf {u}}){\textbf {n}}(\vartheta )\cdot {\textbf {n}}(\vartheta )+\beta \mu {\textbf {1}}\Big )={\textbf {0}}. \end{aligned}$$
(36)
On recognizing that \(\partial _\mu \dot{\omega }=-\dot{c}\), the stationarity with respect to \(\mu \), together with (31)\(_2\), yields the following diffusion equation:
$$\begin{aligned} \kappa \dot{\mu } - \beta \text{ tr }\, \dot{{\textbf {E}}}-\textsf{m}_\textrm{is} \Delta \mu - \alpha \textsf{m}_\textrm{an}\text {div}{(\nabla {\mu }\cdot {\textbf {n}}(\vartheta ))}=0. \end{aligned}$$
(37)
The stationarity with respect to \(\dot{\vartheta }\) yields:
$$\begin{aligned} -E_0\alpha \big ({\textbf {E}}({\textbf {u}})\cdot {\textbf {n}}(\vartheta )\otimes {\textbf {n}}(\vartheta )\big )\Big ( {\textbf {E}}({\textbf {u}})\cdot \big ({\textbf {n}}(\vartheta )\otimes {\textbf {n}}'(\vartheta ) \\+{\textbf {n}}'(\vartheta )\otimes {\textbf {n}}(\vartheta ) \big ) \Big )=0; \end{aligned}$$
(38)
a condition that delivers, at each load step increment, the optimal \(\vartheta \) in terms of the strain and the anisotropy. Finally, the stationarity with respect to \(\dot{\alpha }\) yields the evolution equation for the anisotropy gauge field
$$\begin{aligned} -\frac{E_0}{2}({\textbf {E}}({\textbf {u}}){\textbf {n}}(\vartheta )\cdot {\textbf {n}}(\vartheta ))^2+\textsf{w}(1-\ell ^2\Delta \alpha ) +\partial I^+(\dot{\alpha }) \ni 0. \end{aligned}$$
(39)
Due to the nonsmoothness of the dissipation potential function, the evolution equation for the anisotropy gauge field (39) is written as a differential inclusion. From (39), we determine the non-local consistency conditions:
$$\begin{aligned} \textsf{y}\big ({\textbf {E}}({\textbf {u}}),\vartheta ,\alpha \big )\le 0, \quad \dot{\alpha }\ge 0, \quad \dot{\alpha }\,\textsf{y}\big ({\textbf {E}}({\textbf {u}}),\vartheta ,\alpha \big )=0, \end{aligned}$$
(40)
where the anisotropic yield function
$$\begin{aligned} \textsf{y}\big ({\textbf {E}}({\textbf {u}}),\vartheta ,\alpha \big ):=\frac{E_0}{2}({\textbf {E}}({\textbf {u}}){\textbf {n}}(\vartheta ) \cdot {\textbf {n}}(\vartheta ))^2-\textsf{w}(1-\ell ^2\Delta \alpha ) \end{aligned}$$
(41)
has been introduced.4 The condition (40)\(_1\) defines the isotropic domain, and the constant \(\textsf{w}\) specifies the maximum compressive strain in the fiber direction below which the anisotropy cannot evolve.

4 Numerical implementation and results

This section delves into the examination of the scenario outlined in Fig. 1: a hollow domain experiences a progressively increasing internal negative pressure p at given and constant chemical potential \(\mu _0\). The hole stands for a cell cluster, whose effects on the ECM are represented by its capability to contract the tissue and to maintain chemical equilibrium. Our aim is to showcase that our simple model captures, in qualitative terms,5 the localization phenomenon described in the introduction and the subsequent finger-driven diffusion of the cell phase into the ECM.
Fig. 3
a Setup of the numerical experiment. On the left, it is sketched the whole sample, highlighting the domain implemented \(\Omega \). The green square represents the area showed on the right side of the figure, where details on the refinement of the mesh and the boundary conditions are represented. The graphs in b depicts the trend of boundary conditions and time discretization during the simulation as the iteration changes
In Fig. 3a, due to the symmetry of the problem, is reported the effective domain of the analysis and its discretization.
The energetic properties and the fulfillment of overall conditions within the framework of incremental energy minimization strongly suggest the appropriateness of the alternative minimization strategy as the preferred solution algorithm [27]. In this specific case, it involves four sequential steps repeated over each time increment:
1.
The initial minimization of the total energy is performed concerning the displacement field, \({\textbf {u}}\), equivalent to solving a linear elastic problem.
 
2.
The minimization in \(\vartheta \) involves locally maximizing the norm of the compressive strain \(\varepsilon \). This problem is straightforward since \(\vartheta \) is determined by locally solving an eigenproblem to establish the direction of maximum compressive strain.
 
3.
The optimization related to the anisotropic gauge field \(\alpha \) entails solving a Laplacian problem.
 
4.
Finally, the stationarity with respect to the chemical potential \(\mu \) is enforced.
 
The exponential choice of the time evolution is dictated by the typical exponential solution of the diffusion-like equation as (37). The pseudocode presented in Algorithm 1 offers more details in the process of minimization described just above.
Algorithm 1
Alternate minimization
The code, [29], implemented in python as frontend of FEniCSx, a renowned open-source computing platform for solving partial differential equations [1, 2, 21], utilizes an unstructured mesh to discretize the domain, refined in the vicinity of the hole, as shown in 3a. The displacement field \({\textbf {u}}\), along with the anisotropic gauge fields and chemical potential \(\alpha \) and \(\mu \), are considered in piecewise affine scalar finite element spaces over the domain (using FEniCSx P1 elements). Discontinuous Galerkin approximations suffice for the field \(\vartheta \) (DG elements), due to the absence in derivatives of the orientation field.
We present the outcomes of a numerical experiment involving a progressively increasing loading process in pressure p with a constant chemical potential \(\mu \) applied at the boundary \(\partial _p\Omega \), as depicted in Fig. 3b. The experimental setup is illustrated in Fig. 3a and is tailored to explore the mechanical and diffusive characteristics of the material.
In the initial loading phase, the pressure undergoes a linear increase, inducing compressive hoop strain \(\varepsilon \) across the material, with maximum compression occurring at the hole’s boundary. Until the activation criterion (40)\(_1\) is satisfied, the material maintains isotropy (i.e., \(\alpha =0\) throughout) as well the diffusion process. Upon further pressure increments, the criterion (40)\(_1\) is met simultaneously along the entire internal boundary. At this juncture, a thin layer with \(\alpha >0\) undergoes a sudden bifurcation through a snap-back instability, transforming into a series of radial tethers where \(\alpha \rightarrow 1\). Subsequently, cells diffuse in preferred directions dictated by the anisotropy gauge field. It must be noticed that the initiation points of nucleation are determined by the mesh, whereas in the actual problem, they are influenced by small imperfections, which are inherent to such problems. Similar phenomenon is observed in the fracturing process of ceramic slabs subjected to thermal shocks [4] and in periodic array of crack in thermal shock problem [31].
Fig. 4 tracks the evolution of \(\alpha \) and \(\mu \) at point P\(_1\) of the domain. Additionally, snapshots at three specific time points during the simulation for \(\alpha \) and \(\mu \) are depicted in Fig. 5 and Fig. 6, illustrating the preferential diffusion along the formed fingers. In Fig. 7 we present a magnified view of the dashed area in Fig. 6, illustrating the directions of the flux \({\textbf {j}}\) with white arrows.
Fig. 4
\(\hat{\mu }=\mu /\mu _0\)
Fig. 5
Anisotropic gauge field \(\alpha \)
Fig. 6
Chemical potential field \(\hat{\mu }\)
Fig. 7
Chemical potential field \(\hat{\mu }\) and the cell flux \({\textbf {j}}\) (white arrows)

Declarations

Ethical approval

Not applicable.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Fußnoten
1
Alternatively, the external power can be directly written as \(\int _{\partial \Omega }\dot{c}\xi \), where \(\xi \) has the role of a ‘chemical microtraction’, to be constitutively determined (see [25]).
 
2
In [26], section “5.2.6 Inelastic processes in porous media with diffusion”, the authors face this problem from an analytical point of view and propose a formal analysis to obtain such a representation, which involves the determination of the inverse of a linear operator.
 
3
In the sense that it contains also the chemical potential \(\mu \) as a variable.
 
4
To understand the differential inclusion (39) it is useful to interpret the set \(-\textsf{y}\big ({\textbf {E}}({\textbf {u}}),\vartheta ,\alpha \big )+\partial I^+(\dot{\alpha })\) through the subdifferential (34):
1.
If \(\dot{\alpha }>0\), then \(\partial I^+(\dot{\alpha })=\{0\}\), so \(-\textsf{y}\big ({\textbf {E}}({\textbf {u}}),\vartheta ,\alpha \big )+\partial I^+(\dot{\alpha })\) is simply the set \(\{-\textsf{y}\big ({\textbf {E}}({\textbf {u}}),\vartheta ,\alpha \big )\}\). The condition \(-\textsf{y}\big ({\textbf {E}}({\textbf {u}}),\vartheta ,\alpha \big )+\partial I^+(\dot{\alpha })\ni 0\) means that \(\{-\textsf{y}\big ({\textbf {E}}({\textbf {u}}),\vartheta ,\alpha \big )\}\ni 0\), indicating that \(-\textsf{y}\big ({\textbf {E}}({\textbf {u}}),\vartheta ,\alpha \big )\) must be 0.
 
2.
If \(\dot{\alpha }= 0\) then \(\partial I^+(\dot{\alpha })\) is any negative real number. The condition \(-\textsf{y}\big ({\textbf {E}}({\textbf {u}}),\vartheta ,\alpha \big )+\partial I^+(\dot{\alpha })\ni 0\) therefore indicates that \(-\textsf{y}\big ({\textbf {E}}({\textbf {u}}),\vartheta ,\alpha \big )\ge 0\).
 
 
5
A quantitative comparison would necessitate parameter calibration, which is beyond the scope of this article.
 
Literatur
1.
Zurück zum Zitat Aln MS, Kehlet B, Logg A, Richardson C, Ring J, Rognes E, Wells GN (2015) The FEniCS Project Version 1.5. p 15 Aln MS, Kehlet B, Logg A, Richardson C, Ring J, Rognes E, Wells GN (2015) The FEniCS Project Version 1.5. p 15
2.
Zurück zum Zitat Barrata IA, Dean JP, Dokken S, Jørgen, Habera M, Hale J, Richardson C, Rognes ME, Scroggs MW, Sime N, Wells GN (2023) DOLFINx: the next generation FEniCS problem solving environment Barrata IA, Dean JP, Dokken S, Jørgen, Habera M, Hale J, Richardson C, Rognes ME, Scroggs MW, Sime N, Wells GN (2023) DOLFINx: the next generation FEniCS problem solving environment
3.
Zurück zum Zitat Böger L, Keip M-A, Miehe C (2017) Minimization and saddle-point principles for the phase-field modeling of fracture in hydrogels. Comput Mater Sci 138:474–485CrossRef Böger L, Keip M-A, Miehe C (2017) Minimization and saddle-point principles for the phase-field modeling of fracture in hydrogels. Comput Mater Sci 138:474–485CrossRef
4.
Zurück zum Zitat Bourdin B, Marigo JJ, Maurini C, Sicsic P (2014) Morphogenesis and propagation of complex cracks induced by thermal shocks. Phys Rev Lett 112:014301CrossRef Bourdin B, Marigo JJ, Maurini C, Sicsic P (2014) Morphogenesis and propagation of complex cracks induced by thermal shocks. Phys Rev Lett 112:014301CrossRef
5.
Zurück zum Zitat Coleman BD, Gurtin ME (1967) Thermodynamics with internal state variables. J Chem Phys 47(2):597–613CrossRef Coleman BD, Gurtin ME (1967) Thermodynamics with internal state variables. J Chem Phys 47(2):597–613CrossRef
6.
Zurück zum Zitat Conklin MW, Eickhoff JC, Riching KM, Pehlke CA, Eliceiri KW, Provenzano PP, Friedl A, Keely PJ (2011) Aligned collagen is a prognostic signature for survival in human breast carcinoma. Am J Pathol 178(3):1221–1232CrossRef Conklin MW, Eickhoff JC, Riching KM, Pehlke CA, Eliceiri KW, Provenzano PP, Friedl A, Keely PJ (2011) Aligned collagen is a prognostic signature for survival in human breast carcinoma. Am J Pathol 178(3):1221–1232CrossRef
7.
Zurück zum Zitat Favata A, Rodella A, Vidoli S (2022) An internal variable model for plastic remodeling in fibrous materials. Eur J Mech A Solids 96:104718MathSciNetCrossRef Favata A, Rodella A, Vidoli S (2022) An internal variable model for plastic remodeling in fibrous materials. Eur J Mech A Solids 96:104718MathSciNetCrossRef
8.
Zurück zum Zitat Favata A, Rodella A, Vidoli S (2024) Emerging anisotropy and tethering with memory effects in fibrous materials. Mech Mater 190:104928CrossRef Favata A, Rodella A, Vidoli S (2024) Emerging anisotropy and tethering with memory effects in fibrous materials. Mech Mater 190:104928CrossRef
9.
Zurück zum Zitat Frémond M (2002) Non-smooth thermomechanics. Springer-Verlag, Berlin HeidelbergCrossRef Frémond M (2002) Non-smooth thermomechanics. Springer-Verlag, Berlin HeidelbergCrossRef
10.
Zurück zum Zitat Friedl P (2004) Prespecification and plasticity: shifting mechanisms of cell migration. Curr Opin Cell Biol 16:14–23CrossRef Friedl P (2004) Prespecification and plasticity: shifting mechanisms of cell migration. Curr Opin Cell Biol 16:14–23CrossRef
11.
Zurück zum Zitat Friedl P, Bröcker EB (2000) The biology of cell locomotion within three-dimensional extracellular matrix. Cell Motil Cytoskelet 57:41–64 Friedl P, Bröcker EB (2000) The biology of cell locomotion within three-dimensional extracellular matrix. Cell Motil Cytoskelet 57:41–64
12.
Zurück zum Zitat Friedl P, Wolf K (2003) Tumour-cell invasion and migration: diversity and escape mechanisms. Nat Rev Cancer 3:362–374CrossRef Friedl P, Wolf K (2003) Tumour-cell invasion and migration: diversity and escape mechanisms. Nat Rev Cancer 3:362–374CrossRef
13.
Zurück zum Zitat Germain P, Nguyen QS, Suquet P (1983) Continuum thermodynamics. J Appl Mech 50:1010–1020CrossRef Germain P, Nguyen QS, Suquet P (1983) Continuum thermodynamics. J Appl Mech 50:1010–1020CrossRef
14.
Zurück zum Zitat Grekas G, Proestaki M, Rosakis P, Notbohm J, Makridakis C, Ravichandran G (2021) Cells exploit a phase transition to mechanically remodel the fibrous extracellular matrix. J R Soc Interface 18:20200823CrossRef Grekas G, Proestaki M, Rosakis P, Notbohm J, Makridakis C, Ravichandran G (2021) Cells exploit a phase transition to mechanically remodel the fibrous extracellular matrix. J R Soc Interface 18:20200823CrossRef
15.
Zurück zum Zitat Gurtin M, Fried E, Anand L (2010) The mechanics and thermodynamics of continua. Cambridge University Press, CambridgeCrossRef Gurtin M, Fried E, Anand L (2010) The mechanics and thermodynamics of continua. Cambridge University Press, CambridgeCrossRef
16.
Zurück zum Zitat Halphen B, Nguyen QS (1975) Sur les matériaux standards généralisés. J Mécanique 14(1):39–63MathSciNet Halphen B, Nguyen QS (1975) Sur les matériaux standards généralisés. J Mécanique 14(1):39–63MathSciNet
17.
Zurück zum Zitat Hudson NE, Houser JR, O’Brien ET III, Taylor RM II, Superfine R, Lord ST, Falvo MR (2010) Stiffening of individual fibrin fibers equitably distributes strain and strengthens networks. Biophys J 98:1632–1640CrossRef Hudson NE, Houser JR, O’Brien ET III, Taylor RM II, Superfine R, Lord ST, Falvo MR (2010) Stiffening of individual fibrin fibers equitably distributes strain and strengthens networks. Biophys J 98:1632–1640CrossRef
18.
Zurück zum Zitat Kaunas R, Zemel A (eds) (2018) Cell and matrix mechanics. CRC Press, Boca Raton Kaunas R, Zemel A (eds) (2018) Cell and matrix mechanics. CRC Press, Boca Raton
19.
Zurück zum Zitat Kim OV, Liang X, Litvinov RI, Weisel JW, Alber MS, Purohit PK (2016) Foam-like compression behavior of fibrin networks. Biomech Model Mechanobiol 15:213–228CrossRef Kim OV, Liang X, Litvinov RI, Weisel JW, Alber MS, Purohit PK (2016) Foam-like compression behavior of fibrin networks. Biomech Model Mechanobiol 15:213–228CrossRef
20.
Zurück zum Zitat Lakes R, Rosakis P, Ruina A (1993) A microbuckling instability in elastomeric cellular solids. J Mater Sci 28:4667–4672CrossRef Lakes R, Rosakis P, Ruina A (1993) A microbuckling instability in elastomeric cellular solids. J Mater Sci 28:4667–4672CrossRef
21.
Zurück zum Zitat Logg A, Mardal K-A, Wells G (eds) (2012) Automated Solution of Differential Equations by the Finite Element Method, vol 84. Lecture Notes in Computational Science and Engineering. Springer, Berlin Heidelberg Logg A, Mardal K-A, Wells G (eds) (2012) Automated Solution of Differential Equations by the Finite Element Method, vol 84. Lecture Notes in Computational Science and Engineering. Springer, Berlin Heidelberg
22.
Zurück zum Zitat Maugin GA (1990) Internal variables and dissipative structures. J Non-Equilib Thermodyn 15(2):173–192CrossRef Maugin GA (1990) Internal variables and dissipative structures. J Non-Equilib Thermodyn 15(2):173–192CrossRef
23.
Zurück zum Zitat Miehe C (2011) A multi-field incremental variational framework for gradient-extended standard dissipative solids. J Mech Phys Solids 59(4):898–923MathSciNetCrossRef Miehe C (2011) A multi-field incremental variational framework for gradient-extended standard dissipative solids. J Mech Phys Solids 59(4):898–923MathSciNetCrossRef
24.
Zurück zum Zitat Miehe C, Mauthe S, Teichtmeister S (2015) Minimization principles for the coupled problem of Darcy–Biot-type fluid transport in porous media linked to phase field modeling of fracture. J Mech Phys Solids 82:186–217MathSciNetCrossRef Miehe C, Mauthe S, Teichtmeister S (2015) Minimization principles for the coupled problem of Darcy–Biot-type fluid transport in porous media linked to phase field modeling of fracture. J Mech Phys Solids 82:186–217MathSciNetCrossRef
25.
Zurück zum Zitat Miehe C, Mauthe S, Ulmer H (2014) Formulation and numerical exploitation of mixed variational principles for coupled problems of Cahn–Hilliard-type and standard diffusion in elastic solids. Int J Numer Methods Eng 99(10):737–762MathSciNetCrossRef Miehe C, Mauthe S, Ulmer H (2014) Formulation and numerical exploitation of mixed variational principles for coupled problems of Cahn–Hilliard-type and standard diffusion in elastic solids. Int J Numer Methods Eng 99(10):737–762MathSciNetCrossRef
26.
Zurück zum Zitat Mielke A, Roubíček T (2015) Rate-independent systems. Theory and application. Springer, New YorkCrossRef Mielke A, Roubíček T (2015) Rate-independent systems. Theory and application. Springer, New YorkCrossRef
27.
Zurück zum Zitat Petryk H (2003) Incremental energy minimization in dissipative solids. Comptes Rendus Mécanique 331(7):469–474CrossRef Petryk H (2003) Incremental energy minimization in dissipative solids. Comptes Rendus Mécanique 331(7):469–474CrossRef
28.
Zurück zum Zitat TyrrellRockafellar R (1970) Convex analysis. Princeton University Press, PrincetonCrossRef TyrrellRockafellar R (1970) Convex analysis. Princeton University Press, PrincetonCrossRef
30.
Zurück zum Zitat Rodella A, Favata A, Vidoli S (2023) A variational model for plastic reorientation in fibrous material: numerical experiments on phase segregation. In: Materials research proceedings, pp 17–22 Rodella A, Favata A, Vidoli S (2023) A variational model for plastic reorientation in fibrous material: numerical experiments on phase segregation. In: Materials research proceedings, pp 17–22
31.
Zurück zum Zitat Sicsic P, Marigo J-J, Maurini C (2014) Initiation of a periodic array of cracks in the thermal shock problem: A gradient damage modeling. J Mech Phys Solids 63:256–284MathSciNetCrossRef Sicsic P, Marigo J-J, Maurini C (2014) Initiation of a periodic array of cracks in the thermal shock problem: A gradient damage modeling. J Mech Phys Solids 63:256–284MathSciNetCrossRef
32.
Zurück zum Zitat Van Der Rijt JA, Van Der Werf KO, Bennink ML, Dijkstra PJ, Feijen J (2006) Micromechanical testing of individual collagen fibrils. Macromol Biosci 6(8):697–702CrossRef Van Der Rijt JA, Van Der Werf KO, Bennink ML, Dijkstra PJ, Feijen J (2006) Micromechanical testing of individual collagen fibrils. Macromol Biosci 6(8):697–702CrossRef
33.
Zurück zum Zitat Wolf K, Mazo I, Leung H, Engelke K, von Andrian UH, Deryugina EI, Strongin AY, Bröcker EB, Friedl P (2003) Compensation mechanism in tumor cell migration: mesenchymal–ameboid transition after blocking of pericellular proteolysis. J Cell Biol 160:267–277CrossRef Wolf K, Mazo I, Leung H, Engelke K, von Andrian UH, Deryugina EI, Strongin AY, Bröcker EB, Friedl P (2003) Compensation mechanism in tumor cell migration: mesenchymal–ameboid transition after blocking of pericellular proteolysis. J Cell Biol 160:267–277CrossRef
Metadaten
Titel
A variational model for finger-driven cell diffusion in the extracellular matrix
verfasst von
Antonino Favata
Andrea Rodella
Stefano Vidoli
Publikationsdatum
07.08.2024
Verlag
Springer Netherlands
Erschienen in
Meccanica / Ausgabe 8/2024
Print ISSN: 0025-6455
Elektronische ISSN: 1572-9648
DOI
https://doi.org/10.1007/s11012-024-01835-w

    Marktübersichten

    Die im Laufe eines Jahres in der „adhäsion“ veröffentlichten Marktübersichten helfen Anwendern verschiedenster Branchen, sich einen gezielten Überblick über Lieferantenangebote zu verschaffen.