Introduction

Cardiac biomechanical modeling has made tremendous progress over the past decades, and some accurate models are now available to represent the complex deformations of the organ—among other quantities of interest—over full heartbeats, frequently based on multi-physics and multi-scale formulations, see e.g. [1, 2] and references therein.

As for all natural systems—as e.g. in geophysics—a great challenge consists in dealing with the many unknown or uncertain quantities—initial conditions, boundary conditions, and various physical parameters—that need to be prescribed for running model simulations [3]. In this work, we decide to rely on a data assimilation strategy [4] to estimate the uncertain quantities while allowing predictive simulations.

Concerning the specific problem of estimation in cardiac biomechanical modeling, difficulties arise from both (1) the complexity of the models considered, and (2) the nature of the available measurements, often relying on medical imaging [5]. An effective estimation methodology has been proposed by [6] for this type of model, based on a so-called sequential approach—also known as observer method. In this approach, the dynamical model is corrected at each time using the computed discrepancy between the current simulation and the actual measurements, see also [7]. This strategy was designed to be applicable to measurements concerning displacements, whether they be given internally—in a sub-region of the system—or on a boundary or a part thereof. It was also shown to be extendable to data consisting of segmented surfaces as obtained by processing various types of medical imaging dynamical sequences.

In this paper, we focus on estimation based on data provided by tagged-MR imaging sequences [8, 9]. Tagged-MR is generally considered to be the “gold standard” in cardiac imaging, in particular as regards the assessment of so-called “cardiac mechanical indicators”, namely, indicators pertaining to displacements, strains, and volumes [10]. As a matter of fact, tagged-MR images visualize the deformations of grids associated with the actual tissues, which is of course most valuable for clinical purposes, both from a qualitative standpoint as assessed by the physician’s eye, and with a view to obtaining such quantitative indicators. However, the problem of extracting actual 3D material displacements from a tagged-MR sequence gives rise to serious difficulties [11, 12]. In fact, in many cases only 2D “apparent” displacements are obtained, which introduces specific errors in the displacement-based quantitative indicators, in addition to usual inaccuracies pertaining to image processing. Of course, these difficulties are also of concern when extracted displacements are to be used in an estimation setup, hence this justifies looking more closely into tagged-MR modalities to devise and analyze strategies to adequately employ them for estimation purposes. In this regard, the contributions presented in this communication are twofold.

First, we propose a systematic approach to incorporate within an estimation framework a wide range of data, potentially obtained from prior processing steps applied on tagged-MR images. These data vary in their nature, covering the cases of: full mechanical displacements in a subdomain; sequences of deforming tag planes and tag grids; and 2D apparent displacements. Extracting state—and parameter—corrections from these data is an intricate task. To address this challenge, we devise for every case relevant discrepancy measures. The soundness of our approach is corroborated by a complete mathematical analysis of the state observer in an idealized fully linear case, provided as complementary material in Appendix A.

Secondly, we propose a relevant time-discretization scheme for the state observer, which is a particularly crucial point in the context of sequential data estimation. This scheme is built upon a “prediction–correction” strategy, where the former step corresponds to genuine model time marching, and the latter to discrepancy-based adjustments. This clean decomposition of these steps offers numerous practical benefits. Additionally, we are able to provide evidence that the obtained time-discrete observer retains the convergence properties of the time-continuous observer, with rigorous proofs detailed in Appendix B.

The outline of the paper is as follows. In the forthcoming section we recall the main principles underlying the design of observers, and we provide a quick overview of the mechanical model of a beating heart, as an example of a model formulation. The next section is dedicated to describing the potential information extracted from tagged-MR images and to proposing—for each type of data—the discrepancy measures. We then address the issue of space and time discretization of the observer in order to perform joint state and parameter estimation. Finally, in the last section we present several numerical experiments in which we performed parameter estimation based on synthetic measurements.

Position of the problem

Principles of sequential estimator design

The aim of a sequential estimator—also called observer—is to approximate a real trajectory, in spite of various uncertainties, using the knowledge provided by the measurements obtained on this specific real trajectory. Let us consider a real trajectory \({{\mathrm {y}}}^{{\text {ref}}}(t),\ t \in [0,+\infty )\), belonging to the so-called state space \({\mathcal {Y}}\) and solution, in our case, of a—possibly infinite-dimensional—dynamical system summarized in the state space form

$$\begin{aligned} {\dot{{{\mathrm {y}}}}}^{{\text {ref}}} = {{\mathrm {F}}}({{\mathrm {y}}}^{{\text {ref}}},t), \end{aligned}$$

with an uncertain initial condition

$$\begin{aligned} {{\mathrm {y}}}^{{\text {ref}}}(0) = {{\mathrm {y}}}_0 + \zeta _{{\mathrm {y}}}, \end{aligned}$$

where \({{\mathrm {y}}}_0\) is a known a priori and \(\zeta _{{\mathrm {y}}}\) is the uncertain part in the initial condition. Therefore, any simulation of \({{\mathrm {y}}}\)—based on the discretization of the dynamical system—starting only from \({{\mathrm {y}}}_0\) will be affected by the propagation of this error made in the initial condition. To circumvent this difficulty, we can benefit from the measurements at our disposal on the trajectory. We denote by \({{\mathrm {z}}}\) these measurements—also called observations and belonging to the observation space \({\mathcal {Z}}\)—which are assumed to be generated by a map** \({{\mathrm {H}}}\) on the real trajectory, up to additional measurement errors

$$\begin{aligned} {{\mathrm {z}}}= {{\mathrm {H}}}({{\mathrm {y}}}^{{\text {ref}}},t) + \chi . \end{aligned}$$

The observer denoted by \({\widehat{{{\mathrm {y}}}}}\) is a system that starts from the only part known in the initial condition—namely \({{\mathrm {y}}}_0\)—and uses in time the available measurements \({{\mathrm {z}}}\) to generate a trajectory \({\widehat{{{\mathrm {y}}}}}(t),\ t \in [0,+\infty )\) that converges to \({{\mathrm {y}}}^{{\text {ref}}}\) as fast as possible. Therefore, simulating \({\widehat{{{\mathrm {y}}}}}\) instead of \({{\mathrm {y}}}\) from \({{\mathrm {y}}}_0\) gives a better approximation of the targeted system.

The main categories of observers addressed here are computed by a feedback law based on the measurements in the form

where \({{\mathrm {G}}}\) is called the gain operator, also referred to as filter. The dynamics of \({\widehat{{{\mathrm {y}}}}}\) is corrected when a discrepancy is observed between the actual measurements \({{\mathrm {z}}}\) and the measurement \({{\mathrm {H}}}({\widehat{{{\mathrm {y}}}}})\) that would have been produced by \({\widehat{{{\mathrm {y}}}}}\). This discrepancy

$$\begin{aligned} {{\mathrm {J}}}({\widehat{{{\mathrm {y}}}}},t) = {{\mathrm {z}}}- {{\mathrm {H}}}({\widehat{{{\mathrm {y}}}}}) \end{aligned}$$

is also called innovation since it not only expresses an observation error, but also a source of improvement for the observer. We point out that with certain types of measurements—as is typically the case with image-based observations—it is sometimes difficult to define an adequate observation operator but easier to directly compute a discrepancy [6]. This is not a problem for the observer definition since only the discrepancy appears.

In a fully linear situation, namely, when the dynamics is linear with \({{\mathrm {F}}}({{\mathrm {y}}},t) = {{\mathrm {A}}}(t) {{\mathrm {y}}}+ {{\mathrm {R}}}\) and \({{\mathrm {H}}}({{\mathrm {y}}},t) = {{\mathrm {H}}}(t) {{\mathrm {y}}}\), the most well-known gain operator is given by the Kalman gain, see e.g. [13, 14] and references therein. This operator is expressed as \({{\mathrm {G}}}(t) = {{\mathrm {P}}}(t){{\mathrm {H}}}(t)^*\) where \({{\mathrm {P}}}\) is an operator following the Riccati evolution equation

$$\begin{aligned} \dot{{{\mathrm {P}}}} = {{\mathrm {A}}}{{\mathrm {P}}} + {{\mathrm {P}}}{{{\mathrm {A}}}}^{*} - {{\mathrm {P}}}{{\mathrm {H}}}^*{{\mathrm {H}}}{{\mathrm {P}}}, \quad {{\mathrm {P}}}(0) = {{\mathrm {P}}}_0, \end{aligned}$$

and \({{\mathrm {H}}}^*\) is the adjoint of \({{\mathrm {H}}}\). Although \({{\mathrm {P}}}\) is computable for any dynamics operator \({{\mathrm {A}}}\), it leads after spatial discretization to a discrete operator which is intractable in practice. This phenomenon has been known for decades and called “curse of dimensionality” [14, 15]. Therefore, for specific dynamics other types of gains have been investigated as initiated by [16]. They are based on the fact that, when computing the estimation error \({\widetilde{{{\mathrm {y}}}}} = {{\mathrm {y}}}^{{\text {ref}}} - {\widehat{{{\mathrm {y}}}}}\), we get in a fully linear setting the following dynamics

$$\begin{aligned} {\left\{ \begin{array}{ll} \dot{{\widetilde{{{\mathrm {y}}}}}} = ({{\mathrm {A}}}- {{\mathrm {G}}} {{\mathrm {H}}}) {\widetilde{{{\mathrm {y}}}}} - {{\mathrm {G}}}\chi \\ {\widetilde{{{\mathrm {y}}}}}(0) = \zeta _{{\mathrm {y}}}. \end{array}\right. } \end{aligned}$$

Hence, \({{\mathrm {G}}}\) should be designed to stabilize the estimation error dynamics operator \({{\mathrm {A}}}- {{\mathrm {G}}} {{\mathrm {H}}}\), so that the homogeneous system tends to 0, namely,

$$\begin{aligned} {\widetilde{{{\mathrm {y}}}}}_{\chi =0} \xrightarrow {t \rightarrow +\infty } 0. \end{aligned}$$

In the presence of noise in the measurements, this would also control the error dynamics. This strategy is referred to as the Luenberger observer or nudging—see [17] for a survey. For the elastodynamics system—a particular case of second-order hyperbolic systems— [6] has shown that a very simple choice of

$$\begin{aligned} {{\mathrm {G}}} = \gamma {{\mathrm {H}}}^*, \end{aligned}$$
(1)

with \(\gamma \) a scalar coefficient can be sufficient.

In a nonlinear configuration, fewer theoretical results are available. However, an accepted strategy is to replace in the gain the use of the adjoint of the observation operator, namely \({{\mathrm {H}}}^*\), by the adjoint of the tangent operator \(D{{\mathrm {H}}}({\widehat{{{\mathrm {y}}}}})^*\) around the estimated trajectory. Therefore for small errors we can expect that the linearized error around the trajectory is stable.

One last fundamental aspect that we need to describe in this introduction to observer design is how parameter estimation—also called identification—can be carried out. Let us denote by \(\theta \) the uncertain parameter to be identified. Note that \(\theta \) may be a vector of components or even a distributed field. The main idea is to introduce an augmented state vector and dynamics operator

$$\begin{aligned} {}^{\flat }{{\mathrm {y}}}= \begin{pmatrix} {{\mathrm {y}}}\\ \theta \end{pmatrix}, \quad {}^{\flat }{{\mathrm {F}}}({}^{\flat }{{\mathrm {y}}},t) = \begin{pmatrix} {{\mathrm {F}}}({{\mathrm {y}}},\theta ,t) \\ 0 \end{pmatrix}, \quad \end{aligned}$$

such that we still have \({}^{\flat }{\dot{{{\mathrm {y}}}}} = {}^{\flat }{{\mathrm {F}}}({}^{\flat }{{\mathrm {y}}},t)\). Then, a Kalman observer can be directly defined on this augmented model leading to a covariance operator and gain

$$\begin{aligned} {{\mathrm {P}}}= \begin{pmatrix} {{\mathrm {P}}}_{{{\mathrm {y}}}{{\mathrm {y}}}} &{} {{\mathrm {P}}}_{{{\mathrm {y}}}\theta } \\ {{\mathrm {P}}}_{\theta {{\mathrm {y}}}} &{} {{\mathrm {P}}}_{\theta \theta } \end{pmatrix}, \quad {{\mathrm {G}}}= \begin{pmatrix} {{\mathrm {G}}}_{{\mathrm {y}}}\\ {{\mathrm {G}}}_\theta \end{pmatrix}. \end{aligned}$$

However, it is more intricate to define a relevant Luenberger observer for the augmented system as the observations are frequently linked to the parameters through the state only. Therefore, there is little hope that \(\gamma {{\mathrm {H}}}^*\) will lead to an efficient gain. An alternative strategy was proposed by [18] as a generalization of the adaptive filtering strategy of [19, 20]. The idea is to retain the Luenberger observer on the state while using a Kalman-like gain on the parameters. This strategy can be very effective in practice, since it is common to consider a parameter described much more coarsely than the state discretization, thus alleviating the curse of dimensionality associated with optimal filtering. The complete observer reads

(2)

where \({{\mathrm {U}}}^{-1}\) is a reduced covariance operator on the parameter space and \({{\mathrm {L}}}\) is a “sensitivity” operator from the parameter space to the state space. We see in the dynamics (2) that the state gain is the combination of the Luenberger gain and a gain directly inferred from the parameter filter so that

$$\begin{aligned} {{\mathrm {G}}}_{{\mathrm {y}}}= (\gamma \mathbb {1} + {{\mathrm {L}}} {{\mathrm {U}}}^{-1} {{\mathrm {L}}}^*) D{{\mathrm {H}}}^*. \end{aligned}$$

In [18] the convergence of the complete observer is also established—at least in a linear configuration—based on the idea that the Luenberger state observer reduces the uncertainty to the parameter space where the optimal filter operates. Moreover, the effectiveness of this approach has already been applied to biomechanical identification problems by [7,

$$\begin{aligned} \varvec{\varphi }: \left| \begin{aligned} \Omega _0\times [0,T]&\longrightarrow \Omega (t)\\ (\varvec{\xi },t)&\longmapsto \varvec{x}= \varvec{\varphi }(\varvec{\xi },t) = \varvec{\xi }+ \varvec{u}(\varvec{\xi },t), \end{aligned} \right. \end{aligned}$$

where \(\varvec{u}\) denotes the solid displacement, so that the solid velocity is given by \(\varvec{v}= {\dot{\varvec{u}}}\). The deformation gradient tensor \(\varvec{F}\) is given by

$$\begin{aligned} \varvec{F}(\varvec{\xi },t) = \varvec{\nabla }_{\varvec{\xi }} \varvec{\varphi }= \varvec{\mathbb {1}} + \varvec{\nabla }_{\varvec{\xi }}\varvec{u}. \end{aligned}$$

Furthermore, we introduce the right Cauchy-Green deformation tensor \(\varvec{C}= \varvec{F}^{\intercal }\cdot \varvec{F}\). We finally recall that the Green-Lagrange strain tensor denoted by \(\varvec{e}\) is defined by

$$\begin{aligned} \varvec{e}= \frac{1}{2} (\varvec{C}-\varvec{\mathbb {1}}) = \frac{1}{2} \Bigl (\varvec{\nabla }_{\varvec{\xi }}\varvec{u}+ (\varvec{\nabla }_{\varvec{\xi }}\varvec{u})^\intercal + (\varvec{\nabla }_{\varvec{\xi }}\varvec{u})^\intercal \cdot \varvec{\nabla }_{\varvec{\xi }}\varvec{u}\Bigr ). \end{aligned}$$

Regarding the mechanical quantities notation, we denote by \(\rho \) the tissue mass per unit volume, and by \(\varvec{\sigma }\) the Cauchy stress tensor associated with the deformed configuration. In the reference configuration, we respectively define the associated first and second Piola-Kirchhoff stress tensors as \(\varvec{T}= J \varvec{\sigma }\cdot \varvec{F}^{-\intercal }\) and \(\varvec{\Sigma }= \varvec{F}^{-1} \cdot \varvec{T}= J \varvec{F}^{-1} \cdot \varvec{\sigma }\cdot \varvec{F}^{-\intercal }\), where \(J = \det {\varvec{F}}\).

The constitutive law can be considered as a nonlinear rheological combination of a passive part and an active part \(\varvec{T}= \varvec{\Sigma }_p + \varvec{\Sigma }_a\). The passive part \(\varvec{\Sigma }_p\) is described by a hyperelastic law of potential \({\mathcal {W}}\) and a viscous component chosen proportional to the strain rate \({\dot{\varvec{e}}}\)

$$\begin{aligned} \varvec{\Sigma }_p(\varvec{e},{\dot{\varvec{e}}}) = \frac{\partial {\mathcal {W}}}{\partial \varvec{e}}(\varvec{e}) + \eta _s {\dot{\varvec{e}}}. \end{aligned}$$

Concerning the hyperelastic law, there exists some experimental evidence—based on detailed ex-vivo tri-axial shear testing—in favor of a complete orthotropic passive behavior [22], with a so-called sheet structure providing a second privileged direction, namely, after the muscle fiber direction. However, the sheet direction cannot be characterized in-vivo for patient-specific modeling purposes. Moreover, various studies have shown good agreements of transversely isotropic models with experimental data obtained at the organ level, see e.g. [23, 24]. We thus consider a transversely isotropic law of exponential type earlier proposed by [25], and inspired from the fully orthotropic model of [26], viz.

$$\begin{aligned} {\mathcal {W}}= C_0\exp \big (C_1(J_1 - 3)^2\big ) + C_2\exp \big (C_3(J_4 - 1)^2\big ) + \kappa (J - 1) - \kappa \ln (J), \end{aligned}$$

where \(J_1\) is the standard first reduced invariant, \(J_4\) is the reduced invariant accounting for the anisotropy of the material in the fiber direction \(\varvec{\tau }\), namely \( J_1 = J^{-\frac{2}{3}}{{\text {tr}}}(\varvec{C}),~J_4 = J^{-\frac{2}{3}}\varvec{\tau }\cdot \varvec{C}\cdot \varvec{\tau }, \) and \(\kappa \) is the bulk coefficient.

For the active part \(\varvec{\Sigma }_a\), we rely on the model proposed in [27], with internal variables defining the active strain \(e_c\), the active stiffness \(k_c\) and the associated active stress \(\tau _c\), along the fiber direction \(\varvec{\tau }\) in a chemically-controlled constitutive law describing myofibre mechanics [27, 28]. Therefore, we have \(\varvec{\Sigma }_a = \sigma _{{\scriptscriptstyle 1{{\text {D}}}}}(e_c,k_c,\tau _c) \varvec{\tau }\otimes \varvec{\tau }\). We finally end up with the following second Piola-Kirchhoff stress tensor

$$\begin{aligned} \varvec{\Sigma }(\varvec{e},e_c,k_c,\tau _c) = \frac{\partial {\mathcal {W}}}{\partial \varvec{e}}(\varvec{e}) + \eta _s {\dot{\varvec{e}}} + \sigma _{{\scriptscriptstyle 1{{\text {D}}}}}(e_c,k_c,\tau _c)\varvec{\tau }\otimes \varvec{\tau }. \end{aligned}$$

Concerning the boundary conditions, following [7] we model the interactions with the surrounding organs by visco-elastic boundary conditions on a subpart of the boundary, which gives in the reference configuration \( \varvec{T}\cdot \varvec{n}= k_s \varvec{u}+ c_s \varvec{v}{{\text { on }}} \Gamma _n\), where \(\varvec{n}\) denotes the surface normal in the reference configuration. Regarding the pressure load, we consider a uniform following pressure on the left and right endocardium easily written in the deformed configuration \( \varvec{\sigma }\cdot \varvec{n}_t = - p_{v,i} \, \varvec{n}_t {{\text { on }}} \Gamma _{n,i}(t), i=\{1,2\}, \) where, here, \(\varvec{n}_t\) denotes the normal of the deformed configuration boundary. Finally, the complete mechanical model reads

$$\begin{aligned} \left\{ \begin{array}{ll} {\dot{\varvec{u}}} = \varvec{v}, &{} \quad \text{ in } \Omega _0\\ \rho {\dot{\varvec{v}}} - \mathop {\mathbf {div}}(\varvec{T}) = 0,&{} \quad \text{ in } \Omega _0\\ \varvec{T}\cdot \varvec{n}= k_s \varvec{u}+ c_s \varvec{v}, &{} \quad {{\text {on }}} \Gamma _n \\ \varvec{T}\cdot \varvec{n}= - J p_{v,i} \, \varvec{F}^{-\intercal } \cdot \varvec{n}, &{} \quad {{\text {on }}} \Gamma _{c,i} \\ \varvec{T}\cdot \varvec{n}= 0, &{} \quad {{\text {on }}} \partial \Omega _0\backslash ((\cup _i \Gamma _{c,i}) \cup \Gamma _n). \end{array} \right. \end{aligned}$$
(3)

Together with the internal variable dynamics given in [27], it constitutes a general definition of the dynamical operator denoted by \({{\mathrm {F}}}\) in our above summarized description for a state \({{\mathrm {y}}}\) corresponding to \((\varvec{u},\varvec{v},e_c,k_c,\tau _c)\).