Skip to main content
Log in

An in silico biomechanical analysis of the stent–esophagus interaction

  • Original Paper
  • Published:
Biomechanics and Modeling in Mechanobiology Aims and scope Submit manuscript

An Erratum to this article was published on 20 September 2017

This article has been updated

Abstract

Despite all technological innovations in esophageal stent design over the past 20 years, the association between the stent design’s mechanical behavior and its effect on the clinical outcome has not yet been thoroughly explored. A parametric numerical model of a commercially available esophageal bioresorbable polymeric braided wire stent is set up, accounting for stent design aspects such as braiding angle, strut material, wire thickness, degradation and friction between the wires comprising a predictive tool on the device’s mechanical behavior. Combining this tool with complex multilayered numerical models of the pathological in vivo stressed, actively contracting and buckling esophagus could provide clinicians and engineers with a patient-specific window into the mechanical aspects of stent-based esophageal intervention. This study integrates device and soft tissue mechanics in one computational framework to potentially aid in the understanding of the occurrence of specific symptoms and complications after stent placement.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7

Similar content being viewed by others

Change history

  • 20 September 2017

    In the original publication of the article, Tables 2 and 3 were published with error. The correct tables are provided below (Tables 2, 3). The original version of the article has also been corrected.

References

  • Antoniadis AP et al (2015) Biomechanical modeling to improve coronary artery bifurcation stenting expert review document on techniques and clinical implementation. JACC Cardiovasc Interv 8(10):1281–1296. doi:10.1016/j.jcin.2015.06.015

    Article  Google Scholar 

  • Auricchio F, Constantinescu A, Conti M, Scalet G (2015) A computational approach for the lifetime prediction of cardiovascular balloon-expandable stents. Int J Fatigue 75:69–79

    Article  Google Scholar 

  • Auricchio F, Constantinescu A, Conti M, Scalet G (2016) Fatigue of metallic stents: from clinical evidence to computational analysis. Ann Biomed Eng 44:287–301

    Article  Google Scholar 

  • Auricchio F, Conti M, De Beule M, De Santis G, Verhegghe B (2011) Carotid artery stenting simulation: from patient-specific images to finite element analysis. Med Eng Phys 33:281–289

    Article  Google Scholar 

  • Chiastra C et al (2016) Computational replication of the patient-specific stenting procedure for coronary artery bifurcations: from OCT and CT imaging to structural and hemodynamics analyses. J Biomech 49:2102–2111. doi:10.1016/j.jbiomech.2015.11.024

    Article  Google Scholar 

  • Simulia Corp. (2014) Abaqus Documentation — version 6.14. Dassault Systèmes Simulia Corp. Software Manuals

  • De Beule M (2008) Finite element stent design. Dissertation, Ghent University

  • De Beule M, Van Cauter S, Mortier P, Van Loo D, Van Impe R, Verdonck P, Verhegghe B (2009) Virtual optimization of self-expandable braided wire stents. Med Eng Phys 31:448–453. doi:10.1016/j.medengphy.2008.11.008

    Article  Google Scholar 

  • De Bock S, Iannaccone F, De Santis G, De Beule M, Mortier P, Verhegghe B, Segers P (2012) Our capricious vessels: the influence of stent design and vessel geometry on the mechanics of intracranial aneurysm stent deployment. J Biomech 45:1353–1359. doi:10.1016/j.jbiomech.2012.03.012

    Article  Google Scholar 

  • De Bock S et al (2012) Virtual evaluation of stent graft deployment: a validated modeling and simulation study. J Mech Behav Biomed 13:129–139. doi:10.1016/j.jmbbm.2012.04.021

    Article  Google Scholar 

  • de Jaegere P et al (2016) Patient-specific computer modeling to predict aortic regurgitation after transcatheter aortic valve replacement. JACC Cardiovasc Interv 9:508–512. doi:10.1016/j.jcin.2016.01.003

    Article  Google Scholar 

  • De Putter S, Wolters B, Rutten M, Breeuwer M, Gerritsen F, Van de Vosse F (2007) Patient-specific initial wall stress in abdominal aortic aneurysms with a backward incremental method. J Biomech 40:1081–1090

    Article  Google Scholar 

  • De Santis G et al (2013) Haemodynamic impact of stent-vessel (mal)apposition following carotid artery stenting: mind the gaps!. Comput Methods Biomech Biomed Eng 16:648–659. doi:10.1080/10255842.2011.629997

    Article  Google Scholar 

  • Debusschere N, Segers P, Dubruel P, Verhegghe B, De Beule M (2015) A finite element strategy to investigate the free expansion behaviour of a biodegradable polymeric stent. J Biomech 48:2012–2018

    Article  Google Scholar 

  • Debusschere N, Segers P, Dubruel P, Verhegghe B, De Beule M (2016) A computational framework to model degradation of biocorrodible metal stents using an implicit finite element solver. Ann Biomed Eng 44:382–390

    Article  Google Scholar 

  • Famaey N, Vander Sloten J, Kuhl E (2013) A three-constituent damage model for arterial clamping in computer-assisted surgery. Biomech Model Mechanobiol 12:123–136. doi:10.1007/s10237-012-0386-7

    Article  Google Scholar 

  • Fan Y, Gregersen H, Kassab GS (2004) A two-layered mechanical model of the rat esophagus. Experiment and theory. Biomed Eng Online 3:40

    Article  Google Scholar 

  • Fung Y-C (1970) Mathematical representation of the mechanical properties of the heart muscle. J Biomech 3:381–404

    Article  Google Scholar 

  • Fung Y (1993) Biomechanics: material properties of living tissues. Springer, New York

    Book  Google Scholar 

  • Gasser TC (2016) Biomechanical rupture risk assessment: a consistent and objective decision-making tool for abdominal aortic aneurysm patients. AORTA J 4:42

    Google Scholar 

  • Gasser TC, Ogden RW, Holzapfel GA (2006) Hyperelastic modelling of arterial layers with distributed collagen fibre orientations. J R Soc Interface 3:15–35

    Article  Google Scholar 

  • Gregersen H (2003) Biomechanics of the gastrointestinal tract: new perspectives in motility research and diagnostics. Springer, New York

    Book  Google Scholar 

  • Gregersen H, Kassab G (1996) Biomechanics of the gastrointestinal tract. Neurogastroenterol Motil 8:277–297

    Article  Google Scholar 

  • Gregersen H, Kassab GS, Fung YC (2000) The zero-stress state of the gastrointestinal tract: biomechanical and functional implications. Dig Dis Sci 45:2271–2281

    Article  Google Scholar 

  • Gregersen H, Pedersen J, Drewes AM (2008) Deterioration of muscle function in the human esophagus with age. Dig Dis Sci 53:3065–3070

    Article  Google Scholar 

  • Grogan JA, Leen SB, McHugh PE (2013) Optimizing the design of a bioabsorbable metal stent using computer simulation methods. Biomaterials 34:8049–8060. doi:10.1016/j.biomaterials.2013.07.010

    Article  Google Scholar 

  • Hall GJ, Kasper EP (2006) Comparison of element technologies for modeling stent expansion. J Biomech Eng 128:751–756. doi:10.1115/1.2264382

    Article  Google Scholar 

  • Hill AV (1970) First and last experiments in muscle mechanics. Cambridge University Press, Cambridge

    Google Scholar 

  • Hirdes MM, Vleggaar FP, de Beule M, Siersema PD (2013) In vitro evaluation of the radial and axial force of self-expanding esophageal stents. Endoscopy 45:997–1005. doi:10.1055/s-0033-1344985

    Article  Google Scholar 

  • Hirdes MM, Vleggaar FP, Siersema PD (2011) Stent placement for esophageal strictures: an update. Expert Rev Med Devices 8:733–755. doi:10.1586/erd.11.44

    Article  Google Scholar 

  • Huxley A (1957) Muscle structure and theories of contraction. Prog Biophys Biophys Chem 7:255

    Google Scholar 

  • Iannaccone F et al (2014) The influence of vascular anatomy on carotid artery stenting: a parametric study for damage assessment. J Biomech 47:890–898. doi:10.1016/j.jbiomech.2014.01.008

    Article  Google Scholar 

  • Isayama H et al (2009) Measurement of radial and axial forces of biliary self-expandable metallic stents. Gastrointest Endosc 70:37–44. doi:10.1016/j.gie.2008.09.032

    Article  Google Scholar 

  • Jedwab MR, Clerc CO (1993) A study of the geometrical and mechanical properties of a self-expanding metallic stent-theory and experiment. J Appl Biomater 4:77–85. doi:10.1002/jab.770040111

    Article  Google Scholar 

  • Kajzer W, Kaczmarek M, Marciniak J (2005) Biomechanical analysis of stent-oesophagus system. J Mater Process Technol 162:196–202

    Article  Google Scholar 

  • Kim JH, Kang TJ, Yu W-R (2008) Mechanical modeling of self-expandable stent fabricated using braiding technology. J Biomech 41:3202–3212

    Article  Google Scholar 

  • Kou W, Bhalla APS, Griffith BE, Pandolfino JE, Kahrilas PJ, Patankar NA (2015) A fully resolved active musculo-mechanical model for esophageal transport. arXiv preprint arXiv:1501.02010

  • Kuo B, Urma D (2006) Esophagus-anatomy and development. GI Motil online

  • Li G et al (2013) Biodegradable weft-knitted intestinal stents: fabrication and physical changes investigation in vitro degradation. J Biomed Mater Res A. doi:10.1002/jbm.a.34759

    Google Scholar 

  • Li M, Brasseur JG (1993) Non-steady peristaltic transport in finite-length tubes. J Fluid Mech 248:129–151

    Article  MATH  Google Scholar 

  • Liao D, Fan Y, Zeng Y, Gregersen H (2003) Stress distribution in the layered wall of the rat oesophagus. Med Eng Phys 25:731–738

    Article  Google Scholar 

  • Liao D, Lelic D, Gao F, Drewes AM, Gregersen H (2008) Biomechanical functional and sensory modelling of the gastrointestinal tract. Philos Trans A Math Phys Eng Sci 366:3281–3299. doi:10.1098/rsta.2008.0091

    Article  Google Scholar 

  • Liao D, Zhao J, Fan Y, Gregersen H (2004) Two-layered quasi-3D finite element model of the oesophagus. Med Eng Phys 26:535–543. doi:10.1016/j.medengphy.2004.04.009

    Article  Google Scholar 

  • Liao D, Zhao J, Yang J, Gregersen H (2007) The oesophageal zero-stress state and mucosal folding from a GIOME perspective. World J Gastroenterol 13:1347–1351

    Article  Google Scholar 

  • Liao DH, Zhao JB, Gregersen H (2009) Gastrointestinal tract modelling in health and disease. World J Gastroenterol 15:169–176

    Article  Google Scholar 

  • Miller LS et al (1995) Correlation of high-frequency esophageal ultrasonography and manometry in the study of esophageal motility. Gastroenterology 109:832–837

    Article  Google Scholar 

  • Mittal RK, Padda B, Bhalla V, Bhargava V, Liu J (2006) Synchrony between circular and longitudinal muscle contractions during peristalsis in normal subjects. Am J Physiol Gastrointest Liver Physiol 290:G431–G438

    Article  Google Scholar 

  • Molony DS, Kavanagh EG, Madhavan P, Walsh MT, McGloughlin TM (2010) A computational study of the magnitude and direction of migration forces in patient-specific abdominal aortic aneurysm stent-grafts. Eur J Vasc Endovasc Surg 40:332–339. doi:10.1016/j.ejvs.2010.06.001

    Article  Google Scholar 

  • Murtada S-I, Kroon M, Holzapfel GA (2010) A calcium-driven mechanochemical model for prediction of force generation in smooth muscle. Biomech Model Mechanobiol 9:749–762

    Article  Google Scholar 

  • Murtada S-I, Lewin S, Arner A, Humphrey J (2016) Adaptation of active tone in the mouse descending thoracic aorta under acute changes in loading. Biomech Model Mechanobiol 15:579–592

    Article  Google Scholar 

  • Natali AN, Carniel EL, Gregersen H (2009) Biomechanical behaviour of oesophageal tissues: material and structural configuration, experimental data and constitutive analysis. Med Eng Phys 31:1056–1062

    Article  Google Scholar 

  • Nicosia MA, Brasseur JG, Liu J-B, Miller LS (2001) Local longitudinal muscle shortening of the human esophagus from high-frequency ultrasonography. Am J Physiol Gastrointest Liver Physiol 281:G1022–G1033

    Article  Google Scholar 

  • Nuutinen JP, Clerc C, Tormala P (2003a) Theoretical and experimental evaluation of the radial force of self-expanding braided bioabsorbable stents. J Biomater Sci Polym Ed 14:677–687

    Article  Google Scholar 

  • Nuutinen JP, Clerc C, Tormala P (2003b) Theoretical and experimental evaluation of the radial force of self-expanding braided bioabsorbable stents. J Biomat Sci Polym Ed 14:677–687. doi:10.1163/156856203322274932

    Article  Google Scholar 

  • Peirlinck M (2013) Design of biodegradable esophageal stents. Master thesis, Ghent University

  • Pouderoux P, Lin S, Kahrilas PJ (1997) Timing, propagation, coordination, and effect of esophageal shortening during peristalsis. Gastroenterology 112:1147–1154

    Article  Google Scholar 

  • Puckett J, Bhalla V, Liu J, Kassab G, Mittal R (2005) Oesophageal wall stress and muscle hypertrophy in high amplitude oesophageal contractions. Neurogastroenterol Motil 17:791–799

    Article  Google Scholar 

  • Schultz C et al (2016) Patient-specific image-based computer simulation for theprediction of valve morphology and calcium displacement after TAVI with the medtronic CoreValve and the Edwards SAPIEN valve. EuroIntervention 11:1044–1052

    Article  Google Scholar 

  • Shanahan C, Tofail SA, Tiernan P (2017) Viscoelastic braided stent: finite element modelling and validation of crimping behaviour. Mater Des 121:143–153

    Article  Google Scholar 

  • Sharma P, Kozarek R, Practice Parameters Committee of American College of G (2010) Role of esophageal stents in benign and malignant diseases. Am J Gastroenterol 105:258–273. doi:10.1038/ajg.2009.684 (quiz 274)

  • Simulia Corp (2014) Abaqus Documentation—version 6.14. Dassault Systèmes Simulia Corp. Software Manuals

  • Sokolis DP (2010) Strain-energy function and three-dimensional stress distribution in esophageal biomechanics. J Biomech 43:2753–2764. doi:10.1016/j.jbiomech.2010.06.007

    Article  Google Scholar 

  • Sokolis DP (2013) Structurally-motivated characterization of the passive pseudo-elastic response of esophagus and its layers. Comput Biol Med 43:1273–1285. doi:10.1016/j.compbiomed.2013.06.009

    Article  Google Scholar 

  • Sommer G, Schriefl A, Zeindlinger G, Katzensteiner A, Ainodhofer H, Saxena A, Holzapfel GA (2013) Multiaxial mechanical response and constitutive modeling of esophageal tissues: impact on esophageal tissue engineering. Acta Biomater 9:9379–9391. doi:10.1016/j.actbio.2013.07.041

    Article  Google Scholar 

  • Stavropoulou EA, Dafalias YF, Sokolis DP (2009) Biomechanical and histological characteristics of passive esophagus: experimental investigation and comparative constitutive modeling. J Biomech 42:2654–2663. doi:10.1016/j.jbiomech.2009.08.018

    Article  Google Scholar 

  • Uitdehaag MJ, Siersema PD, Spaander MC, Vleggaar FP, Verschuur EM, Steyerberg EW, Kuipers EJ (2010) A new fully covered stent with antimigration properties for the palliation of malignant dysphagia: a prospective cohort study. Gastrointest Endosc 71:600–605. doi:10.1016/j.gie.2009.09.023

    Article  Google Scholar 

  • Villadsen G, Storkholm J, Zachariae H, Hendel L, Bendtsen F, Gregersen H (2001) Oesophageal pressure–cross sectional area distributions and secondary peristalsis in relation to subclassification of systemic sclerosis. Neurogastroenterol Motil 13:199–210

    Article  Google Scholar 

  • Walter D et al (2015) A randomized trial comparing biodegradable stent placement and endoscopic dilation for recurrent benign esophageal strictures (Destiny study). United Eur Gastroenterol J 3:A24. doi:10.1177/2050640615610034

    Google Scholar 

  • Yang J, Zhao J, Liao D, Gregersen H (2006a) Biomechanical properties of the layered oesophagus and its remodelling in experimental type-1 diabetes. J Biomech 39:894–904

    Article  Google Scholar 

  • Yang W, Fung T, Chian K, Chong C (2006b) 3D Mechanical properties of the layered esophagus: experiment and constitutive model. J Biomech Eng 128:899–908

    Article  Google Scholar 

  • Yang W, Fung TC, Chian KS, Chong CK (2006c) Directional, regional, and layer variations of mechanical properties of esophageal tissue and its interpretation using a structure-based constitutive model. J Biomech Eng 128:409–418. doi:10.1115/1.2187033

    Article  Google Scholar 

  • Yang W, Fung TC, Chian KS, Chong CK (2006d) Viscoelasticity of esophageal tissue and application of a QLV model. J Biomech Eng 128:909–916. doi:10.1115/1.2372473

    Article  Google Scholar 

  • Yang W, Fung TC, Chian KS, Chong CK (2007a) Finite element simulation of food transport through the esophageal body. World J Gastroenterol 13:1352–1359

    Article  Google Scholar 

  • Yang W, Fung TC, Chian KS, Chong CK (2007b) Instability of the two-layered thick-walled esophageal model under the external pressure and circular outer boundary condition. J Biomech 40:481–490. doi:10.1016/j.jbiomech.2006.02.020

  • Yang W, Fung TC, Chian KS, Chong CK (2007c) Three-dimensional finite element model of the two-layered oesophagus, including the effects of residual strains and buckling of mucosa. Proc Inst Mech Eng H 221:417–426

    Article  Google Scholar 

  • Zhao S, Liu XC, Gu L (2012) The impact of wire stent fabrication technique on the performance of stent placement. J Med Devices 6:011007

    Article  Google Scholar 

Download references

Acknowledgements

The authors gratefully acknowledge Peter Dubruel, Ph.D., David De Wilde, Ph.D., and Wouter Kappelle, MD, for their valuable support and assistance. We also thank MPT Europe for giving us access to the radial stent compression unit, which greatly assisted the mechanical characterization of the studied stent samples. This research was supported by the Flanders Innovation & Entrepreneurship Agency, strategic basic research Grant No.141014.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Mathias Peirlinck.

Ethics declarations

Conflicts of interest

The authors declare that they have no conflict of interest.

Additional information

The original version of this article was revised: The tables 2 and 3 have been corrected.

An erratum to this article is available at https://doi.org/10.1007/s10237-017-0963-x.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary material 1 (mp4 7189 KB)

Appendices

Appendix 1: Esophageal histology

The esophageal wall is composed of four well-defined layers: the mucosa, the submucosa, the muscularis externa and the adventitia (Fig. 8) (Gregersen 2003; Gregersen et al. 2000). The mucosal layer is build up from connective tissue and can be subdivided in the epithelium, the lamina propria and the muscularis mucosa. Contrary to the mucosa, where very fine collagen fibrils are loosely and randomly organized, collagen fibrils in the submucosa are organized in thick aligned fibers running helically down the esophagus (Natali et al. 2009). Due to this difference in structure, the submucosa plays a stronger mechanical role in offering resistance to the deformation of the esophageal wall than the mucosa does. The submucosa is loosely interconnected (Sommer et al. 2013) with the muscularis externa which is mainly composed of smooth and striated muscle cells (the transition from striated to smooth muscle is spread throughout the middle third of the esophagus). This muscle layer can be subdivided in two layers, the inner one having muscle cells arranged in the circumferential and the outer one arranged in the axial direction. The muscle layer is surrounded by a layer of soft connective tissue, the adventitia, which supports the flexibility of the esophagus and its integration in the thorax (Gregersen 2003). In this study, the esophagus was modeled as a five-layered system consisting of the mucosa (M), the collagen-rich submucosa (SM), a virtual interfacial layer (IF) and the inner and the outer muscle layer of the muscularis externa (ME). The interfacial layer was introduced to incorporate the loose interconnection between the submucosa and the muscularis externa into our models.

Fig. 8
figure 8

Image credit: biophoto associates/science photo library

Light micrograph of a cross section through the human esophagus depicting the mucosa, submucosa, the loose connective tissue in between the submucosa and the muscle layers (modeled as the interfacial layer in the finite element models), the circular and the longitudinal muscle layers.

Appendix 2: Constitutive modeling

The esophagus exhibits, as most other biologic soft tissues, a large deformability, together with nonlinear, pseudo-elastic and anisotropic material properties due to its aforementioned layer-specific composition (Fan et al. 2004; Yang et al. 2006a). Constitutive models are capable of characterizing this mechanical behavior through a functional relation between stresses and strains (Famaey et al. 2013). Following Fung et al., anisotropic hyperelastic material properties can be adequately represented constitutively by means of a strain energy function (SEF) (Fung 1993). The general material model applied in this study describes active esophageal tissue. Active contraction of the smooth (and striated) muscle cells is the most important feature of the peristaltic movement of the esophagus. In this model, the muscle cells were modeled to be oriented with their longest axes along the contractile direction and evenly distributed along the depth and height of the muscle tissue (Murtada et al. 2010). The contractile unit consists of two thin filaments (actin) and one thick filament (myosin) in a side-polar structure with cross-bridges attaching these filaments. These units are modeled to be arranged in series, separated by dense bodies forming long contractile fibers inside the muscle tissue (see Fig. 2, middle). In addition, they are assumed to contract uniformly and acting as a single unit. The three element Hill model has been developed to handle this muscle behavior in more detail (Hill 1970). The model considers the tissue as a composite of a ’contractile element’ connected with a ‘series elastic element’ to describe the active contraction of the muscle, and a ’parallel element’ describing the connective tissue (Gregersen et al. 2000). These contractile elements behave similarly in smooth muscle and striated muscle. The proposed constitutive model was additively decomposed in three constituents: an isotropic matrix material constituent, an anisotropic constituent attributed to the dispersion of collagen fibers in the tissue and an anisotropic smooth cell constituent. Due to the esophagus’s composition and our choice for a five-layered model, the three constituents will never contribute simultaneously in a specific layer. Several research groups have modeled the matrix and fiber constituent through phenomenological (Sokolis 2010; Stavropoulou et al. 2009) or microstructurally motivated (Sommer et al. 2013; Yang et al. 2006c) constitutive descriptors. In this study, we chose to model the passive deformational response of the intact esophagus based upon the pioneering work of Gasser et al. in arterial tissue mechanics (Gasser et al. 2006). The third, active contraction, constituent was motivated by Murtada et al.’s model of mechanical smooth-muscle activation (Murtada et al. 2010). Famaey et al. (2013) similarly combined Gasser’s passive and Murtada’s active material model in a cardiovascular setting, studying the stiffness degradation and identifying the critical loading regimes during arterial clamping.

1.1 A.2.1 Kinematics

To set forth our material model, we first must introduce some key kinematic quantities. For soft tissues undergoing large deformation, the deformation is typically quantified through the deformation gradient \({\varvec{F}}\), that is the gradient of the deformation map \(\varvec{\chi } \) with respect to the undeformed position \({\varvec{X}}\).

$$\begin{aligned} {\varvec{F}}=\frac{\partial {\varvec{x}}}{\partial {\varvec{X}}}=\frac{\partial \varvec{\chi } \left( {\varvec{X}} \right) }{\partial {\varvec{X}}} \end{aligned}$$

The local volume ratio of the material, which is close to 1 for nearly incompressible materials, is defined by the Jacobian J

$$\begin{aligned} J=\hbox {det}\left( {\varvec{F}} \right) \end{aligned}$$

The deformation gradient \({\varvec{F}}\) can be decomposed multiplicatively into a spherical (dilatational) part \(J^{1/3}{\varvec{I}}\) and a unimodular (distortional) part \(\bar{{\varvec{F}}}\), so that \(\det \left( {\bar{{\varvec{F}}}} \right) =1\).

The deformation of the material will be defined in terms of the invariants of the right and left Cauchy–Green tensors, denoted C and b, respectively, and their isochoric counterparts, denoted \(\bar{{\varvec{C}}} \) and \(\bar{{\varvec{b}}}\), respectively, associated with \( \bar{{\varvec{F}}}\).

With \(\lambda {_r} \), \(\lambda {_\theta } \) and \(\lambda {_z} \) being the principal distortional stretches in the radial, circumferential and axial direction, the invariants describing the deformation of the isotropic matrix material can be expressed as follows:

$$\begin{aligned} \bar{I}_{1}= & {} \lambda {_r} ^{2}+\lambda _\theta ^{2}+\lambda _z ^{2} \\ \bar{I}_{2}= & {} \frac{1}{2}\left[ \bar{I}_{1}^{2}-\left( {\lambda _r^{4}+\lambda _\theta ^{4}+\lambda _z ^{4}} \right) \right] \\ \bar{I}_{3}= & {} \lambda _r ^{2}\lambda _\theta ^{2}\lambda _z ^{2} \end{aligned}$$

while the stretches along the fiber or muscle cells direction are described as follows:

$$\begin{aligned} \bar{I}_{4} ^\mathrm{fib}= & {} \lambda _\theta ^{2}\mathrm{cos}^{2}\alpha ^\mathrm{fib1}+\lambda _z ^{2}\mathrm{sin}^{2}\alpha ^\mathrm{fib1} \\ \bar{I}_6 ^\mathrm{{fib}}= & {} \lambda _\theta ^{2}\mathrm{cos}^{2}\alpha ^\mathrm{fib2}+\lambda _z ^{2}\mathrm{sin}^{2}\alpha ^\mathrm{fib2} \\ \bar{I} ^\mathrm{smc}= & {} \lambda _\theta ^{2}\mathrm{cos}^{2}\alpha ^\mathrm{smc}+\lambda _z ^{2}\mathrm{sin}^{2}\alpha ^\mathrm{smc} \end{aligned}$$

with \(\alpha ^\mathrm{fib1}, \alpha ^\mathrm{fib2}, \alpha ^\mathrm{smc}\) the angle between the first fiber, the second fiber, the smooth/striated muscle cell, respectively, and the circumferential direction.

To account for the dispersion of collagen fibers along their averaged direction, Gasser et al. (2006) proposed the use of the following pseudo-invariants

$$\begin{aligned} \begin{array}{l} \bar{I}_{4} ^\mathrm{fib{\prime }}=\kappa \bar{I}_{1} +\left( {1-3\kappa } \right) \bar{I}_{4} ^\mathrm{{fib}} \\ \bar{I}_{6} ^\mathrm{fib{\prime }}=\kappa \bar{I}_{1} +\left( {1-3\kappa } \right) \bar{I}_{6} ^\mathrm{{fib}} \\ \end{array} \end{aligned}$$

in which \(\kappa \) is a structure parameter representing the fiber distribution in an integral sense and describes its ‘degree of anisotropy’ (\(\kappa =0\) corresponds to an anisotropic non-dispersed fiber state, while \(\kappa =1/3\) corresponds to the totally dispersed fiber, and thus isotropic, state). This parameter should be determined from histological data.

1.2 A.2.2 Strain energy function

As the esophageal tissue can be considered nearly incompressible due to the high liquid content of the ground substance (Gregersen et al. 2000; Natali et al. 2009), we additively decomposed the SEF \(\Psi \) in a volumetric \(\Psi ^{vol}\) and a deviatoric \(\Psi ^\mathrm{dev}\) part.

$$\begin{aligned} \Psi =\Psi ^\mathrm{vol}+\Psi ^\mathrm{dev} \end{aligned}$$

The deviatoric part was additively decomposed in the isotropic contribution of the matrix material \(\Psi ^\mathrm{mat}\), the anisotropic contribution of two families of collagen fibers \(\Psi ^\mathrm{fib1}\) and \(\Psi ^\mathrm{fib2}\), and the contribution of the smooth/striated muscle cells \(\Psi ^\mathrm{smc}\).

$$\begin{aligned} \Psi ^\mathrm{dev}=\Psi ^\mathrm{mat}+\Psi ^\mathrm{fib1}+\Psi ^\mathrm{fib2}+\Psi ^\mathrm{smc} \end{aligned}$$

1.2.1 A.2.2.1 Volumetric bulk material

The volumetric free energy \(\Psi ^\mathrm{vol}\) was characterized by the hyperelastic constitutive Arruda–Boyce model, which can be expressed as:

$$\begin{aligned} \Psi ^\mathrm{{vol}}=\frac{K_0 }{2}\left( {\frac{J^{2}-1}{2}-\hbox {ln}\left( J \right) } \right) \end{aligned}$$

with \(K_{0}\) being the bulk modulus. To ensure near-incompressibility, the bulk modulus had to be set high enough to ensure the ratio of the bulk modulus to the initial shear modulus \(\frac{K_{0}}{\mu _{0}}\) is large enough. In this study, this ratio was chosen to be equal to 20.

1.2.2 A.2.2.1 Non-collagenous ground matrix

The extracellular matrix is characterized by means of an incompressible neo-Hookean model, i.e.,

$$\begin{aligned} \Psi ^\mathrm{{mat}}=\frac{1}{2}c\left( {\bar{I}_{1}-3 }\right) \end{aligned}$$

where \(\bar{I}_{1}\) denotes the previously defined first invariant of the Cauchy–Green tensor, and c the neo-Hookean parameter, which characterizes the matrix’s stiffness through its initial shear modulus (\(c=\mu _{0})\). The neo-Hookean material model is a special case of the polynomial hyperelastic material model (also called generalized Rivlin model).

1.2.3 A.2.2.3 Collagen fibers

The collagen fibers have no strength under compression and can thus only contribute to the SEF when under tension. Following Gasser et al. (2006), we particularized their transversely isotropic free energy function for the \(i\mathrm{th}\) family of collagen fibers as follows:

$$\begin{aligned} \Psi ^\mathrm{fib_i }=\frac{k_1 }{2k_2 }\left[ {\mathrm{exp}\left( {k_2 \left[ {\bar{I}_i ^\mathrm{fib^{\prime }}-1} \right] ^{2}-1} \right) } \right] \end{aligned}$$

in which \(k_2 >0\) is a dimensionless parameter and \(k_1 >0\) is a stress-like parameter to be determined from mechanical tests of the tissue.

1.2.4 A.2.2.4 Smooth and striated muscle cells

The muscle cells form an integral part of the matrix constituent, both in their passive and active state. Murtada et al. developed a mechanochemical model describing muscle contraction and relaxation (Murtada et al. 2010). This model is based on the sliding filament theory first described by Huxley (1957). The theory states how the muscle contraction is the result of the relative sliding between thin actin and thick myosin filaments, which is accomplished by configurational changes in the cross-bridges between these filaments. These configurational cross-bridges changes are modeled through a four-state model in which cycling between these states is governed by chemically catalyzed (de)phosphorylation of the myosin heads. The rate parameters of transition between these states may differ significantly depending on the method of activation and the muscle type. For esophageal muscle tissue, no isometric and/or isotonic contraction experiments are available (ex infra- Sect. A.2.5), so we simplified the model by assuming all the myosin heads to be in their force generating state for this model. This lead to the following strain energy component:

$$\begin{aligned} {\Psi ^\mathrm{smc}}=\frac{1}{2}\mu _\mathrm{smc} \left[ {\sqrt{\bar{I}^\mathrm{{smc}}}+u_\mathrm{{rs}} -1 }\right] ^{2} \end{aligned}$$

where \(\mu _\mathrm{{smc}}\) characterized the stiffness of the actin-myosin filament apparatus and \(u_\mathrm{{rs}} \) stands for the averaged normalized relative sliding between both filaments. Due to the strong hyperelasticity of the esophageal muscle tissue (ex infra), a quadratic degree did not suffice to phenomenologically describe the tissue. A quartic degree proved to be more appropriate (ex infra):

$$\begin{aligned} {\Psi ^\mathrm{{smc}}}=\frac{1}{2}\mu _\mathrm{{smc}} \left[ {\sqrt{\bar{I} ^\mathrm{{smc}}}+u_\mathrm{{rs}} -1} \right] ^{4} \end{aligned}$$

1.2.5 A.2.2.5 Layer-specific equations and material properties

The full three-constituent material model summarized to:

$$\begin{aligned} \Psi =\Psi ^\mathrm{{vol}}+\Psi ^\mathrm{{dev}}=\Psi ^\mathrm{{vol}}+\Psi ^\mathrm{{mat}}+\Psi ^\mathrm{fib1}+\Psi ^\mathrm{fib2}+\Psi ^\mathrm{{smc}} \end{aligned}$$

For the mucosa (M) with its loosely and randomly thin collagen fibrils, the full material model was simplified to isotropic hyperelastic behavior. The same behavior was used for the interfacial layer (IF) between the submucosa and the muscle tissue.

$$\begin{aligned} \Psi ^{\textit{M}}=\Psi ^\mathrm{{IF}}=\Psi ^\mathrm{{vol}}+\Psi ^\mathrm{{mat}} \end{aligned}$$

The submucosa with its thick aligned fibers running helically down the esophagus was modeled as follows:

$$\begin{aligned} \Psi ^\mathrm{{SM}}=\Psi ^\mathrm{{vol}}+\Psi ^\mathrm{{mat}}+\Psi ^\mathrm{fib1}+\Psi ^\mathrm{fib2} \end{aligned}$$

Both muscle layers consist fully out of smooth/striated muscle cells and no cumulative collagen fibers are present, which lead to:

$$\begin{aligned} \Psi ^\mathrm{{ME}}=\Psi ^\mathrm{{vol}}+\Psi ^\mathrm{{mat}}+\Psi ^\mathrm{{smc}} \end{aligned}$$
Table 3 Fitted constitutive parameters
Fig. 9
figure 9

Comparison of averaged experimental results on ovine tissue samples (extracted from Sommer et al. (2013)) and the constitutive material model fit

1.3 A.2.3 Evaluation of the constitutive parameters

The material parameters introduced by the outlined constitutive model need to be calibrated based on specific experimental setups. As we did not dispose of any esophageal tissue, we had to rely on literature to pinpoint the quantitative values of these parameters. Several studies have focused on characterizing the passive material properties of each layer of the esophageal wall (Fan et al. 2004; Liao et al. 2003; Natali et al. 2009; Sokolis 2010, 2013; Sommer et al. 2013; Stavropoulou et al. 2009; Yang et al. 2006b, c). Unfortunately, these articles are not conclusive on layer-specific mechanical behavior. While Liao et al. (2003), Fan et al. (2004) and Yang et al. (2006a) concluded that the muscle tissue is stiffer than the (sub)mucosal tissue, the opposite was concluded in Stavropoulou et al. (2009), whereas Yang et al. (2006b) and Sommer et al. (2013) stated that the mucosa–submucosa is initially more compliant and becomes stiffer than the muscle layer at increasing pressures. These discrepancies can be caused by tissue samples originating from different species (where different ratios in layer thicknesses might be seen), differences in chemical tissue preparation to ensure passive relaxed muscle behavior and the experimental protocol using only uniaxial or multiaxial data. Regarding the latter, only a fraction of these studies (Liao et al. 2003; Sokolis 2010, 2013; Stavropoulou et al. 2009) focused on the characterization of the multiaxial passive behavior of esophageal tissue, mostly based on extension–inflation tests. Sommer et al. (2013) investigated the heterogeneity of the multilayered composite esophageal structure by performing (layer-specific) uniaxial tensile, biaxial tensile and extension–inflation tests on ovine esophagi. We chose to calibrate the constitutive model based on this study for two reasons. First, Sommer et al.’s experimental results constitute the most complete multiaxial data set available up to date. Second, the mucosa–submucosa was found to be initially more compliant and become stiffer at increasing pressures, similar to the results described in the only available study on porcine data (Yang et al. 2006b), which is often considered as the best mechanical surrogate for human tissue. The constitutive parameters for each layer of our model were calibrated in a nonlinear least squares fitting procedure in which the constitutive biaxial response of the model was compared to the ex vivo results. As Sommer et al. studied the combined mucosa–submucosa layer and the full double-layered muscle layer, some assumptions had to be made while fitting the constitutive parameters to their averaged experimental results. Focusing on the mucosa and submucosa, the esophageal histology (“Appendix 1”) allowed us to assume a mechanically similar non-collagenous ground matrix for both these layers. All stiffening of the tissue based on the recruitment of the anisotropic distributed embedded wavy collagen fibrils was assumed to be attributed to the submucosa’s collagen content only. For the virtual interfacial layer introduced to incorporate the loose connective tissue connection between the submucosa and the muscularis externa, we assumed isotropic hyperelastic behavior one order of magnitude softer than the mucosal layer. Both the circumferential and longitudinal muscle layers were assumed to mechanically behave identically except for their smooth muscle cell orientations. To fit the material parameters to the experimentally measured passive strength of the full (two-layered) muscularis externa slab, we assumed the muscularis externa to behave as a single layer with their smooth muscle cells oriented per Sommer et al.’s fitted ‘collagen fiber’ orientations, leading to an angle of \(55.3^{\circ }\) with respect to the circumferential direction. The fitted constitutive material parameters can be found in Table 3, and the fit is depicted in Fig. 9.

As can be seen in Fig. 9, a second-degree muscle cell invariant dependency of the strain energy function proved inadequate to describe the strong hyperelastic behavior of the muscle cells. The aforementioned quartic muscle cell invariant dependency strain energy potential model resulted in a better fit.

1.4 A.2.4 FEA implementation

The outlined constitutive model was implemented in the Abaqus user subroutines UMAT and VUMAT, a family of implicit (Abaqus-Standard) and explicit (Abaqus-Explicit) subroutines, respectively, designed to define a material’s mechanical behavior, in which the stress state in the local material orientations gets updated at the end of the increment for which it is called. In the implicit subroutine, the material Jacobian matrix \(\frac{\partial \sigma }{\partial \epsilon }\) also needs to be fed back to the FEA solver at the end of each increment (Simulia Corp 2014).

1.5 A.2.5 Limitations of the constitutive model

Although the used dataset (Sommer et al. 2013) for the constitutive material fit comprises, to the authors’ knowledge, the most complete multiaxial study on esophageal tissue, the fitted mechanical parameters can only fully reflect the correct passive behavior of the material following an appropriate design of the experiments. The active response of the esophageal muscle layers in situ should be quantified in ex vivo or in vivo isometric and/or isotonic contraction experiments. Due to a lack of these types of experiments on esophageal smooth/striated muscle cell tissue in literature, the four-state model describing the kinetics of the actin-myosin power stroke (Murtada et al. 2010) and the viscous evolution law of the average normalized relative sliding between both filaments \(\dot{u}_\mathrm{rs.}\) could not be quantified. For that reason, future work on the experimental characterization of active esophageal muscle tissue (through ex vivo or in vivo isometric and/or isotonic contraction experiments) should be applauded. Murtada et al. (2016) recently proposed a newer mathematical model in which the intrinsically coupled active and passive wall contributions by smooth muscle cells to the total Cauchy stress in the wall was described as the sum of the passive and active stresses, in which the active stress contribution is no longer part of the strain energy function, which could be very interesting if fitted to esophageal muscle tissue. As the smooth muscle cell filaments act like active dashpots which contract proportionally to the resulting force acting upon them (three element Hill model (Fung 1970)), the active tension of smooth muscle cells typically increases until it reaches a local maximum point. The ability of the muscle to develop active tension depends on the initial stretch state of the muscle, e.g., (Gregersen et al. 2008; Murtada et al. 2016). This mechanical phenomenon is not included into the model due to the simplifications (no viscous evolution law for \(\dot{u}_\mathrm{rs.})\) to the smooth muscle cell strain energy potential contributions (which had to be taken due to unavailable experimental data). Consequently, the applied constitutive model might overestimate the contractile force of the esophageal muscle tissue at higher stretches, which is schematically illustrated in Fig. 10.

Fig. 10
figure 10

Schematic illustration of a stress–strain diagram for esophageal muscle tissue. The outlined and calibrated constitutive model will overestimate the esophageal muscle’s total force compared to the physiological total force

Additionally, as this study sets forth to develop patient-specific esophagus models, we have to note that Gregersen et al. (2008) have recently shown that the human esophageal properties and function deteriorate with age (the wall becomes stiffer and the maximal active esophageal muscle tension deteriorated after the age of 40 years). The proposed constitutive material parameter fit was not tuned and/or scaled to the considered patient’s esophageal function or age, but based on the passive multiaxial study of ovine tissue by Sommer et al. (2013).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Peirlinck, M., Debusschere, N., Iannaccone, F. et al. An in silico biomechanical analysis of the stent–esophagus interaction. Biomech Model Mechanobiol 17, 111–131 (2018). https://doi.org/10.1007/s10237-017-0948-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10237-017-0948-9

Keywords

Navigation