Skip to main content
Log in

Interlayer micromechanics of the aortic heart valve leaflet

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

Abstract

While the mechanical behaviors of the fibrosa and ventricularis layers of the aortic valve (AV) leaflet are understood, little information exists on their mechanical interactions mediated by the GAG-rich central spongiosa layer. Parametric simulations of the interlayer interactions of the AV leaflets in flexure utilized a tri-layered finite element (FE) model of circumferentially oriented tissue sections to investigate inter-layer sliding hypothesized to occur. Simulation results indicated that the leaflet tissue functions as a tightly bonded structure when the spongiosa effective modulus was at least 25 % that of the fibrosa and ventricularis layers. Novel studies that directly measured transmural strain in flexure of AV leaflet tissue specimens validated these findings. Interestingly, a smooth transmural strain distribution indicated that the layers of the leaflet indeed act as a bonded unit, consistent with our previous observations (Stella and Sacks in J Biomech Eng 129:757–766, 2007) of a large number of transverse collagen fibers interconnecting the fibrosa and ventricularis layers. Additionally, when the tri-layered FE model was refined to match the transmural deformations, a layer-specific bimodular material model (resulting in four total moduli) accurately matched the transmural strain and moment-curvature relations simultaneously. Collectively, these results provide evidence, contrary to previous assumptions, that the valve layers function as a bonded structure in the low-strain flexure deformation mode. Most likely, this results directly from the transverse collagen fibers that bind the layers together to disable physical sliding and maintain layer residual stresses. Further, the spongiosa may function as a general dampening layer while the AV leaflets deforms as a homogenous structure despite its heterogeneous architecture.

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
Fig. 8

Similar content being viewed by others

Abbreviations

AV:

Aortic valve

AC:

Flexure direction directed against the natural curvature of the leaflet

ECM:

Extracellular matrix

FE:

Finite element

GAG:

Glycosaminoglycans

I :

Second moment of inertia

\(\Delta \kappa \) :

Change in valve leaflet curvature during flexure testing

M :

Applied bending moment

PG:

Proteoglycan

\(\mu \) :

Shear modulus

TE:

Tissue engineering

W :

Strain energy function

WC:

Flexure direction directed with the natural curvature of the leaflet

References

  • Ateshian GA (2007) Anisotropy of fibrous tissues in relation to the distribution of tensed and buckled fibers. J Biomech Eng 129:240–249

    Article  Google Scholar 

  • Billiar KL, Sacks MS (2000) Biaxial mechanical properties of the natural and glutaraldehyde treated aortic valve cusp-part I: experimental results. J Biomech Eng 122:23–30

    Article  Google Scholar 

  • Butler DL, Goldstein SA, Guilak F (2000) Functional tissue engineering: the role of biomechanics. J Biomech Eng 122:570–575

    Google Scholar 

  • Carruthers CA, Alfieri CM, Joyce EM, Watkins SC, Yutzey KE, Sacks MS (2012) Gene expression and collagen fiber micromechanical interactions of the semilunar heart valve interstitial cell. Cell Mol Bioeng 135:1–12

    Google Scholar 

  • Christie G, Stephenson R (1989) Modelling the mechanical role of the fibrosa and ventricularis in the porcine bioprosthesis. In: International symposium on surgery for heart valve disease, ICR Publishers, London, pp 815–824

  • Curnier A, He QC, Zysset P (1995) Conewise linear elastic-materials. J Elast 37:1–38

    Article  MATH  MathSciNet  Google Scholar 

  • Eckert CE, Fan R, Mikulis B, Barron M, Carruthers CA, Friebe VM et al (2013) On the biomechanical role of glycosaminoglycans in the aortic heart valve leaflet. Acta Biomater 9:4653–4660

    Article  Google Scholar 

  • Engelmayr GC Jr, Hildebrand DK, Sutherland FW, Mayer JE Jr, Sacks MS (2003) A novel bioreactor for the dynamic flexural stimulation of tissue engineered heart valve biomaterials. Biomaterials 24:2523–2532

    Article  Google Scholar 

  • Engelmayr GC Jr, Rabkin E, Sutherland FW, Schoen FJ, Mayer JE Jr, Sacks MS (2005) The independent role of cyclic flexure in the early in vitro development of an engineered heart valve tissue. Biomaterials 26:175–187

    Article  Google Scholar 

  • Gloeckner D, Billiar K, Sacks M (1998) The bending behavior of fixed porcine aortic cusp. In: Third world congress of biomechanics, Hokkaido, Japan.

  • Gloeckner DC, Billiar KL, Sacks MS (1999) Effects of mechanical fatigue on the bending properties of the porcine bioprosthetic heart valve. Asaio J 45:59–63

    Article  Google Scholar 

  • Go AS, Mozaffarian D, Roger VL, Benjamin EJ, Berry JD, Borden WB et al (2013) Heart disease and stroke statistics-2013 update: a report from the American Heart Association. Circulation 127:e6–e245

    Article  Google Scholar 

  • Grashow JS, Sacks MS, Liao J, Yoganathan AP (2006a) Planar biaxial creep and stress relaxation of the mitral valve anterior leaflet. Ann Biomed Eng 34:1509–1519

    Article  Google Scholar 

  • Grashow JS, Yoganathan AP, Sacks MS (2006b) Biaxial stress-stretch behavior of the mitral valve anterior leaflet at physiologic strain rates. Ann Biomed Eng 34:315–325

    Article  Google Scholar 

  • Iyengar AKS, Sugimoto H, Smith DB, Sacks MS (2001) Dynamic in vitro quantification of bioprosthetic heart valve leaflet motion using structured light projection. Ann Biomed Eng 29:963–973

    Article  Google Scholar 

  • Lam T (2004) The mechanical properties of native porcine aortic and pulmonary heart valve leaflets (thesis). Type, University of Pittsburgh, Pittsburgh, PA

  • Merryman WD, Huang HY, Schoen FJ, Sacks MS (2006) The effects of cellular contraction on aortic valve leaflet flexural stiffness. J Biomech 39:88–96

    Article  Google Scholar 

  • Mirnajafi A, Raymer JM, McClure LR, Sacks MS (2006) The flexural rigidity of the aortic valve leaflet in the commissural region. J Biomech 39:2966–2973

    Article  Google Scholar 

  • Missirlis Y, Chong M (1978) Aortic valve mechanics—part I: material properties of natural porcine aortic valves. J Bioeng 2:287–300

    Google Scholar 

  • Mohri H, Reichenback D, Merendino K (1972) Biology of homologous and heterologous aortic valves. In: Ionescu M, Ross D, Wooler G (eds) Biological tissue in heart valve replacement. Butterworths, London, p 137

    Google Scholar 

  • Rivlin RS (1949) Large elastic deformations of isotropic materials. 5. The problem of flexure. Proc R Soc Lond Ser 195:463–473

    Article  MATH  MathSciNet  Google Scholar 

  • Sacks MS, David Merryman W (2009) On the biomechanics of heart valve function. J Biomech 42:1804–1824

    Article  Google Scholar 

  • Sacks MS, Smith DB, Hiester ED (1998) The aortic valve microstructure: effects of transvalvular pressure. J Biomed Mater Res 41:131–141

    Article  Google Scholar 

  • Sacks MS, Yoganathan AP (2007) Heart valve function: a biomechanical perspective. Philos Trans R Soc Lond B Biol Sci 362:1369–1391

    Article  Google Scholar 

  • Sauren AA, Kuijpers W, van Steenhoven AA, Veldpaus FE (1980) Aortic valve histology and its relation with mechanics-preliminary report. J Biomech 13:97–104

    Article  Google Scholar 

  • Schoen F, Levy R (1999) Tissue heart valves: current challenges and future research perspectives. J Biomed Mater Res 47:439–465

    Article  Google Scholar 

  • Scott M, Vesely I (1995) Aortic valve cusp microstructure: the role of elastin. Ann Thorac Surg 60:S391–S394

    Article  Google Scholar 

  • Song T, Vesely I, Boughner D (1990) Effects of dynamic fixation on shear behaviour of porcine xenograft valves. Biomaterials 11:191–196

    Article  Google Scholar 

  • Stella JA, Liao J, Sacks MS (2007) Time-dependent biaxial mechanical behavior of the aortic heart valve leaflet. J Biomech 40:3169– 3177

    Article  Google Scholar 

  • Stella JA, Sacks MS (2007) On the biaxial mechanical properties of the layers of the aortic valve leaflet. J Biomech Eng 129:757–766

    Article  Google Scholar 

  • Stephens EH, Chu CK, Grande-Allen KJ (2008) Valve proteoglycan content and glycosaminoglycan fine structure are unique to microstructure, mechanical loads, and age: relevance to an age-specific tissue engineered heart valve. Acta Biomater (in press)

  • Sugimoto B, Sacks MS (2013) Effects of leaflet stiffness on in-vitro dynamic bioprosthetic heart valve leaflet shape. Cardiovasc Eng Technol (in-press)

  • Talman EA, Boughner DR (1995) Glutaraldehyde fixation alters the internal shear properties of porcine aortic heart valve tissue. Ann Thorac Surg 60:S369–S373

    Article  Google Scholar 

  • Thubrikar M (1990) The aortic valve. CRC, Boca Raton

    Google Scholar 

  • Thubrikar M, Bosher LP, Nolan SP (1979) The mechanism of opening of the aortic valve. J Thorac Cardiovasc Surg 77:863–870

    Google Scholar 

  • Thubrikar M, Harry R, Nolan SP (1977) Normal aortic valve function in dogs. Am J Cardiol 40:563–568

    Article  Google Scholar 

  • Thubrikar M, Piepgrass WC, Bosher LP, Nolan SP (1980) The elastic modulus of canine aortic valve leaflets in vivo and in vitro. Circ Res 47:792–800

    Article  Google Scholar 

  • Thubrikar MJ, Aouad J, Nolan SP (1986) Comparison of the in vivo and in vitro mechanical properties of aortic valve leaflets. J Thorac Cardiovasc Surg 92:29–36

    Google Scholar 

  • Timoshenko S (1953) History of strength of materials, with a brief account of the history of theory of elasticity and theory of structures. McGraw-Hill, New York

    Google Scholar 

  • Tseng H, Grande-Allen KJ (2012) Elastic fibers in the aortic valve spongiosa: a fresh perspective on its structure and role in overall tissue function (vol 7, pg 2101, 2011). Acta Biomater 8, 924

  • Vesely I (1996) Reconstruction of loads in the fibrosa and ventricularis of porcine aortic valves. Asaio J 42:M739–M746

    Article  Google Scholar 

  • Vesely I (1998) The role of elastin in aortic valve mechanics. J Biomech 31:115–123

    Article  Google Scholar 

  • Vesely I, Boughner D (1989) Analysis of the bending behaviour of porcine xenograft leaflets and of neutral aortic valve material: bending stiffness, neutral axis and shear measurements. J Biomech 22:655–671

    Article  Google Scholar 

  • Vesely I, Noseworthy R (1992) Micromechanics of the fibrosa and the ventricularis in aortic valve leaflets. J Biomech 25:101–113

    Article  Google Scholar 

  • Wiltz D, Arevalos CA, Balaoing LR, Blancas AA, Sapp MC, Zhang X, Grande-Allen KJ (2013) Extracellular matrix organization, structure, and function. In Aikawa E (ed) Calcific Aortic Valve Disease, InTech. Available from http://www.intechopen.com/books/calcific-aortic-valve-disease/extracellular-matrix-organization-structure-and-function

  • Woo SLY, Seguchi Y (1989)Tissue engineering—1989. BED (ed), vol 14, Asme, New York

  • Yacoub MH, Cohn LH (2004) Novel approaches to cardiac valve repair: from structure to function: part I. Circulation 109:942–950

    Article  Google Scholar 

Download references

Acknowledgments

This research was supported by NIH Grants HL-068816, HL-089750, HL-070969, and HL-108330. The authors would like to thank Thanh V. Lam for the development of the flexure-testing device and Brett Zubiate for the later improvements made to the transmural strain system. Also, recognition goes to Kristen Feaver for her contribution of the bimodular schematic (Fig. 8).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Michael S. Sacks.

Appendices

Appendix 1

1.1 Transmural deformation analysis

The images captured using the micro-imaging system (Fig. 2d) were analyzed by locating the markers that had been airbrushed onto the edge of the tissue. A custom program was written in LabVIEW (National Instruments, Austin, TX) to post-process the images taken by the micro-imaging system so that displacement fields could be determined. The markers were identified, numbered, and their areas and centroid coordinates were determined. This procedure was performed simultaneously for the reference image and for the deformed image. The software then displayed both altered images concurrently so that the user of the program could match markers between the reference image and the deformed image.

From the resulting images, the coordinates of the reference markers were referred to as the (\(X_{1}, X_{2})\) system, and the deformed coordinates were referred to as the (\(x_{1}, x_{2})\) system. The displacements, \(u\) and \(v\), were calculated from the former using \(u=x_1 -X_1 \) and \(v=x_2 -X_2\), respectively. These quantities were then fitted to the surface described by Eq. (5).

$$\begin{aligned} u&= a_0 +a_1 X_1 + a_2 X_2 + a_3 X_1 X_2 +a_4 X_1^2 + a_5 X_2^2\nonumber \\ v&= b_0 + b_1 X_1 +b_2 X_2 + b_3 X_1 X_2 + b_4 X_1^2 +b_5 X_2^2 \end{aligned}$$
(5)

The surface fit to the \(u\) and \(v\) coordinates achieved an \(r^{2}\) value of approximately 0.9. A higher order fit could have been used resulting in a higher \(r^{2}\), this would produce a rough surface due to variations in marker location from thresholding. The lower order fit maintains a smooth surface, true to the nature of the sample tested, by not overfitting the curve to all variations in marker location. By evaluating Eq. (6), the deformation gradient was obtained.

$$\begin{aligned} \mathbf{F}=\mathbf{H}+\mathbf{I}=\left[ {{\begin{array}{ll} {\frac{\partial u }{\partial X_1 }}&{} {\frac{\partial u }{\partial X_2 }} \\ {\frac{\partial v}{\partial X_1 }}&{} {\frac{\partial v}{\partial X_2 }} \\ \end{array} }} \right] +\left[ {{\begin{array}{ll} 1&{} 0 \\ 0&{} 1 \\ \end{array} }} \right] =\left[ {{\begin{array}{ll} {\lambda _1 }&{} {k _1 } \\ {k _2 }&{} {\lambda _2 } \\ \end{array} }} \right] \end{aligned}$$
(6)

The F was then decomposed into its stretch and rotation tensors, U and R, respectively, Eq. (7). The polar decomposition of the deformation tensor removes rigid body motion effects into the rotation tensor, leaving only the stretch deformation information in the stretch tensor. The rigid body rotation information in R was calculated to determine the degree of rotation experienced by the tissue during flexure. Higher levels of rigid body rotation were determined to be coincident with measurements taken away from the center of the tissue.

$$\begin{aligned}&\mathbf{F}=\mathbf{RU}\nonumber \\&\mathbf{U}^{2}=\mathbf{F}^\mathbf{T}\mathbf{F}\nonumber \\&\mathbf{R}=\mathbf{FU}^{-1} \end{aligned}$$
(7)

The stretch tensor components \(U_{11}\) and \(U_{22 }\) correspond to local tissue strains in the \(X_{1}\) and \(X_{2}\) directions, respectively. Thus, the location of the neutral axis was determined by plotting \(U_{11}\) against the thickness of the tissue. The depth of the tissue that coincided with the \(U_{11}\) value of unity was the corresponding location of the neutral axis. Rotation that occurred in the displacement field was characterized by determining the angle of rotation, \(\alpha \), incurred in the deformed system from the reference state.

Appendix 2

1.1 Parametric interlayer bonding study supplement

The reported results for the parametric bonding study represented a curvature change of \(0.2\,\hbox {mm}^{-1 }\) solely for sake of clarity, as the same trends were observed at curvature changes of 0.1, 0.2 and 0.3 mm\(^{-1}\) (Fig. 9a–c). These choices of curvature change were taken from our in vitro measurements (Sugimoto and Sacks 2013). Not surprisingly, we noted that the presence of interlayer sliding (not magnitude) either simulated (Fig. 9) or experimentally (Fig. 10) was not a function of the level of bending (i.e., \(\Delta \kappa )\), but only of the ratio of the ventricularis and fibrosa:spongiosa moduli for the simulations. Thus, the estimated \(\mu _{S}\) threshold is independent of imposed curvature. Greater bending simply created greater sliding.

Fig. 9
figure 9

Transmural deformation results of parametric bonding simulation, \(\varLambda _{1}\) is plotted against the normalized leaflet thickness for a curvature change of 0.1, 0.2, and 0.3 mm\(^{-1}\). A tri-layered rectangular strip represented the AV and shear moduli values ranging from 1.0 Pa to 45 kPa were assigned to the central spongiosa layer to emulate varying degrees of connectivity between the outside layers and identify the following relationship: \(\mu _{F}=\mu _{V}:\mu _{S}\). Results indicate for all curvature levels that for measurable sliding to occur between the fibrosa and ventricularis, the spongiosa must possess a shear modulus less than 1 kPa

Fig. 10
figure 10

Experimental results obtained from transmural bending tests performed on native aortic valve tissue. The tissue was bent to three different changes of curvature, 0.1, 0.2, and 0.3 mm\(^{-1}\). As curvature increased, the deformation increased as expected, yet no sliding is observed

1.2 Parametric out-of-plane warping study supplement

To investigate the effects of leaflet geometry on the simulation findings of out-of-plane warping effects, a parametric simulation was performed varying the thickness of the leaflets as well as the curvature change. The specimen length and width remained constant for the simulations. Figure 7 demonstrates the significant change in net axial stretch between the center of the specimen and the edge. Therefore, this change in absolute axial stretch (\(\varLambda _{1}\)) was used as a metric of warping and plotted against the change in specimen thickness (Fig. 11). Results found that increasing thickness of the specimens exaggerated the degree of warping. Furthermore, as expected, this warping effect increases with increasing curvature change.

Fig. 11
figure 11

The effect of leaflet thickness on the out-of-plane warping estimated by the simulation (Fig. 7). The degree of warping, measured by the absolute change in net axial stretch (\(\varLambda _{1})\), increases with increasing leaflet thickness. Additionally, this relationship is maintained and exaggerated with increasing curvature (0.1, 0.2 and 0.3 mm\(^{-1})\)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Buchanan, R.M., Sacks, M.S. Interlayer micromechanics of the aortic heart valve leaflet. Biomech Model Mechanobiol 13, 813–826 (2014). https://doi.org/10.1007/s10237-013-0536-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10237-013-0536-6

Keywords

Navigation