Skip to main content
Log in

Modeling function–perfusion behavior in liver lobules including tissue, blood, glucose, lactate and glycogen by use of a coupled two-scale PDE–ODE approach

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

Abstract

This study focuses on a two-scale, continuum multicomponent model for the description of blood perfusion and cell metabolism in the liver. The model accounts for a spatial and time depending hydro-diffusion–advection–reaction description. We consider a solid-phase (tissue) containing glycogen and a fluid-phase (blood) containing glucose as well as lactate. The five-component model is enhanced by a two-scale approach including a macroscale (sinusoidal level) and a microscale (cell level). The perfusion on the macroscale within the lobules is described by a homogenized multiphasic approach based on the theory of porous media (mixture theory combined with the concept of volume fraction). On macro level, we recall the basic mixture model, the governing equations as well as the constitutive framework including the solid (tissue) stress, blood pressure and solutes chemical potential. In view of the transport phenomena, we discuss the blood flow including transverse isotropic permeability, as well as the transport of solute concentrations including diffusion and advection. The continuum multicomponent model on the macroscale finally leads to a coupled system of partial differential equations (PDE). In contrast, the hepatic metabolism on the microscale (cell level) was modeled via a coupled system of ordinary differential equations (ODE). Again, we recall the constitutive relations for cell metabolism level. A finite element implementation of this framework is used to provide an illustrative example, describing the spatial and time-depending perfusion–metabolism processes in liver lobules that integrates perfusion and metabolism of the liver.

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
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18

Similar content being viewed by others

References

  • Ateshian GA (2007) On the theory of reactive mixtures for modeling biological growth. Biomech Model Mechanobiol 6:423–445

    Article  Google Scholar 

  • Ateshian GA, Costa KD, Azeloglu EU, III BM, Hung CT (2009) Continuum modeling of biological tissue growth by cell division, and alteratin of intracellular osmolytes and extracellular fixed charge density. J Biomech Eng 131(10):101001. doi:10.1115/1.3192138

  • Babcock M, Cardell J (1974) Hepatic glycogen patterns in fasted and fed rats. Am J Anat 140:299–337

    Article  Google Scholar 

  • Biot MA (1935) Le probl\(\acute{e}\)me de la consolidation des mati\(\grave{e}\)res argileuses sous une charge. Annales de la Societ\(\acute{e}\) scientifique de Bruxelles 55:110–113

  • Biot MA (1941) General theory of three dimensional consolidation. J Appl Phys 12:155–164

    Article  MATH  Google Scholar 

  • Blouin A, Bolender RP, Weibel ER (1977) Distribution of organelles and membranes between hepatocytes and nonhepatocytes in the rat liver parenchyma. A stereological study. J Cell Biol 72(2):441–455

  • Bluhm J (2002) Modelling of saturated thermo-elastic porous solids with different phase temperatures. In: Ehlers W, Bluhm J (eds) Porous media: theory, experiments and numerical applications. Springer, Berlin

    Google Scholar 

  • Bowen RM (1976a) Theory of mixtures. In: Eringen AC (ed) Continuum physics, vol III. Academic Press, New York, pp 1–127

    Google Scholar 

  • Bowen RM (1976b) Theory of mixtures. In: Eringen AC (ed) Continuum physics. Academic Press, New York

  • Brown JD, Rosen J, Kim YS, Chang L, Sinanan M, Hannaford B (2003) In-vivo and in-situ compressive properties of porcine abdominal soft tissues. In: Westwood JD, Hoffman HM, Mogel GT, Phillips R, Robb RA, Stredney D (eds) Medicine meets virtual reality 11, vol 94. IOS Press, Amsterdam, pp 26–32

    Google Scholar 

  • Burt A, Portmann B, Ferrell L (2007) MacSweens pathology of the liver. Elsevier, Churchill Livingstone

    Google Scholar 

  • Coleman BD, Noll W (1963) The thermodynamics of elastic materials with heat conduction and viscosity. Arch Ration Mech Anal 13:167–178

    Article  MATH  MathSciNet  Google Scholar 

  • Corrin B, Aterman K (1968) The pattern of glycogen distribution in the liver. Am J Anat 122:57–72

    Article  Google Scholar 

  • de Boer R (1996) Highlights in the historical development of the porous media theory: toward a transistent macroscopic theory. Appl Mech Rev 4:201–262

    Article  Google Scholar 

  • de Boer R (2000) Theory of porous media: highlights in the historical development and current state. Springer, Berlin

    Book  Google Scholar 

  • Debbaut C, Vierendeels J, Siggers JH, Repetto R, Monbaliu D, Segers P (2014) A 3d porous media liver lobule model: the importance of vascular septa and anisotropic permeability for homogeneous perfusion. Comput Methods Biomech Biomed Eng 17(12):1295–1310

  • Dirsch O, Madrahimov N, Chaudri N, Deng M, Madrahimova F, Schenk A, Dahmen U (2008) Recovery of liver perfusion after focal outflow obstruction and liver resection. Transplantation 85:748–756

  • Ehlers W (2002) Foundations of multiphasic and porous materials. In: Ehlers W, Bluhm J (eds) Porous media: theory, experiments and numerical applications. Springer, Berlin, pp 3–86

    Chapter  Google Scholar 

  • Eipper G (1998) Theorie und Numerik finiter elastischer Deformationen in fluidgesättigten Porösen Medien. Dissertation, Bericht Nr. II-1 aus dem Institut für Mechanik (Bauwesen), Universität Stuttgart

  • Evans DW, Moran EC, Baptista PM, Soker S, Sparks J (2013) Scale-dependent mechanical properties of native and decellularized liver tissue. Biomech Model Mechanobiol 12(3):569–580

    Article  Google Scholar 

  • Garikipati K, Olberding JE, Narayanan H, Arruda EM, Grosh K, Calve S (2006) Biological remodelling: stationary energy, configurational change, internal variables and dissipation. J Mech Phys Solids 54(7):1493–1515

    Article  MATH  MathSciNet  Google Scholar 

  • Gerich J (1993) Control of glycaemia. Baillieres Clin Endocrinol Metab 7(3):551–86

    Article  Google Scholar 

  • Greve R (2003) Kontinuumsmechanik: Ein Grundkurs. Springer, Berlin

    Book  Google Scholar 

  • Grytz R, Girkin CA, Libertiaux V, Downs JC (2012a) Perspectives on biomechanical growth and remodeling mechanisms in glaucoma. Mech Res Commun 42:92–106

    Article  Google Scholar 

  • Grytz R, Sigal IA, Ruberti JW, Meschke G, Downs C (2012b) Lamina cribrosa thickening in early glaucoma predicted by a microstructure motivated growth and remodeling approach. Mech Mater 44:99–109

    Article  Google Scholar 

  • Hall C (2006) Regulation der hepatischen mikrozirkulation bei schrittweiser resektion im rattenlebermodell. PhD thesis, Medizinische Fakultät der Universität Duisburg-Essen, Zentrum für Chirurgie Klinikum für Allgemeinchirurgie, Viszeral-und Transplantationschirurgie

  • Hariton I, deBotton G, Gasser TC, Holzapfel GH (2007) Stress-driven collagen fiber remodeling in arterial walls. Biomech Model Mechanobiol 6(3):163–175

    Article  MathSciNet  Google Scholar 

  • Humphrey JD, Rajagopal KR (2002) A constrained mixture model for growth and remodelling of soft tissues. Math. Models Methods Appl Sci 12(3):407–430

    Article  MATH  MathSciNet  Google Scholar 

  • Hutter K, Jöhnk K (2004) Continuum methods of physical modeling: continuum mechanics, dimensional analysis, turbulence. Springer, Berlin

    Book  Google Scholar 

  • Job G, Herrmann F (2006) Chemical potentiala quantity in search of recognition. Eur J Phys 27:353–371

    Article  Google Scholar 

  • Jungermann K (1986) Functional heterogeneity of periportal and perivenous hepatocytes. Enzyme 35:161–180

    Google Scholar 

  • Kerdok AE (2006) Characterizing the nonlinear mechanical response of liver to surgical manipulation. PhD thesis, Harvard University, Engineering and Applied Sciences

  • König M, Bulik S, Holzhütter H (2012) Quantifying the contribution of the liver to glucose homeostasis: a detailed kinetic model of human hepatic glucose metabolism. PLoS Comput Biol 8(6):e1002,577

    Article  Google Scholar 

  • Kuhl E, Holzapfel GA (2007) A continuum model for remodeling in living structures. J Mater Sci 42:8811–8823

    Article  Google Scholar 

  • Kuntz E, Kuntz HD (2006) Hepatology: principles and practice. Springer, Berlin

    Google Scholar 

  • Lai WM, Hou JS, Mow VC (1991) A triphasic theory for the swelling and deformation behaviours of articular cartilage. ASME J Biomech Eng 113:245–258

    Article  Google Scholar 

  • Lautt WW (2009) Hepatic circulation. University of Manitoba, San Rafael

    Google Scholar 

  • Lu JF, Hanyga A (2005) Linear dynamic model for porous media saturated by two immiscible fluids. Int J Solids Struct 42:2689–2709

    Article  MATH  Google Scholar 

  • Maass-Moreno R, Rothe CF (1997) Distribution of pressure gradients along hepatic vasculature. Am J Physiol Heart Circ Physiol 272(6):H2826–H2832. http://ajpheart.physiology.org/content/272/6/H2826.full.pdf

  • MacPhee PJ, Schmidt EE, Groom AC (1995) Intermittence of blood flow in liver sinusoids, studied by high-resolution in vivo microscopy. Am J Physiol Gastrointest Liver Physiol 269(5):G692–G698. http://ajpgi.physiology.org/content/269/5/G692.full.pdf

  • Millonig G, Friedrich S, Adolf S, Fonouni H, Golriz M, Mehrabi A, Stiefel P, Pschl G, Bchler MW, Seitz HK, Mueller S (2010) Liver stiffness is directly influenced by central venous pressure. J Hepatol 52(2):206–210

    Article  Google Scholar 

  • Mow VC, Gibbs MC, Lai WM, Zhu WB, Athanasiou KA (1989) Biphasic indentation of articular cartilage-ii. A numerical algorithm and an experimental study. J Biomech 22:853–861

    Article  Google Scholar 

  • Nava A, Mazza E, Furrer M, Villiger P, Reinhart WH (2008) In vivo mechanical characterization of human liver. Med Image Anal 12(2):203–216

    Article  Google Scholar 

  • Nuttal FQ, Ngo A, Gannon MC (2008) Regulation of hepatic glucose production and the role of gluconeogenesis in humans: is the rate of gluconeogenesis constant? Diabetes Metab Res Rev 24:438–458

    Article  Google Scholar 

  • Pierce DM, Ricken T, Holzapfel GA (2013a) A hyperelastic biphasic fibre-reinforced model of articular cartilage considering distributed collagen fibre orientations: continuum basis, computational aspects and applications. Comput Methods Biomech Biomed Eng 16(12):1344–1361

    Article  Google Scholar 

  • Pierce DM, Ricken T, Holzapfel GA (2013b) Modeling sample/patient-specific structural and diffusional responses of cartilage using DT-MRI. Int J Numer Methods Biomed Eng 29(8):807–821

    Article  Google Scholar 

  • Radziuk J, Pye S (2001) Hepatic glucose uptake, gluconeogenesis and the regulation of glycogen synthesis. Diabetes Metab Res Rev 17:250–272

    Article  Google Scholar 

  • Ricken T (2002) Kapillarität in porösen medien—theoretische untersuchung und numerische simulation. PhD thesis, Dissertation, Shaker Verlag, Aachen

  • Ricken T, Bluhm J (2009) Evolutional growth and remodeling in multiphase living tissue. Comput Mater Sci 45(3):806–811

    Article  Google Scholar 

  • Ricken T, Bluhm J (2010) Remodeling and growth of living tissue: a multiphase theory. Arch Appl Mech 80(5):453–465

    Article  MATH  Google Scholar 

  • Ricken T, de Boer R (2003) Multiphase flow in a capillary porous medium. Comput Mater Sci 28:704–713

    Article  Google Scholar 

  • Ricken T, Dahmen U, Dirsch O (2010) A biphasic model for sinusoidal liver perfusion remodeling after outflow obstruction. Biomech Model Mechanobiol 9:435–450

    Article  Google Scholar 

  • Saitoh Y, Terada N, Saitoh S, Ohno N, Fujii Y, Ohno S (2010) Histochemical approach of cryobiopsy for glycogen distribution in living mouse livers under fasting and local circulation loss conditions. Histochem Cell Biol 133(2):229–239

    Article  Google Scholar 

  • Samur E, Sedef M, Basdogan C, Avtan L, Duzgun O (2005) A robotic indenter for minimally invasive characterization of soft tissues. In: International congress series, CARS 2005: computer assisted radiology and surgery proceedings of the 19th international congress and exhibition, vol 1281, no 0, pp 713–718

  • Sasse D (1975) Dynamics of liver glycogen: the topochemistry of glycogen synthesis, glycogen content and glycogenolysis under the experimental conditions of glycogen accumulation and depletion. Histochemistry 45:237–254

  • Schwartz JM, Denninger M, Rancourt D, Moisan C, Laurendeau D (2005) Modelling liver tissue properties using a non-linear visco-elastic model for surgery simulation. Med Image Anal 9(2):103–112

  • Siggers JH, Leungchavaphongse K, Ho CH, Repetto R (2014) Mathematical model of blood and interstitial flow and lymph production in the liver. Biomec Model Mechanobiol 13(2):363–378

  • Stanhope KL, Griffen SC, Bair BR, Swarbrick MM, Keim NL, Havel PJ (2008) Twenty-four-hour endocrine and metabolic profiles following consumption of high-fructose corn syrup-, sucrose-, fructose-, and glucose-sweetened beverages with meals. Am J Clin Nutr 87:1194–1203

    Google Scholar 

  • Teutsch HF (2005) The modular microarchitecture of human liver. Hepatology 42(2):317–325

    Article  Google Scholar 

  • Teutsch HF, Schuerfeld D, Groezinger E (1999) Three-dimensional reconstruction of parenchymal units in the liver of the rat. Hepatology 29(2):494–505

    Article  Google Scholar 

  • Truesdell C (1984) Thermodynamics of diffusion. In: Truesdell C (ed) Rational thermodynamics. Springer, New York, pp 219–236

    Chapter  Google Scholar 

  • Truesdell C, Toupin R (1960) The classical field theory. In: Flügge S (ed) Handbuch der Physik, vol 3/1. Springer, Berlin

    Google Scholar 

  • Valentin A, Humphrey JD, Holzapfel GA (2013) A finite element-based constrained mixture implementation for arterial growth, remodeling, and adaptation: theory and numerical verification. Int J Numer Methods Biomed Eng 29(8):822–849

    Article  MathSciNet  Google Scholar 

  • Villeneuve JP, Dagenais PM, Huet P, Roy A, Lapointe R, Marleau D (1996) The hepatic microcirculation in the isolated perfused human liver. Hepatology 23(1):24–31

    Article  Google Scholar 

  • Wahren J, Ekberg K (2007) Splanchnic regulation of glucose production. Annu Rev Nutr 27:329–345

    Article  Google Scholar 

  • Walpole J, Papin J, Peirce S (2013) Multiscale computational models of complex biological systems. Ann Rev Biomed Eng 15:137–154

    Article  Google Scholar 

  • Warren A, Chaberek S, Ostrowski K, Cogger VC, Hilmer SN, McCuskey RS, Fraser R, Le Couteur DG (2008) Effects of old age on vascular complexity and dispersion of the hepatic sinusoidal network. Microcirculation 15(3):191–202

    Article  Google Scholar 

  • Werner D, Ricken T, Dahmen U, Dirsch O (2012) A biphasic fem model for the microperfusion in liver lobules. PAMM 12:89–90

    Article  Google Scholar 

  • Yuen K, McDaniel P, Riddle M (2012) Twenty-four-hour profiles of plasma glucose, insulin, c-peptide and free fatty acid in subjects with varying degrees of glucose tolerance following short-term, medium-dose prednisone (20 mg/day) treatment: evidence for differing effects on insulin secretion and action. Clin Endocrinol 77:224–232

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to T. Ricken.

Appendix

Appendix

1.1 Kinematics

The saturated porous solid is treated as an immiscible mixture of all constituents \(\varphi ^{\varvec{\alpha }}\) with particles \(X_{{\varvec{\alpha }}}\) with an own independent motion function

$$\begin{aligned} \mathbf x ={\varvec{\chi }}_{\small {\varvec{\alpha }}}(\mathbf{X}_{\varvec{\alpha }},\mathrm{t})\>,\quad \mathbf{X}_{\varvec{\alpha }}={\varvec{\chi }}_{\small {\varvec{\alpha }}}^{\mathrm{-1}}(\mathbf{x},\mathrm{t})\>. \end{aligned}$$
(40)

where (40)\(_1\) represents the Lagrange description of motion. The function \({\varvec{\chi }}_{\small {\varvec{\alpha }}}\) is postulated to be unique and uniquely invertible at any time t. The existence of a function inverse to (40)\(_1\) leads to the Euler description of motion, see (40)\(_2\). A mathematical condition, which is necessary and sufficient for the existence of equation (40)\(_2\), is given if the Jacobian \(\mathrm J_{\varvec{\alpha }}=\mathrm{det}\>\mathbf{F}_{\varvec{\alpha }}\) differs from zero. Therein, \(\mathbf{F}_{\varvec{\alpha }}\) is the deformation gradient. The tensor \(\mathbf{F}_{\varvec{\alpha }}\) and its inverse \(\mathbf{F}_{\varvec{\alpha }}^{\mathrm{-1}}\) are defined as \(\mathbf{F}_{\varvec{\alpha }}=\partial \mathbf{x}/\partial \mathbf{X}_{\varvec{\alpha }}=\mathrm{Grad}_{{\varvec{\alpha }}}\>{\varvec{\chi }}_{\small {\varvec{\alpha }}}\) and \(\mathbf{F}_{\varvec{\alpha }}^{\mathrm{-1}}=\partial \mathbf{X}_{\varvec{\alpha }}/\partial \mathbf{x}= \mathrm{grad}\>\mathbf{X}_{\varvec{\alpha }}\). The differential operator “\(\mathrm{Grad}_{{\varvec{\alpha }}}\)” is referring to a partial differentiation with respect to the reference position \(\mathbf{X}_{\varvec{\alpha }}\) of the constituent \(\varphi ^{\varvec{\alpha }}\) and the differential operator \(``\mathrm{grad}\)” referring to the spatial point \(\mathbf{x}\). During the deformation process, \(\mathbf{F}_{\varvec{\alpha }}\) is restricted to \(\mathrm{det}\>\mathbf{F}_{\varvec{\alpha }}> \mathrm 0\). The spatial velocity gradient \(\mathbf{L}_{\varvec{\alpha }}= ( \, \mathrm{Grad}_{\varvec{\alpha }}\>\mathbf{x}^\prime _{\varvec{\alpha }}\, ) \, \mathbf{F}_{\varvec{\alpha }}^\mathrm{-1} = \mathrm{grad}\>\mathbf{x}^\prime _{\varvec{\alpha }}\), where \(( \mathbf{F}_{\varvec{\alpha }})^\prime _{\varvec{\alpha }}= \partial \mathbf{x}^\prime _{\varvec{\alpha }}/\partial \mathbf{X}_{\varvec{\alpha }}= \mathrm{Grad}_{\varvec{\alpha }}\>\mathbf{x}^\prime _{\varvec{\alpha }}\) denotes the material velocity gradient, can be additively decomposed into a symmetric part \(\mathbf{D}_{\varvec{\alpha }}= ( \, {\mathbf{L}_{{\varvec{\alpha }}}} \, + \, {\mathbf{L}_{{\varvec{\alpha }}}^\mathrm{T}} \, )\,/\, 2\) and a skew-symmetric part \({\mathbf{W}_{{\varvec{\alpha }}}} = ( \, {\mathbf{L}_{{\varvec{\alpha }}}} \, - \, {\mathbf{L}_{{\varvec{\alpha }}}^\mathrm{T}} \, ) \,/\,2\) with \(\mathbf{L}_{\varvec{\alpha }}= \mathbf{D}_{\varvec{\alpha }}+ \mathbf{W}_{\varvec{\alpha }}\).

With the Lagrange description of motion (40)\(_1\), the velocity and acceleration fields of the constituents \(\varphi ^{\varvec{\alpha }}\) are defined as material time derivatives of the motion function (40)\(_1\), see Fig. 19,

$$\begin{aligned} \mathbf{x}^{\prime }_{{\varvec{\alpha }}}=\frac{\partial {\varvec{\chi }}_{\small {\varvec{\alpha }}}(\mathbf{X}_{\varvec{\alpha }},\mathrm{t})}{\partial \mathrm{t}}\>,\quad {\mathbf{x}}''_{\varvec{\alpha }}=\frac{\partial ^{\mathrm{2}}{\varvec{\chi }}_{\small {\varvec{\alpha }}}(\mathbf{X}_{\varvec{\alpha }},\mathrm{t})}{\partial \mathrm{t}^{\mathrm{2}}}\>. \end{aligned}$$
(41)

For scalar fields depending on \({\mathbf{x}}\) and t, the material time derivatives are defined as

$$\begin{aligned} (...)^\prime _{\varvec{\alpha }}=\partial (...)/\partial \mathrm{t}+\mathrm{grad}\>(...)\cdot {\mathbf{x}}'_{\varvec{\alpha }}, \end{aligned}$$
(42)

see, e.g., de Boer (2000).

Fig. 19
figure 19

Motion of solid and fluid particle in a fluid saturated porous solid

1.2 Constitutive theory

With respect to the assumptions made in Sect. 3, the material time derivative of the saturation condition (24) is an equation in excess that restricts the motion of the incompressible constituents. Therefore, the set of unknown field quantities must be extended by a scalar, namely the Lagrange multiplier \(\bar{\lambda }\), which is understood as an indeterminate reaction force due to the saturation condition. Thus, the set of unknown field quantities is:

$$\begin{aligned} {\mathcal {U}}=\left\{ \,\varvec{\chi }_{\varvec{\alpha }},\>{\mathrm{c}}^{\alpha \beta },\>\bar{\lambda }\,\right\} \end{aligned}$$
(43)

Considering that the external acceleration \(\mathbf{b}\), i.e., the acceleration of gravity, is known, the remaining quantities

$$\begin{aligned} {\mathcal {C}}=\left\{ \,\mathbf{T}^{\varvec{\alpha }},\>\hat{\mathbf{p}}^\mathrm{F},\>\hat{\mathbf{p}}^{\alpha \beta },\>{\hat{\mathrm{c}}}^{\alpha \beta }\,\right\} \end{aligned}$$
(44)

in the set of field equations, see Sect. 4, require constitutive relations in order to close the system of equations. In view of the treatment of the entropy inequality in analogy to Coleman and Noll (1963) the following set of free but not overall independent process variables is chosen:

$$\begin{aligned}&\mathcal {P}=\left\{ \,\mathbf{C}_\mathrm{S},{\mathrm{n}}^{\mathrm{S}},{\mathrm{n}}^{\mathrm{F}},\mathbf{w}_\mathrm{FS},\mathbf{w}_{\mathrm{F} \beta S},{\mathrm{c}}^{\mathrm{S}\beta },\mathrm{c}^{\mathrm{F}\beta },\mathrm{grad}\mathrm{n}^\mathrm{F},\right. \nonumber \\&\quad \left. \mathrm{grad}\mathrm{c}^\mathrm{S\beta },\mathrm{grad}\mathrm{c}^{\mathrm{F}\beta }\right\} \end{aligned}$$
(45)

In (45) the right Cauchy-Green deformation tensor \(\mathbf{C}_\mathrm{S}=\mathbf{F}_\mathrm{S}^\mathrm{T}\,\mathbf{F}_\mathrm{S}\) is considered. The entropy inequality for the mixture body

$$\begin{aligned}&\displaystyle \sum _{{\varvec{\alpha }}}^{\mathrm{S,F,\alpha \beta }}\left\{ \,-\,\rho ^{\varvec{\alpha }}\,(\psi ^{\varvec{\alpha }})^\prime _{\varvec{\alpha }}-\hat{\rho }^{\varvec{\alpha }}\,(\,\psi ^{\varvec{\alpha }}-\frac{1}{2}\,\mathbf{x}^\prime _{\varvec{\alpha }}\cdot \mathbf{x}^\prime _{\varvec{\alpha }}\,)\right. \nonumber \\&\quad \left. +\mathbf{T}^{\varvec{\alpha }}\cdot \mathbf{D}_{\varvec{\alpha }}+\hat{\mathrm{e}}^{\varvec{\alpha }}-\hat{\mathbf{p}}^{\varvec{\alpha }}\cdot \mathbf{x}^\prime _{\varvec{\alpha }}\,\right\} \ge \mathrm{0} \end{aligned}$$
(46)

will be rearranged to ensure no neglecting of dependencies which can influence constitutive modeling. In order to keep the complexity of the evaluation in a justifiable scope, the dependency of the Helmholtz free energies \(\psi ^{\varvec{\alpha }}\) on the process variables \(\mathcal {P}\) will be restricted as follows:

$$\begin{aligned} \psi ^\mathrm{S}=\psi ^\mathrm{S}\,\left\{ \,\mathbf{C}_\mathrm{S}\,\right\} \>,\quad \psi ^\mathrm{F}=\psi ^\mathrm{F}\,\{\,-\,\}\>, \quad \psi ^{\alpha \beta }=\psi ^{\alpha \beta }\,\{\,{{\mathrm{c}}^{\alpha \beta }}\,\}\>.\nonumber \\ \end{aligned}$$
(47)

For further investigations, the relations \(\rho ^{\varvec{\alpha }}\,(\psi ^{\varvec{\alpha }})^\prime _{\varvec{\alpha }}\) will be replaced by

$$\begin{aligned} \begin{array}{llllll} \rho ^{\mathrm{S}}\,(\psi ^{\mathrm{S}})^{\prime }_{\mathrm{S}}&{}=&{}\displaystyle 2\,{\mathrm{n}}^{\mathrm{S}}\,\rho ^{\mathrm{SR}}\,\mathbf{F}_{\mathrm{S}}\,\frac{\partial \psi ^{\mathrm{S}}}{\partial \mathbf{C}_{\mathrm{S}}}\,\mathbf{F}_{\mathrm{S}}^{\mathrm{T}}\cdot \mathbf{D}_{\mathrm{S}},\\ \rho ^{\mathrm{F}}\,(\psi ^{\mathrm{F}})^{\prime }_{\mathrm{F}}&{} = &{} 0,\\ \rho ^{\alpha \beta }\,(\psi ^{\alpha \beta })^{\prime }_{\beta }&{}=&{}\displaystyle {\rho ^{\alpha \beta }\,\frac{\partial \psi ^{\alpha \beta }}{\partial {\mathrm{c}}^{\alpha \beta }}\,(\mathrm{c}^{\alpha \beta })^{\prime }_{\beta }}\>. \end{array} \end{aligned}$$
(48)

For saturated porous media consisting of incompressible constituents the saturation condition (20) is a constraint with respect to the overall volumetric deformation. Therefore, the saturation condition must be considered in view of the evaluation of the entropy inequality. Here, the material time derivative of the saturation condition following the motion of the solid phase with

$$\begin{aligned} (\mathrm{n}^{\mathrm{S}})^{\prime }_{\mathrm{S}}+ (\mathrm{n}^{\mathrm{F}})^{\prime }_{\mathrm{S}}= (\mathrm{n}^{\mathrm{S}})^{\prime }_{\mathrm{S}}+ (\mathrm{n}^{\mathrm{F}})^{\prime }_{\mathrm{F}}- \mathrm{grad}\, {\mathrm{n}}^{\mathrm{F}}\cdot \mathbf{w}_{\mathrm{FS}}= \mathrm 0 \end{aligned}$$
(49)

will be considered, where (42) and the difference velocity

$$\begin{aligned} \mathbf{w}_{\mathrm{FS}}= \mathbf{x}^{\prime }_{\mathrm{F}}- \mathbf{x}^{\prime }_{\mathrm{S}}\end{aligned}$$
(50)

are used. Applying the identity

$$\begin{aligned} \begin{array}{lll} (\mathrm{n}^{\alpha })^{\prime }_{\alpha }&{} = &{} \displaystyle {\frac{{\mathrm{n}}^{\alpha }}{\rho ^{\alpha }}} \, (\rho ^{\alpha })^{\prime }_{\alpha }- \displaystyle {\frac{{\mathrm{n}}^{\alpha }}{\rho ^{\alpha \mathrm{R}}}} \, (\rho ^{\alpha })^{\prime }_{\alpha }\\ &{} = &{} - \, {\mathrm{n}}^{\alpha }\, ( \, \mathbf{D}_{\alpha }\, \cdot \, \mathbf{I}\, ) -\displaystyle {\frac{{\mathrm{n}}^{\alpha }}{\rho ^{\alpha \mathrm{R}}}}\,\underbrace{(\rho ^{\alpha \mathrm{R}})^{\prime }_{\alpha }}_\mathrm{=\,0},\\ &{} = &{} - \, {\mathrm{n}}^{\alpha }\, ( \, \mathbf{D}_{\alpha }\, \cdot \, \mathbf{I}) \end{array} \end{aligned}$$
(51)

the rate of the saturation condition (49) can be rearranged to

$$\begin{aligned} {\mathrm{n}}^{\mathrm{S}}\, ( \, \mathbf{D}_{\mathrm{S}}\cdot \mathbf{I}\, ) + {\mathrm{n}}^{\mathrm{F}}\, ( \, \mathbf{D}_{\mathrm{F}}\cdot \, \mathbf{I}\, ) + \mathrm{grad}\, {\mathrm{n}}^{\mathrm{F}}\, \cdot \, \mathbf{w}_{\mathrm{FS}}= \mathrm 0. \end{aligned}$$
(52)

In view of the constitutive modeling, we use the concept of Lagrange multipliers. Using the identity \(\mathrm{div}\>\mathbf{x}^{\prime }_{\alpha }=\mathbf{D}_{\alpha }\cdot \mathbf{I}\), we add (21)\(_\mathrm{1,2}\) to the entropy inequality (46), multiplied by the Lagrange multiplier \(\bar{\lambda }\), cf. de Boer (1996). The interconnection between the spatial velocity deformation gradients \(\mathbf{D}_{{\varvec{\alpha }}}\) with the volume fractions and their material time derivative \({\mathrm{n}}^{\alpha }\) and \((\mathrm{n}^{\alpha })^{\prime }_{\alpha }\) as well as the mass supplies \(\hat{\rho }^{\alpha }\) will be considered with the balance equations of mass (14)\(_1\) multiplied with a respective Lagrange multiplier:

$$\begin{aligned}&\bar{\lambda }^\mathrm{S}\,\left[ \,(\mathrm{n}^{\mathrm{S}})^{\prime }_{\mathrm{S}}+{\mathrm{n}}^{\mathrm{S}}\>(\mathbf{D}_{\mathrm{S}}\cdot \mathbf{I})\,\right] =\mathrm{0}\nonumber \\&\bar{\lambda }^\mathrm{F}\,\left[ \,(\mathrm{n}^{\mathrm{F}})^{\prime }_{\mathrm{F}}+{\mathrm{n}}^{\mathrm{F}}\>(\mathbf{D}_{\mathrm{F}}\cdot \mathbf{I})\,\right] =\mathrm{0}\nonumber \\&{\bar{\lambda }}^{\mathrm{S}\beta }\,\left[ \,(\mathrm{n}^{\mathrm{S}})^{\prime }_{\mathrm{S}}\,{\mathrm{c}}^{\mathrm{S}\beta }+{\mathrm{n}}^{\mathrm{S}}\,({\mathrm{c}}^{\mathrm{S} \beta })^{\prime }_\mathrm{S}+{\mathrm{n}}^{\mathrm{S}}\,{\mathrm{c}}^{\mathrm{S}\beta }\,\mathbf{D}_{\mathrm{S}}\cdot \mathbf{I}-\hat{\mathrm{c}}^{\mathrm{S} \beta }\right] =\mathrm{0}\nonumber \\&\bar{\lambda }^\mathrm{F\beta }\,\left[ \,(\mathrm{n}^{\mathrm{F}})^{\prime }_{\mathrm{F}}\,\mathrm{c}^{\mathrm{F}\beta }+{\mathrm{n}}^{\mathrm{F}}\,(\mathrm{c}^{\mathrm{F} \beta })^{\prime }_{\beta }+\mathrm{grad}\>{\mathrm{n}^{\mathrm{F}}}(\mathbf{w}_{\mathrm{F}\beta \mathrm{S}}-\mathbf{w}_{\mathrm{FS}})\,\mathrm{c}^{\mathrm{F}\beta }\right. \nonumber \\&\quad \left. +\,{\mathrm{n}}^{\mathrm{F}}\,\mathrm{c}^{\mathrm{F}\beta }\,\mathbf{D}_{\mathrm{F} \beta }\cdot \mathbf{I}-{\hat{\mathrm{c}}}^{\mathrm{F} \beta }\right] =\mathrm{0}\nonumber \\&\quad -\,\bar{\lambda }\,\left[ \,-{\mathrm{n}}^{\mathrm{S}}\,\mathbf{D}_{\mathrm{S}}\cdot \mathbf{I}-{\mathrm{n}}^{\mathrm{F}}\,\mathbf{D}_{\mathrm{F}}\cdot \mathbf{I}-\mathrm{grad}\>{\mathrm{n}^{\mathrm{F}}}\,\mathbf{w}_{\mathrm{FS}}\right] =\mathrm{0} \end{aligned}$$
(53)

The entropy inequality (46), supplemented with the material time derivative of the saturation condition (53)\(_5\) and the balance equations of mass (53)\(_{1-4}\), all multiplied with the corresponding Lagrange multiplier, reads

$$\begin{aligned}&\mathbf{D}_{\mathrm{S}}\>\cdot \left\{ \,\displaystyle {\mathbf{T}^{\mathrm{S}}+\mathbf{T}^\mathrm{S \beta }-\mathrm{2}\,{{\mathrm{n}}^{\mathrm{S}}\,\rho ^{\mathrm{SR}}}\,{\mathbf{F}_{\mathrm{S}}}\,{\frac{\partial \psi ^{\mathrm{S}}}{\partial \mathbf{C}_{\mathrm{S}}}}\,{\mathbf{F}_{\mathrm{S}}^{\mathrm{T}}}}\right. \nonumber \\&\left. \quad \,+\displaystyle {\bar{\lambda }\,{\mathrm{n}}^{\mathrm{S}}\,\mathbf{I}\,+\bar{\lambda }^\mathrm{S}\,{\mathrm{n}}^{\mathrm{S}}\,\mathbf{I}\,+{\bar{\lambda }}^{\mathrm{S}\beta }{\mathrm{c}}^{\mathrm{S}\beta }\,{\mathrm{n}}^{\mathrm{S}}\,\mathbf{I}}\,\right\} \,\nonumber \\&\quad +\,\mathbf{D}_{\mathrm{F}}\>\cdot \left\{ \,\displaystyle {\mathbf{T}^{\mathrm{F}}}+\displaystyle {\bar{\lambda }\,{\mathrm{n}}^{\mathrm{F}}\,\mathbf{I}}+\displaystyle {\bar{\lambda }^\mathrm{F}\,{\mathrm{n}}^{\mathrm{F}}\,\mathbf{I}}\,\right\} \,\nonumber \\&\quad +\,\displaystyle {\mathbf{D}_{\mathrm{F} \beta }}\>\cdot \left\{ \,\displaystyle {\mathbf{T}^\mathrm{F \beta }+\displaystyle {\bar{\lambda }^\mathrm{F\beta }\,{\mathrm{n}}^{\mathrm{F}}\,\mathrm{c}^{\mathrm{F}\beta }\,\mathbf{I}}}\,\right\} \,\nonumber \\&\quad -\,(\mathrm{n}^{\mathrm{S}})^{\prime }_{\mathrm{S}}\left\{ \, -\bar{\lambda }^\mathrm{S}-{\bar{\lambda }}^{\mathrm{S}\beta }\,{\mathrm{c}}^{\mathrm{S}\beta }\right\} \,-(\mathrm{n}^{\mathrm{F}})^{\prime }_{\mathrm{F}}\left\{ \, -\bar{\lambda }^\mathrm{F}-\bar{\lambda }^\mathrm{F\beta }\,\mathrm{c}^{\mathrm{F}\beta }\right\} \,\nonumber \\&\quad -\,\displaystyle {(\mathrm{c}^{\mathrm{F}\beta })^{\prime }_{\beta }}\left\{ \,\displaystyle {\rho ^{\mathrm{F} \beta }\,\frac{\partial \psi ^{\mathrm{F} \beta }}{\partial \mathrm{c}^{\mathrm{F}\beta }}-\bar{\lambda }^\mathrm{F\beta }\,{\mathrm{n}}^{\mathrm{F}}} \right\} \,\nonumber \\&\quad -\,\displaystyle {({\mathrm{c}}^{\mathrm{S}\beta })^{\prime }_{\beta }}\left\{ \,\displaystyle {\rho ^{\mathrm{S} \beta }\,\frac{\partial \psi ^{\mathrm{S} \beta }}{\partial {\mathrm{c}}^{\mathrm{S}\beta }}-{\bar{\lambda }}^{\mathrm{S}\beta }\,{\mathrm{n}}^{\mathrm{S}}} \right\} \,\nonumber \\&\quad -\,\hat{\rho }^{\mathrm{S} \beta }\displaystyle {\left\{ \psi ^{\mathrm{S} \beta }-\frac{\mathrm{1}}{\mathrm{2}}\,\mathbf{x}^{\prime }_{\mathrm{S}}\cdot \mathbf{x}^{\prime }_{\mathrm{S}}+\frac{{\bar{\lambda }}^{\mathrm{S}\beta }}{\mathrm{M}^{\beta }_\mathrm{mol}}\right\} }\,\nonumber \\&\quad -\,\hat{\rho }^{\mathrm{F} \beta }\displaystyle {\left\{ \psi ^{\mathrm{F} \beta }-\frac{\mathrm{1}}{\mathrm{2}}\,\mathbf{x}^{\prime }_\mathrm{F \beta }\cdot \mathbf{x}^{\prime }_\mathrm{F \beta }+\frac{\bar{\lambda }^\mathrm{F\beta }}{\mathrm{M}^{\beta }_\mathrm{mol}}\right\} }\,\nonumber \\&\quad -\,\displaystyle {\mathbf{w}_{\mathrm{FS}}\>\cdot }\left\{ \,\displaystyle {{\hat{\mathbf{p}}}^{\mathrm{F}}-\bar{\lambda }\,\,\mathrm{grad}\>{\mathrm{n}^{\mathrm{F}}}+\bar{\lambda }^\mathrm{F\beta }\,\mathrm{c}^{\mathrm{F}\beta }\,\mathrm{grad}\>{\mathrm{n}^{\mathrm{F}}}}\,\right\} \,\nonumber \\&\quad -\,\displaystyle {\mathbf{w}_{\mathrm{F}\beta \mathrm{S}}\>\cdot }\left\{ \,\displaystyle {{\hat{\mathbf{p}}}^{\mathrm{F}\beta }-\bar{\lambda }^\mathrm{F\beta }\,\mathrm{c}^{\mathrm{F}\beta }\,\mathrm{grad}\>{\mathrm{n}^{\mathrm{F}}}}\,\right\} \,\ge 0\>, \end{aligned}$$
(54)

where use has been made of \(\sum \rho ^{\alpha \beta }= 0\) and \({\hat{\mathbf{p}}}^{\mathrm{S}}=-{\hat{\mathbf{p}}}^{\mathrm{F}}-{\hat{\mathbf{p}}}^{\mathrm{S}\beta }-{\hat{\mathbf{p}}}^{\mathrm{F}\beta }\), see (19). The inequality must hold for fixed values of the process variables, see (45), and for arbitrary values of the so-called free-available quantities \(\mathcal {A}=\displaystyle {\{\mathbf{D}_{{\varvec{\alpha }}},\>(\mathrm{n}^{\alpha })^{\prime }_{\alpha },\>(\mathrm{c}^{\alpha \beta })^{\prime }_{\alpha \beta }\}}\) which contains selective derivatives of the values contained in \(\mathcal {P}\). Thus, the entropy inequality (54) can be satisfied if the following structure is obtained:

$$\begin{aligned}&\displaystyle {\mathbf{D}_{\mathrm{S}}\cdot \left\{ \underbrace{(\ldots )}_{=\,\mathbf{0}}\right\} +\mathbf{D}_{\mathrm{F}}\cdot \left\{ \underbrace{(\ldots )}_{=\,\mathbf{0}}\right\} +\mathbf{D}_{\mathrm{F} \beta }\cdot \left\{ \underbrace{(\ldots )}_{=\,\mathbf{0}}\right\} }\,\nonumber \\&-(\mathrm{n}^{\mathrm{S}})^{\prime }_{\mathrm{S}}\,\left\{ \underbrace{(\ldots )}_{=\,\mathrm{0}}\right\} -(\mathrm{n}^{\mathrm{F}})^{\prime }_{\mathrm{F}}\,\left\{ \underbrace{(\ldots )}_{=\,\mathrm{0}}\right\} \,\\&-\displaystyle {(\mathrm{c}^{\mathrm{S} \beta })^{\prime }_\mathrm{S\beta }\,\left\{ \underbrace{(\ldots )}_{=\,\mathrm{0}}\right\} \,-(\mathrm{c}^{\mathrm{F} \beta })^{\prime }_\mathrm{F\beta }\,\{\underbrace{(\ldots )}_{=\,\mathrm{0}}\}\,+}\,\,\displaystyle {\underbrace{\mathcal {D}is}_{\ge {\mathrm{\,0}}}\,}\,\ge 0\>,\nonumber \end{aligned}$$
(55)

where the dissipation mechanism

$$\begin{aligned} \mathcal {D}is&= -\hat{\rho }^{\mathrm{S} \beta }\displaystyle {\left\{ \psi ^{\mathrm{S} \beta }-\frac{\mathrm{1}}{\mathrm{2}}\,\mathbf{x}^{\prime }_{\mathrm{S}}\cdot \mathbf{x}^{\prime }_{\mathrm{S}}+\frac{{{\bar{\lambda }}^{\mathrm{S}\beta }}}{\mathrm{M}^{\beta }_\mathrm{mol}}\right\} }\,\nonumber \\&-\hat{\rho }^{\mathrm{F} \beta }\displaystyle {\left\{ \psi ^{\mathrm{F} \beta }-\frac{\mathrm{1}}{\mathrm{2}}\,\mathbf{x}^{\prime }_\mathrm{F \beta }\cdot \mathbf{x}^{\prime }_\mathrm{F \beta }+\frac{\bar{\lambda }^\mathrm{F\beta }}{\mathrm{M}^{\beta }_\mathrm{mol}}\right\} }\,\\&-\displaystyle {\mathbf{w}_{\mathrm{FS}}\>\cdot }\left\{ \,\displaystyle {{\hat{\mathbf{p}}}^{\mathrm{F}}-\bar{\lambda }\,\,\mathrm{grad}\>{\mathrm{n}^{\mathrm{F}}}+\bar{\lambda }^\mathrm{F\beta }\,\mathrm{c}^{\mathrm{F}\beta }\,\mathrm{grad}\>{\mathrm{n}^{\mathrm{F}}}}\,\right\} \,\nonumber \\&-\displaystyle {\mathbf{w}_{\mathrm{F}\beta \mathrm{S}}\>\cdot }\{\,\displaystyle {{\hat{\mathbf{p}}}^{\mathrm{F}\beta }-\bar{\lambda }^\mathrm{F\beta }\,\mathrm{c}^{\mathrm{F}\beta }\,\mathrm{grad}\>{\mathrm{n}^{\mathrm{F}}}\,\}}\,\ge \,0\nonumber \end{aligned}$$
(56)

must hold. Considering the aforementioned remarks, we obtain necessary and sufficient conditions for the unrestricted validity of the second law of thermodynamics. The Lagrange multipliers associated with the solutes can be identified from (55), terms 4–7, as

$$\begin{aligned} \begin{array}{ll} \bar{\lambda }^{\alpha }&{} = -\bar{\lambda }^{\alpha \beta }\,{\mathrm{c}}^{\alpha \beta },\\ \bar{\lambda }^{\alpha \beta }&{} = \displaystyle \frac{\rho ^{\alpha \beta }}{{\mathrm{n}}^{\alpha }}\, \frac{\partial \psi ^{\alpha \beta }}{\partial {\mathrm{c}}^{\alpha \beta }}= {\mathrm{c}}^{\alpha \beta }\,{\mathrm{M}}^{\beta }_\mathrm{mol}\, \frac{\partial \psi ^{\alpha \beta }}{\partial {\mathrm{c}}^{\alpha \beta }}. \end{array} \end{aligned}$$
(57)

Thus, we obtain from (55), terms 1–3 for the Cauchy stress tensors

$$\begin{aligned} \mathbf{T}^\mathbf{S}&= \mathbf{T}^{\mathrm{S}}+ \mathbf{T}^\mathrm{S\beta }\nonumber \\&= \!\displaystyle {-\bar{\lambda }\,{\mathrm{n}}^{\mathrm{S}}\,\mathbf{I}\,\underbrace{-\bar{\lambda }^\mathrm{S}\,{\mathrm{n}}^{\mathrm{S}}\,\mathbf{I}\,\!-\!{\bar{\lambda }}^{\mathrm{S}\beta }{\mathrm{c}}^{\mathrm{S}\beta }\,{\mathrm{n}}^{\mathrm{S}}\,\mathbf{I}}_\mathrm{= 0}\!+\!2\,\rho ^{\mathrm{S}}\,\mathbf{F}_{\mathrm{S}}\,{\displaystyle {\frac{\partial \psi ^{\mathrm{S}}}{\partial \mathbf{C}_{\mathrm{S}}}}}\,\mathbf{F}_{\mathrm{S}}^{\mathrm{T}}\,}\nonumber \\&= -\,\bar{\lambda }\,{\mathrm{n}}^{\mathrm{S}}\,\mathbf{I} +\underbrace{2\,\rho ^{\mathrm{S}}\,\mathbf{F}_{\mathrm{S}}\,{\displaystyle {\frac{\partial \psi ^{\mathrm{S}}}{\partial \mathbf{C}_{\mathrm{S}}}}}\,\mathbf{F}_{\mathrm{S}}^{\mathrm{T}}}_{\mathbf{T}^\mathbf{S}_\mathrm{E}} \>, \nonumber \\ \mathbf{T}^{\mathrm{F}}&= +\displaystyle {\bar{\lambda }\,{\mathrm{n}}^{\mathrm{F}}\,\mathbf{I}}+\displaystyle \bar{\lambda }^\mathrm{F}\,{\mathrm{n}}^{\mathrm{F}}\,\mathbf{I}\nonumber \\&= -\,\bar{\lambda }\,{\mathrm{n}}^{\mathrm{F}}\,\mathbf{I}+ {\mathrm{n}}^{\mathrm{F}}\, (\mathrm{c}^{\mathrm{F}\beta })^2 \mathrm M^{\beta }_\mathrm{mol}\,\displaystyle {\frac{\partial \psi ^{\mathrm{F} \beta }}{\partial \mathrm{c}^{\mathrm{F}\beta }}} \,\mathbf{I}\>,\nonumber \\ \mathbf{T}^\mathrm{F \beta }&= -\bar{\lambda }^\mathrm{F\beta }\,{\mathrm{n}}^{\mathrm{F}}\,\mathrm{c}^{\mathrm{F}\beta }\,\mathbf{I}\nonumber \\&= -\,{\mathrm{n}}^{\mathrm{F}}\,(\mathrm{c}^{\mathrm{F}\beta })^2\,\mathrm M^{\beta }_\mathrm{mol}\, \displaystyle \frac{\partial \psi ^{\mathrm{F} \beta }}{\partial \mathrm{c}^{\mathrm{F}\beta }}\,\mathbf{I}\nonumber \\&= -{\mathrm{n}}^{\mathrm{F}}\,\mu ^{\mathrm{F} \beta }\,\mathbf{I}\>, \end{aligned}$$
(58)

using (57)\(_\mathrm{1, 2}\) and the abbreviations \(\mathbf{T}^\mathbf{S} = \mathbf{T}^{\mathrm{S}}+ \mathbf{T}^\mathrm{S\beta }\) for the overall solid stress and \(\mu ^{\mathrm{F} \beta }=(\mathrm{c}^{\mathrm{F}\beta })^2\,\mathrm M^{\beta }_\mathrm{mol}\, (\partial \psi ^{\mathrm{F} \beta }/ \partial \mathrm{c}^{\mathrm{F}\beta })\) for the chemical potential. Moreover, the Lagrange multiplier \(\bar{\lambda }\) is understood as the reaction force between the solid and fluid phase, which can be identify as the pore pressure \(\lambda \) with \(\bar{\lambda }=\lambda \), cf. eg. Bluhm (2002). In view of the dissipative mechanism (56), the following approaches for the interaction forces \({\hat{\mathbf{p}}}^{\mathrm{F}}\) and \({\hat{\mathbf{p}}}^{\mathrm{F}\beta }\) and for the mass supply \(\rho ^{\alpha \beta }\)are postulated:

$$\begin{aligned} {\hat{\mathbf{p}}}^{\mathrm{F}}&= \,\bar{\lambda }\,\mathrm{grad}\>{\mathrm{n}^{\mathrm{F}}}-\bar{\lambda }^\mathrm{F\beta }\,\mathrm{c}^{\mathrm{F}\beta }\,\mathrm{grad}\>{\mathrm{n}^{\mathrm{F}}}-\alpha _{\mathbf{w}_{\mathrm{FS}}}\,\mathbf{w}_{\mathrm{FS}}\\&+\alpha _{\mathbf{w}_{\mathrm{FF}\beta }}\,\mathbf{w}_{\mathrm{FF}\beta }\>,\\ {\hat{\mathbf{p}}}^{\mathrm{F}\beta }&= \,\bar{\lambda }^\mathrm{F\beta }\, \mathrm{c}^{\mathrm{F}\beta }\,\mathrm{grad}\>{\mathrm{n}^{\mathrm{F}}}- \alpha _{\mathbf{w}_{\beta \mathrm{S}}}\,\mathbf{w}_{\beta \mathrm{S}}-\alpha _{\mathbf{w}_{\mathrm{FF}\beta }}\,\mathbf{w}_{\mathrm{FF}\beta }\>,\\ \hat{\rho }^\mathrm{FLc}&= \,-\delta ^\mathrm{FLc}(\Psi ^\mathrm{FLc}-\Psi ^\mathrm{SGy})\>,\\ \hat{\rho }^\mathrm{FGu}&= \,-\delta ^\mathrm{FGu}(\Psi ^\mathrm{FGu}-\Psi ^\mathrm{SGy})\>, \end{aligned}$$
(59)

with the abbreveation \(\Psi ^{\alpha \beta }=\psi ^{\alpha \beta }-\displaystyle \mathrm{{1}}/{2} \, {\mathbf{x}^{\prime }_{\alpha \beta }}\cdot {\mathbf{x}^{\prime }_{\alpha \beta }}+ \displaystyle {\bar{\lambda }^{\alpha \beta }}/{{\mathrm{M}}^{\beta }_\mathrm{mol}}\). The factors connected to the relative permeabilities \(\alpha _{\mathbf{w}_{\mathrm{FS}}}\), \(\alpha _{\mathbf{w}_{\mathrm{FF}\beta }}\), \(\alpha _{\mathbf{w}_{\beta \mathrm{S}}}\) and connected to the mass supply terms \(\delta ^\mathrm{FLc}\), \(\delta ^\mathrm{FGu}\) are material parameters depending on the set \(\mathcal {U}\) given in (43) and are restricted to positive values (\(\ge 0\)).

Regarding the liver lobe level (microperfusion scale), we receive constitutive restrictions with respect to the concentration exchange of lactate and glucose from (59)\(_{3,4}\) with

$$\begin{aligned} \hat{\rho }^\mathrm{FLc}&= \mathrm{M}^\mathrm{Lc}_\mathrm{mol}\,\hat{{\mathrm{c}}}^\mathrm{FLc}= -\delta ^\mathrm{FLc}(\Psi ^\mathrm{FLc}-\Psi ^\mathrm{SGy})\>,\\ \hat{\rho }^\mathrm{FGu}&= \mathrm{M}^\mathrm{Gu}_\mathrm{mol}\,\hat{{\mathrm{c}}}^\mathrm{FGu}= -\delta ^\mathrm{FGu}(\Psi ^\mathrm{FGu}-\Psi ^\mathrm{SGy})\>, \end{aligned}$$
(60)

where \(\Psi ^{\alpha \beta }=\psi ^{\alpha \beta }-\mathrm{{1}}/2 \, \mathbf{x}^{\prime }_{\alpha \beta }\cdot \mathbf{x}^{\prime }_{\alpha \beta }+ (\bar{\lambda }^{\alpha \beta }/{\mathrm{M}}^{\beta }_\mathrm{mol})\). We assume now that the concentration exchange depends neither on kinetic energy \(\mathrm 1/2 \,\, \mathbf{x}^{\prime }_{\alpha \beta }\cdot \mathbf{x}^{\prime }_{\alpha \beta }\) nor on the pressure state \(\bar{\lambda }^{\alpha \beta }\). Thus, the concentration exchange rates reduce with (18) to

$$\begin{aligned} \hat{{\mathrm{c}}}^\mathrm{FLc}&= \displaystyle \frac{\hat{\rho }^\mathrm{FLc}}{\mathrm{M}^\mathrm{Lc}_\mathrm{mol}} = \displaystyle -\frac{\delta ^\mathrm{FLc}}{\mathrm{M}^\mathrm{Lc}_\mathrm{mol}}(\psi ^\mathrm{FLc}-\psi ^\mathrm{SGy})\>,\nonumber \\ \hat{{\mathrm{c}}}^\mathrm{FGu}&= \displaystyle \frac{\hat{\rho }^\mathrm{FGu}}{\mathrm{M}^\mathrm{Gu}_\mathrm{mol}} = \displaystyle -\frac{\delta ^\mathrm{FGu}}{\mathrm{M}^\mathrm{Gu}_\mathrm{mol}}(\psi ^\mathrm{FGu}-\psi ^\mathrm{SGy})\>,\\ \hat{{\mathrm{c}}}^\mathrm{SGy}&= -\hat{{\mathrm{c}}}^\mathrm{FLc}-\hat{{\mathrm{c}}}^\mathrm{FGu}.\nonumber \end{aligned}$$
(61)

1.3 Derivation of field equations

Considering that the solid and fluid carrier phases are treated to be material incompressible (\((\rho ^{\alpha \mathrm{R}})^{\prime }_{\alpha }=0\)), see Sec. 3, leads to the simplification

$$\begin{aligned} (\rho ^{\alpha })^{\prime }_{\alpha }=({\mathrm{n}}^{\alpha }\,\rho ^{\alpha \mathrm{R}})'_S=(\mathrm{n}^{\alpha })^{\prime }_{\alpha }\,\rho ^{\alpha R} + {\mathrm{n}}^{\alpha }\,\underbrace{(\rho ^{\alpha R})}_{=0} = (\mathrm{n}^{\alpha })^{\prime }_{\alpha }\,\rho ^{\alpha R}.\nonumber \\ \end{aligned}$$
(62)

Incorporating additionally that mass exchanges are only acts between the solute (\(\hat{\rho }^{\alpha }=0\)), the balance equation of mass (14)\(_1\) reduce for the solid and fluid carrier phases to

$$\begin{aligned} (\mathrm{n}^{\mathrm{S}})^{\prime }_{\mathrm{S}}+{\mathrm{n}}^{\mathrm{S}}\>\mathrm{div}\>\mathbf{x}^{\prime }_{\mathrm{S}}=\mathrm 0\,,\quad (\mathrm{n}^{\mathrm{F}})^{\prime }_{\mathrm{F}}+{\mathrm{n}}^{\mathrm{F}}\>\mathrm{div}\>\mathbf{x}^{\prime }_{\mathrm{F}}=\mathrm 0. \end{aligned}$$
(63)

In addition, with (11) the mass for the solutes is given with

$$\begin{aligned} \begin{array}{lll} {\mathrm{m}}^{\beta }(\mathbf{x},\mathrm{t}) &{} = &{}\displaystyle \int _\mathrm{M_{{\varvec{\beta }}}}\,\mathrm d{\mathrm{m}}^{\beta }\\ &{} = &{}\displaystyle \int _{\mathrm{B}_{\varvec{\alpha }}} {\mathrm{c}}^{\alpha \beta }\,{\mathrm{M}}^{\beta }_\mathrm{mol}\,{\mathrm{d}\mathrm{v}}^{{\alpha }}\\ &{} = &{}\displaystyle \int _{\mathrm{B}_{\varvec{\alpha }}} {\mathrm{n}}^{\alpha }\,{\mathrm{c}}^{\alpha \beta }\,{\mathrm{M}}^{\beta }_\mathrm{mol}\,{\mathrm{d}\mathrm{v}}, \end{array} \end{aligned}$$
(64)

which directly leads to the solute balance equation

$$\begin{aligned} \left( {\mathrm{n}}^{\alpha }\,{\mathrm{c}}^{\alpha \beta }\,{\mathrm{M}}^{\beta }_\mathrm{mol}\right) '_{\alpha \beta }+{\mathrm{n}}^{\alpha }\,{\mathrm{c}}^{\alpha \beta }\,{\mathrm{M}}^{\beta }_\mathrm{mol}\>\mathrm{div}\>\mathbf{x}^{\prime }_{\alpha \beta }=\hat{\rho }^{\alpha \beta }. \end{aligned}$$
(65)

Further, we consider that \({\mathrm{M}}^{\beta }_\mathrm{mol}\) is a constant of the substance \(\varphi ^{\beta }\) and applies for the time derivatives for scalar fields (42). Thus, the solute balance equation of mass (65) yields with (13) to

$$\begin{aligned} (\mathrm{n}^{\alpha })^{\prime }_\mathrm{S}\,{\mathrm{c}}^{\alpha \beta }+{\mathrm{n}}^{\alpha }\,(\mathrm{c}^{\alpha \beta })^{\prime }_\mathrm{S}+ \mathrm{div}\left( \mathbf{j}^{\alpha \beta }\right) + {\mathrm{n}}^{\alpha }\,{\mathrm{c}}^{\alpha \beta }\,\mathrm{div}\>\mathbf{x}^{\prime }_{\mathrm{S}}= {\hat{\mathrm{c}}}^{\alpha \beta }.\nonumber \\ \end{aligned}$$
(66)

Herein,

$$\begin{aligned} \mathbf{j}^{\alpha \beta }\!=\! {\mathrm{c}}^{\alpha \beta }\,{\mathrm{n}}^{\alpha }\,\mathbf{w}_{\alpha \beta \mathrm{S}}\!=\! -\mathrm{D}_{\alpha \beta }\,[-{\mathrm{n}}^{\alpha }\,\mathrm{grad}{\mathrm{c}}^{\alpha \beta }]\!+\!\hat{\mathrm{D}}_{\alpha \beta }\,\mathbf{w}_{\alpha \mathrm{S}}\end{aligned}$$
(67)

denotes the molar flux, with \(\mathrm{D}_\mathrm{\alpha \beta }\) as the diffusion coefficient (Fick’s part), \(\hat{\mathrm{D}}_{\alpha \beta }\) as the advective coefficient (advective part) and \(\mathbf{w}_{\alpha \beta \mathrm{S}}=\mathbf{x}^{\prime }_{\alpha \beta }-\mathbf{x}^{\prime }_{\mathrm{S}}\) as the velocity of the solute \(\varphi ^{\alpha \beta }\) solved in the carrier phase \(\varphi ^{\alpha }\) relative to the solid velocity.

Fig. 20
figure 20

Developing different angel \(\theta _{\mathrm{ap}}^\mathrm{t}\) in the \(\mathbf a_\mathrm{0}\times \mathbf p_\mathrm{0}\) plain spanned by the pressure gradient \(\mathbf p_\mathrm{0}\) and preferred flow direction \(\mathbf a_\mathrm{0}\) or \(\mathbf a_\mathrm{0}'\)

1.4 Evolutionary approach for preferred flow direction

The stationary solution for the anisotropic flow is achieved following the phenomenological approach that sinusoids tend to be oriented in the direction of the pressure gradient \(\mathbf{\mathbf p_\mathrm{0}} =\Vert \mathrm{Grad}\,\bar{\lambda }\Vert \) with \(\mathbf{|\mathbf p_\mathrm{0}|} = \mathrm{1}\); see Fig. 20. Thus, starting from an arbitrary preferred flow direction \(\mathbf{a}_\mathrm{0}\) with \(| \mathbf{a}_\mathrm{0} | = \mathrm 1\) an updated preferred flow direction \(\mathbf{a}_\mathrm{0}'\) is calculated with the relation

$$\begin{aligned} \mathbf{a}_\mathrm{0}' = \mathbf{a}_\mathrm{0} + \triangle \mathbf{a}_\mathrm{0}' \quad \mathrm{with} \quad \mathbf{a}_\mathrm{0} \cdot \triangle \mathbf{a}_\mathrm{0} = \mathrm 0 \end{aligned}$$
(68)

where \(\triangle \mathbf{a}_\mathrm{0}'\) denotes the incremental update of the preferred flow direction \(\mathbf{a}_\mathrm{0}\)Werner et al. (2012). The incremental update can be expressed by a rigid body rotation of \(\mathbf a_\mathrm{0}\) by \(\triangle \mathbf a_\mathrm{0}= \varvec{\omega }\times \mathbf a_\mathrm{0}\, \) where \(\varvec{\omega }\) denotes the angular velocity of reorientation. Since \(\varvec{\omega }\) is perpendicular to the plane spanned by \(\mathbf a_\mathrm{0}\) and \(\mathbf p_\mathrm{0}\) the condition \({\varvec{\omega }} = \frac{\delta _\mathrm{t}}{\delta _\mathrm{d}} ( \mathbf a_\mathrm{0}\times \mathbf p_\mathrm{0})\, \) must hold with \(\delta _\mathrm{t}\) denoting a virtual damping coefficient with respect to the time-dependent model and \(\delta _\mathrm{d} = |\mathbf a_\mathrm{0}-\mathbf p_\mathrm{0}|\) weighting the distance between the preferred flow direction and the pressure gradient. Thus, the incremental update reads

$$\begin{aligned} \triangle \mathbf a_\mathrm{0}= \frac{\delta _\mathrm{t}}{\delta _\mathrm{d}} ( [\mathbf a_\mathrm{0}\times \mathbf p_\mathrm{0}] \times \mathbf a_\mathrm{0}). \end{aligned}$$
(69)

1.5 Supplemental information model reduction

\({V}_\mathrm{liver}\)

1.5 l

liver volume in full kinetic model

\({V}_\mathrm{sim}\)

 

liver volume for simulation in liter

\({V}_\mathrm{f}\)

\(\displaystyle \frac{{V}_\mathrm{sim}}{{V}_\mathrm{liver}}\)

Relative volume for adaptions of model fluxes and properties like the glycogen storage capacity to the actual liver volume

\({C}_\mathrm{HGU}, {C}_\mathrm{GLY}\)

 

Coefficients for polynomial approximation of quasi-steady state

\({m}_\mathrm{bw}\)

75 kg

Body weight used for conversion of fluxes given per body weight in concentration changes

\({k}_\mathrm{z} = {k}_\mathrm{x}\)

0.05 mM

Saturation parameter accounting for substrate limitation: HGU and glycolysis are limited by the available \(\varphi ^{\mathrm{FGu}}\). HGP and gluconeogenesis are limited by available gluconeogenic substrate (handles the border conditions for integration in the polynoms)

$$\begin{aligned} {x}&= \varphi ^{\mathrm{FGu}}\,y=\displaystyle \frac{\varphi ^{\mathrm{SGy}}}{V_f},\,z=\varphi ^{\mathrm{FLc}}\\ C_\mathrm{HGU}&= \begin{pmatrix} 0.002037960420379\\ -0.000367490977632\\ -0.069301032419012\\ -0.000002823120484\\ 0.011282864074433\\ 3.740276159346358\\ -0.000000181515364\\ 0.000157328485157\\ -0.100050917438436\\ -18.414978834613287\\ \end{pmatrix}\\ {C}_\mathrm{GLY}&= \begin{pmatrix} -0.013260401508103\\ -0.000078240970095\\ 0.478235644004833\\ 0.000002861605817\\ 0.000932752106971\\ -2.492569641130055\\ 0.000000166945924\\ -0.000125285017396\\ 0.015354944655784\\ -4.975026288067225\\ \end{pmatrix} \\ \end{aligned}$$
$$\begin{aligned}&{p}_{\mathrm{HGU}}(\varphi ^{\mathrm{FGu}},\varphi ^{\mathrm{SGy}},\varphi ^{\mathrm{FLc}})\nonumber \\&\quad =\displaystyle \frac{{V}_{\mathrm{f}}\,{{m}}_{\mathrm{bw}}}{60\cdot 10^{-3}}\cdot ({C}_{\mathrm{HGU}}[1]\cdot {{x}}^3+{{C}}_{\mathrm{HGU}}[2]\cdot {x}^2\mathrm{y}\nonumber \\&\quad +\,{C}_{\mathrm{HGU}}[3]\cdot {x}^2 +{C}_{\mathrm{HGU}}[4]\cdot {xy}^2+{C}_{\mathrm{HGU}}[5]\cdot {xy}\nonumber \\&\quad +\,{C}_{\mathrm{HGU}}[6]\cdot {x}+{{C}}_{\mathrm{HGU}}[7]\cdot {y}^3 +{C}_{\mathrm{HGU}}[8]\cdot {y}^2\nonumber \\&\quad +\,{C}_{\mathrm{HGU}}[9]\cdot {y} +{C}_{\mathrm{HGU}}[10])\end{aligned}$$
(70)
$$\begin{aligned}&{v}_{\mathrm{HGU}}(\varphi ^{\mathrm{FGu}},\varphi ^{\mathrm{SGy}},\varphi ^{\mathrm{FLc}})\nonumber \\&\quad ={p}_{\mathrm{HGU}}(\varphi ^{\mathrm{FGu}},\varphi ^{\mathrm{SGy}},\varphi ^{\mathrm{FLc}})\cdot {\left\{ \begin{array}{ll} \displaystyle \frac{{z}}{{z}+{k}_{\mathrm{z}}} \quad \forall \,{p}_{\mathrm{HGU}}<0\\ \displaystyle \frac{{x}}{{ x+k}_{\mathrm{x}}} \quad \forall \,{p}_{\mathrm{HGU}}\ge 0\\ \end{array}\right. }\end{aligned}$$
(71)
$$\begin{aligned}&{p}_{\mathrm{GLY}}(\varphi ^{\mathrm{FGu}},\varphi ^{\mathrm{SGy}},\varphi ^{\mathrm{FLc}})\nonumber \\&\quad =\displaystyle \frac{{V}_\mathrm{f}\,\mathrm{m}_\mathrm{bw}}{60\cdot 10^{-3}}\cdot ({C}_\mathrm{GLY}[1]\cdot \mathrm{x}^3+\mathrm{C}_\mathrm{GLY}[2]\cdot { x}^{2}{ y+C}_\mathrm{GLY}[3]\cdot {x}^2\nonumber \\&\quad +\,{ C}_\mathrm{GLY}[4]\cdot {xy}^{2}+{C}_\mathrm{GLY}[5]\cdot {xy+C}_\mathrm{GLY}[6]\cdot {x+C}_\mathrm{GLY}[7]\cdot { y}^3\nonumber \\&\quad +\,{C}_\mathrm{GLY}[8]\cdot {y}^{2}+{C}_\mathrm{GLY}[9]\cdot {y+C}_\mathrm{GLY}[10])\end{aligned}$$
(72)
$$\begin{aligned}&{v}_\mathrm{GLY}(\varphi ^{\mathrm{FGu}},\varphi ^{\mathrm{SGy}},\varphi ^{\mathrm{FLc}})\nonumber \\&\quad ={p}_\mathrm{GLY}(\varphi ^{\mathrm{FGu}},\varphi ^{\mathrm{SGy}},\varphi ^{\mathrm{FLc}})\cdot {\left\{ \begin{array}{ll} \displaystyle \frac{{z}}{{z+k}_\mathrm{z}} \quad \forall \,{p}_\mathrm{GLY}<0\\ \displaystyle \frac{{x}}{{x+k}_\mathrm{x}} \quad \forall \,{p}_\mathrm{GLY}\ge 0\\ \end{array}\right. }\end{aligned}$$
(73)
$$\begin{aligned}&{ v}_\mathrm{GS}(\varphi ^{\mathrm{FGu}},\varphi ^{\mathrm{SGy}},\varphi ^{\mathrm{FLc}})\nonumber \\&\quad ={v}_\mathrm{HGU}(\varphi ^{\mathrm{FGu}},\varphi ^{\mathrm{SGy}},\varphi ^{\mathrm{FLc}})-{\mathrm{v}}_\mathrm{GLY}(\varphi ^{\mathrm{FGu}},\varphi ^{\mathrm{SGy}},\varphi ^{\mathrm{FLc}})\end{aligned}$$
(74)
$$\begin{aligned}&\displaystyle \frac{\mathrm{d}\,\varphi ^{\mathrm{FGu}}}{\mathrm{dt}}=\mathrm{-v}_{\mathrm{HGU}}(\varphi ^{\mathrm{FGu}}, \varphi ^{\mathrm{SGy}}, \varphi ^{\mathrm{FLc}})= \hat{{\mathrm{c}}}^\mathrm{FGu}, \nonumber \\&\displaystyle \frac{\mathrm{d}\,\varphi ^{\mathrm{SGy}}}{\mathrm{dt}}=\mathrm{-v}_{\mathrm{HGU}}(\varphi ^{\mathrm{FGu}}, \varphi ^{\mathrm{SGy}}, \varphi ^{\mathrm{FLc}})= \hat{{\mathrm{c}}}^\mathrm{FGu}, \\&\displaystyle \frac{\mathrm{d}\,\varphi ^{\mathrm{FLc}}}{\mathrm{dt}}= 2\cdot \mathrm{v}_{\mathrm{GLY}}(\varphi ^{\mathrm{FGu}}, \varphi ^{\mathrm{SGy}}, \varphi ^{\mathrm{FLc}})= \hat{{\mathrm{c}}}^\mathrm{FLc}.\nonumber \end{aligned}$$
(75)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ricken, T., Werner, D., Holzhütter, H.G. et al. Modeling function–perfusion behavior in liver lobules including tissue, blood, glucose, lactate and glycogen by use of a coupled two-scale PDE–ODE approach. Biomech Model Mechanobiol 14, 515–536 (2015). https://doi.org/10.1007/s10237-014-0619-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10237-014-0619-z

Keywords

Navigation