Survey
* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project
* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project
REVIEWS PHYSIOLOGY 20: 316–325, 2005; 10.1152/physiol.00022.2005 A Strategy for Integrative Computational Physiology Peter Hunter and Poul Nielsen Bioengineering Institute, University of Auckland, Auckland, New Zealand [email protected] Organ function (the heart beat for example) can only be understood through knowledge of molecular and cellular processes within the constraints of structurefunction relations at the tissue level. A quantitative modeling framework that can deal with these multiscale issues is described here under the banner of the International Union of Physiological Sciences Physiome Project. The wealth of data now available from 50 years of research in molecular and cellular biology, coupled with recent developments in computational modeling, presents a wonderful opportunity to understand the physiology of the human body across the relevant 109 (from nanometers to meters) range of spatial scales and 1015 (from microseconds to a human lifetime) range of temporal scales. The challenge is to develop mathematical models of structure-function relations appropriate to each (limited) spatial and temporal domain and then to link the parameters of a model at one scale to a more detailed description of structure and function at the level below. To span from molecules to organ systems requires databases of models at many spatial and temporal levels and software tools for authoring, visualizing, and running models based on widely adopted modeling standards. It also requires the development of ontologies dealing with anatomy, physiology, and molecular and cellular biology to uniquely identify and link model components. In this review, we outline a framework for computational physiology being developed under the auspices of the Physiome and Bioengineering Committee (co-chaired by P. Hunter and A. Popel) of the International Union of Physiological Sciences (IUPS). We give examples of its application to several organs and organ systems, particularly the heart, and highlight the challenges for this so-called Physiome Project (1, 3, 10, 11, 19, 20). Note that the focus on computational physiology is the distinguishing feature of the Physiome Project but that it also encompasses the area traditionally known as systems biology, which has arisen out of molecular and cellular biology and which typically deals with subcellular signaling cascades, metabolic pathways, and gene-regulation networks. Lessons from Physics and Engineering: Continuum Fields and Constitutive Laws One of the most important concepts of mathemat316 ical analysis in the physical and engineering sciences is that of a “field,” which, together with appropriate mathematical operations on fields, is used to express the physical conservation laws of nature, such as conservation of mass, momentum, charge, etc. Fields were first introduced by the great 19th-century British scientist Michael Faraday to represent the intensity and direction of the force experienced by a small charged particle in the presence of magnetic dipoles and electric charges throughout a three-dimensional space. The greatest success of 19th- and early 20th-century physics was the discovery that most observable phenomena could be described by differential equations operating on these fields: Maxwell’s field equations for electricity and magnetism, Einstein’s field equation for gravity, the Yang-Mills field equations for subatomic particles, and the field equations of continuum mechanics were derived from Newton’s laws of conservation of mass and conservation of momentum. There is an interesting analogy with molecular biology here: just as the classical field theory world of physics was overshadowed in the early 20th century by quantum theory and particle physics, the integrative systems-level physiology of the first half of the 20th century gave way to the reductionist molecular biology of the second half. In both cases the challenge now is to reconcile the continuum/systems view with the particle/molecular view. For physicists, this may be achievable by describing their equations in a higher-dimensional space via superstring theory. For mathematical biologists, it is likely to be achieved via multiscale modeling. Many of the great accomplishments of electrical and mechanical engineering, respectively, in the 20th century were built (for the former), on the sound theoretical basis of Maxwell’s field theory and quantum mechanics and (for the latter) on continuum mechanics. In both cases, however, the theoretical field theory models had to be supplemented with empirical descriptions of material behavior that hide molecular or atomic detail behind approximate “constitutive laws.” For elec- 1548-9213/05 8.00 ©2005 Int. Union Physiol. Sci./Am. Physiol. Soc. Downloaded from http://physiologyonline.physiology.org/ by 10.220.33.6 on June 15, 2017 The Need for a Multiscale Modeling Framework REVIEWS Human Genome Project recognized, the outcome will be much more than the sum of the parts: an anatomically and biophysically based model of the human body will reveal characteristics and interactions that could not be predicted by focusing on only a single spatial scale, time scale, or region of the body. Organ-Level Modeling Computational modeling based on the above physical principles must be applied at many levels (see Table 1): at the scale of whole organs, in which the overall geometry and structure is modeled and used to solve the coupled systems of field equations (see FIGURE 1); at the tissue level, typically dealing with millimeter-sized blocks of tissue in which structures are defined at the micrometer level; at the cell level, typically dealing with a resolution of 0.1 m; and even at the protein level. Good progress is being made on modeling the anatomy and biophysics of the heart (FIGURE 1A), the lungs (FIGURE 1B), the digestive system (FIGURE 1C), and the musculoskeletal system (FIGURE 1D). In each case the field equations are a little different (the lungs, for example, do not require the electrical field equations), and, despite very different anatomy and material properties (expressed via the constitutive laws), the numerical modeling framework required to solve these field equations is the same for all. At the current rate of progress, models of all 12 organ systems should be available within a few years, at least for physiologically normal human anatomy. Linking the organ and organ systems together to yield models that can predict and interpret multiorgan physiological behavior is the focus of systems physiology. To Downloaded from http://physiologyonline.physiology.org/ by 10.220.33.6 on June 15, 2017 trical engineering, for example, Ohm’s law gives a linear relationship between voltage gradient and current flow (analogous to Fick’s law giving solute flux from concentration gradient via a coefficient of diffusion). The proportionality constant (electrical “resistance”) can be measured directly and used in the solution of the governing Maxwell’s field equation (which reduces to Kirchoff’s current law in this simple example) without dealing with the underlying structure of the material. An implementation of multiscale modeling would be to derive the constitutive relation (Ohm’s law) from field equations applied to the detailed material structure. For mechanical engineering, the equivalent is the empirically derived relationship between stress and strain (or strain rate for a fluid), although in this case these two measures of force and deformation, respectively, require the use of tensors, each with six components (three components represent forces or deformations acting normally to the three orthogonal faces of a cube, and three represent the shearing forces or deformations). The development of stress and strain tensors as the appropriate quantities to represent the mechanical state of the material at a point was a major intellectual achievement that, together with an understanding of how to simplify these tensors and their relationships for various solids (beams, plates, shells, etc.) and liquids (e.g., Newtonian fluids) led to an ability to design complex engineering structures. The application of continuum field concepts and constitutive laws, whose parameters are derived from separate, finer-scale models, is the key to linking molecular systems biology (with its characterization of molecular processes and pathways) to larger-scale systems physiology (with its characterization of the integrated function of the body’s organ systems). The appropriate name for this application of physical and engineering principles to physiology is computational physiology. The term systems biology, currently inappropriately limited to the molecular scale, needs to be associated with all spatial scales. Two principles guiding the development of all mathematical models are 1) “Occam’s razor”—the simpler the better, consistent with the level of experimental data available (which, as Einstein noted, also implies not too simple!), and 2) that every model must be validated against experimental measurement. This journal’s space limitations mean that both of these issues are delegated to the references for the models we discuss. Similarly, the physiological insights achieved by these models are left to the references, because the focus here is on presenting a strategy for integrative multiscale modeling. There is also a third guiding principle for Physiome Project models: just as the leaders of the “The appropriate name for this application of physical and engineering principles to physiology is computational physiology.” examine the behavior of these systems over long time intervals, the models are often lumpedparameter models in which system-identification techniques are applied to the detailed field-equation models illustrated in FIGURE 1. This approach allows the parameters of the lumped-parameter models to be linked to the more biophysically based organ models (see the cardiac example below). The organ-level models shown in FIGURE 1 are based on finite-element models of the anatomic PHYSIOLOGY • Volume 20 • October 2005 • www.physiologyonline.org 317 REVIEWS Table 1. The multiscale modeling hierarchy Physiological Level Organ systems Cardiovascular, respiratory, musculoskeletal, digestive, skin, urinary, nervous, endocrine, lymphatic, male reproductive, female reproductive, special sense organs Types of Model, etc. Imaging Technologies Systems theory Lumped-parameter models Software JSIM MRI, CT, PET 3-D Ultrasound 1 mm or 10–3 m BioPSE Organs 226 bones 850 muscles 2817 arteries, arterioles, and capillary beds … Continuity Tissues Epithelial Connective Muscle Nerve Cells Approximately 200 cell types Organelles Cell membrane, mitochondria, nucleus, endoplasmic reticulum, ribosomes, Golgi apparatus, centrioles, lysosomes, peroxisomes Cell function Membrane receptors Membrane ion channels Signaling pathways Metabolic pathways Transport Motility Maintenance Conservation equations Passive flux equations Carrier-mediated transport Electroneutrality constraints Continuum models Lattice-Boltzmann equations Stochastic models Markov models Hodgkin-Huxley models Network models Numerical methods Finite-element method Boundary-element method Finite-difference method MicroCT, Optical coherence tomography 10 m or 10–5 m Serial sections 1 m or 10–6 m Confocal microscopy 1 m or 10–6 m 2-photon microscopy 100 nm or 10–7 m Electron tomography 1 nm or 10–9 m Cell cycle Proteins, carbohydrates, and lipids Quantum mechanics Molecular dynamics Poisson-Boltzmann eqs. Coarse-grained models Software AMBER X-ray diffraction, NMR 1 Å or 10–10 m CHARMM Posttranslational modifications Protein folding Translation Posttranscriptional modifications Gene regulation Transcription Genes 20,000 genes Stochastic models Boolean network models Microarrays 2-D gel electrophoresis Mass spectrometry Nucleic acid sequencing URLs for the software packages mentioned above are as follows: JSim, http://nsr.bioeng.washington.edu/PLN/Software; BioPSE, http://software.sci.utah.edu; CMISS, http://www.cmiss.org; Continuity, http://www.continuity.ucsd.edu; Virtual Cell, http://www.ibiblio.org/virtualcell; Gepasi, http://www.gepasi.org; AMBER, http://amber.scripps.edu; CHARMM, http://www.charmm.org. 318 PHYSIOLOGY • Volume 20 • October 2005 • www.physiologyonline.org Downloaded from http://physiologyonline.physiology.org/ by 10.220.33.6 on June 15, 2017 Continuum theory Conservation of mass Conservation of momentum Conservation of charge Software CMISS REVIEWS A D B FIGURE 1. Organ-level models A: geometry and microstructure of ventricular myocardium shown in a 3-D finite-element model of the heart. Left: finite-element surfaces fitted to measurements from the left and right ventricles of the pig heart. Middle: 2 layers of streamlines (one on the epicardial surface and one midway through the wall) are used to visualize the epicardial and midwall fiber directions. Right: coronary arteries modeled from pig heart data (52). B: models of the lung (6, 54). Left: airways. Right: pulmonary arteries. C: models of the digestive system, including the esophagus, stomach, intestines, and colon (4). Middle: diaphragm. Right: stomach and large intestines. D: models of the musculoskeletal system (14). Left: human skeleton with all bones modeled and also showing muscles in the left leg as well as heart, lungs, and digestive system. Right: muscles of the left arm. fields (geometry and tissue structure) encoded in a markup language called FieldML (http:// www.physiome.org.nz/fieldml/pages). In all of these models, the anatomic description is designed to facilitate the efficient solution of the field equations representing organ function. The dependent variables used in these field equations (such as the deformed geometry, membrane potential, oxygen concentration, etc.) can also be exported from computational codes in FieldML-formatted files. The use of FieldML as a standard for encoding field information allows the models and their solution fields to be read by a range of simulation and visualization codes (see Table 1). The spatially varying physical properties of the tissues are also described with FieldML files. Note that an important difference between engineering materials and biological tissues is that the latter are almost always anisotropic (e.g., material properties are different in orthogonal material directions), inhomogeneous (properties vary with position in the materi- al), and nonlinear. This makes the characterization of biological tissue properties much more difficult than it is for engineering materials and also reinforces the need to derive the material parameters of the constitutive laws used in organ models from more detailed models of the underlying tissue structure (13). Cell-Level Models: CellML, Associated Tools, and Databases A framework for modeling cell function has been developed over the past five years by the Bioengineering Institute at the University of Auckland. The key elements are: 1) the development of a markup language called CellML to standardize the mathematical description and metadata (information about the model) of cell models (12, 15, 30); 2) application programming interfaces (APIs) to define the way that information is passed between the CellML model files and computer pro- PHYSIOLOGY • Volume 20 • October 2005 • www.physiologyonline.org 319 Downloaded from http://physiologyonline.physiology.org/ by 10.220.33.6 on June 15, 2017 C REVIEWS at http://www.cellml.org/examples/repository includes ~300 models in the following categories: Signal transduction pathway models Metabolic pathway models Cardiac electrophysiological models Calcium dynamics models Immunology models Cell cycle models Simplified electrophysiological models Other cell type electrophysiological models Smooth and skeletal muscle models Mechanical models and constitutive laws For each model, the information on the website describes which biological processes are represented, references the associated publication, and links to the CellML model file itself. Although the use of CellML can never replace the need for a peerreviewed publication of a model, it can enhance the availability of the model and ease the burden of implementation validity checking by the model users. A model-publishing workflow is being developed to allow reviewers of submitted papers to access and test the models and also to allow updating of models if errors or omissions are found after publication. Note that another XML-based standard called systems biology markup language (SBML; see http://www.sbml.org) has been developed for describing models of biochemical reaction networks (18). Multiscale User Interfaces and Ontologies To facilitate access to the anatomically based FIGURE 2. MozCellML simulation tool for models from the CellML database The CellML file, in this case describing the cardiac cell ion channels that generate the membrane action potential, is read and transformed into code, which is complied and linked on the fly. The resulting executable program integrates the differential equations to yield the time course of the (typically 10–50) dependent variables in the model. The computed action potential is seen in the graphical output window. Another set of windows (not shown here) is designed for creating CellML models. Executable and source code versions of MozCellML are available for several different operating systems from http://cellml.sourceforge.net. 320 PHYSIOLOGY • Volume 20 • October 2005 • www.physiologyonline.org Downloaded from http://physiologyonline.physiology.org/ by 10.220.33.6 on June 15, 2017 grams written in languages such as C, C++, Fortran, Java, Perl, Python, or Matlab (see http:// www.cellml.org/private/api); 3) the development of authoring tools to facilitate the creation of the computational model from the mathematical description; 4) the creation of a simulation code that reads the chosen CellML files, generates the source code, and compiles and links this into an executable that can then be run for a specified time interval to generate model results (see FIGURE 2); and 5) the development, in collaboration with other groups, of ontologies to represent human anatomy and physiological processes (9, 41). Note that the CellML language is designed to represent mathematical models of any biological system and achieves the following goals: it is machine readable because it is based on extensible markup language (XML); it represents concepts from mathematical modeling (model, variables, units, and mathematics); and it uses structures in common with objectoriented programming (abstraction, interfaces, encapsulation, containment, and importing). Where appropriate, the language builds upon existing XML standards, such as MathML (http://www.w3.org/Math) for the specification of mathematical equations and the resource description framework (http://www.w3.org/RDF) for the encapsulation of metadata. Other tools have been written to check, for example, the consistency of units in the CellML model descriptions. All of these software tools are freely available, for both academic and commercial use, at http://cellml.sourceforge.net. The goal is to make all published models of cellular function available in CellML format and to encourage the authors of these models (or others) to ensure that the model parameters and initial conditions available on the CellML site are consistent with the model results published in the paper. The repository of models currently available REVIEWS FIGURE 3. A generic framework for generating graphical user interfaces for FieldML and CellML files The example here is a user interface designed to teach the principles of EKG to medical students. The window on the left includes 3-D anatomically based models of the heart and torso that can be rotated and zoomed. EKG leads are shown at the standard positions on the torso, but these can be moved. An activation sequence is shown on the heart (hidden in this view) and torso. Moving current dipole vectors are also computed and displayed. Potential maps are shown on the torso surface as a color map, and the time course of EKG leads is shown on the right. Image courtesy of Carey Stevens (Bioengineering Institute, Auckland, New Zealand). An Example: Computational Physiology of the Heart The development of the Auckland-Oxford-UCSD Downloaded from http://physiologyonline.physiology.org/ by 10.220.33.6 on June 15, 2017 organ-level models and their links downward to tissue properties and cell-level processes, we are developing browser-based graphical user interfaces that exploit the open-source software framework provided by the Mozilla community (http://www.mozilla.org). These use an XML-based description of the user interfaces, called XUL, which allows problem-specific interfaces to be easily created, as illustrated in FIGURE 3. Providing a set of unique names for the components of a biological model (a “controlled vocabulary”), together with the relationships between the components, allows the application of reasoning algorithms and access to the growing number of protein databases that give, for example, data on protein-protein interactions. We are working with a number of groups who are developing domainspecific ontologies such as the Gene Ontology project (http://www.geneontology.org) and the Foundational Model of Anatomy (http://sig. biostr.washington.edu/projects/fm/AboutFM.htm l) (41). A web interface based on these ontologies is being developed for accessing models (http:// n2.bioeng5.bioeng. auckland.ac.nz/ontology). The specific biological information that is relevant to each model in the CellML model repository is represented in the metadata associated with a model and is bound directly to parts of the ontology. The CellML metadata editor will be extended to use the ontologies, supplying modelers with a system of biological entities and relations that they can associate with their modeling structures. heart model is used here as an example of multiscale physiome modeling. The long-term goal is to be able to predict the effect of a channel mutation or drug-binding event on arrhythmogenesis at the whole-heart level, together with its clinically observable effect on the body surface. Although there are still many gaps, this example will serve to illustrate what has been achieved with the physiome approach and where further research is needed. All models described here are encoded in CellML and FieldML and are available from the IUPS Physiome Project databases. The various levels referred to here are illustrated in FIGURE 4. Atomistic models Molecular dynamics (MD) models of the atomic structure of ion channels, pumps, exchangers, etc. are needed that can predict the open-channel permeation of the channels, the voltage dependence of the channel permeability, and the time- and voltage-dependent gating behavior. These models describe the atomic mass and bonded (covalent) and nonbonded (electrostatic, van der Waals) forces operating on all (or a sizable subset of) atoms in the protein and then solve Newton’s laws of motion. The covalent forces can be computed as a function of bond length, bond angle, and dihedral angle from quantum-mechanical calculations. Water is usually included in discrete molecular form or, for greater distances, as a bipolar continuum field. These models require the protein’s atomic structure to be known from X-ray diffraction or NMR, and this has proved very difficult for membrane-bound proteins. Ideally of course this structure would come from protein-folding computations, but this is some way off. Structures are currently known for a few bacterial potassium chan- PHYSIOLOGY • Volume 20 • October 2005 • www.physiologyonline.org 321 REVIEWS 1 Atomic 3 Subcellular pathways IKATP IK1 K+ 2 IKto IKr1 IKr2 K+ K+ K+ K+ IKs INa Calcium buffering TropC Calmodulin Ca2+ Ca2+ Vleak K+ Magnesium buffering MgATP MgADP Mg2+ MgPi 3MgADP + 3Pi H2+ Protein TT Vtr 2+ Ca Vrel Glycogenesis 3MgATP 2Lac IbNa IpNa Na+ Na+ Na+ JSR Ca2+ IbCa Vup NSR MgATP+ 2K+ Ca2+ MgATP 3H+ Ca2+ Contraction ADP MgADP ImKATP H2PO4– H+ AK K+ MgADP + Pi + H OxPhos AMP PCr Ca2+ MgATP Lac 4 3-D cell 5 Na+ H+ LAT Tissue ATP Na+ HCO3– H+ H2PO4– NPE NHE 6 Intrinsic buffers CK Cr 2Na+ MgATP ADP + H+ VCO2transport NBC Whole heart Cl– Cl– OH– CHE 3Na+ HCO3– CBE 7 Ca2+ NCE Torso FIGURE 4. An illustration of the hierarchy of spatial scales used in the IUPS Heart Physiome Project The levels are 1) atomic, shown here by the atomic coordinates for the sarco(endo)plasmic reticulum calcium ATPase (SERCA) from the PDB database (note the 2 bound calcium atoms in white); 2) protein A, demonstrating a coarser-grained model of structure for the same protein; 3) subcellular pathways, which include the cell electrophysiology, calcium transport, proton transport, myofilament mechanics, metabolic pathways, and some cell-signaling pathways; 4) 3-D cell, in this case cardiac muscle cell from electron micrograph; 5) tissue organization, shown here with collagen fibers in a transmural tissue block from a rat heart; 6) whole heart (The Auckland Textured Virtual Heart); and 7) torso (the Auckland Human Torso Model). nels such as KcsA, KirBac, and KvAP, which have close homology (at least in the pore region) to mammalian potassium channels (43). MD calculations, based on ~100,000 atoms in current models, are very expensive and are typically run for periods of only 10 ns (7). Sometimes homology modeling is used in combination with MD simulation to generate, test, and refine models of mammalian potassium channels based on bacterial templates (8). The structures of sodium and calcium channels are also on the horizon, as well as those of key pumps and exchangers. Protein models A major challenge now is to develop coarse-grained models of these ion channels and other proteins with parameters calculated from the MD models. This will allow the models to include transient gating behavior for time intervals up to ~100 ms. 322 PHYSIOLOGY • Volume 20 • October 2005 • www.physiologyonline.org Models of channel kinetics based on whole-cell voltage-clamp data—the now classic HodgkinHuxley equations (16)—were developed over 50 years ago for sodium, potassium, and chloride channels in giant squid axons and over 40 years ago by Denis Noble for cardiac ion channels (36). The subsequent development of the single-channel patch clamp led to refinement of these models with Markov-state variable models (53). One of the challenges now for the Heart Physiome Project is to derive the parameters of the Hodgkin-Huxley or Markov models from the MD models via coarsegrained intermediate models as the molecular structures of these proteins become available. Subcellular pathways Lumped-parameter models (i.e., without diffusion modeling) of various processes within cardiac myocytes are now well advanced. The composite Downloaded from http://physiologyonline.physiology.org/ by 10.220.33.6 on June 15, 2017 + CO2 3Na+ REVIEWS cell models encompass all of the ion channels, ATPase pumps, and exchangers known to support the cardiac action potential (34), together with equations governing calcium transport (47), proton exchange (55, 56), myofilament mechanics (21), metabolic pathways (2), and signal-transduction pathways that modulate the phosphorylation state of intracellular proteins (44, 45, 46). All of these models are available in the CellML repository. 3-D cell models Tissue models Cardiac cells (myocytes and fibroblasts) are organized into layers three or four cells thick (27, 28, 29) that ensure high-conductivity gap-junction-mediated connections within sheets to ensure rapid activation of the myocardium but allow mechanical shearing to occur between sheets to allow the wall thickening that ensures a high ejection fraction for the ventricles. This tissue structure has been modeled mechanically and electrically to bridge the gap between cell-level and tissue-level models (23, 31, 33). For example, computing the propagation of an activation wavefront through a tissue block with a detailed model of tissue structure allows the continuum orthotropic conductivity parameters, which are then used in the wholeheart models, to be evaluated in different regions of the myocardium (17). Whole-heart models Models at the whole-heart level can now deal with the large deformation mechanics of the ventricles by using orthotropic constitutive properties based on the fiber-sheet orientations (29, 35) and with the activation wavefront propagation computed from bidomain field equations (23) by using a conductivity tensor based on the fiber-sheet orientations and the ion-channel cell models referred to above. The coupling between the mechanical and electrical models is also achieved with calcium released from ryanodine receptors binding to troponin-C to initiate cross-bridge interaction (21) and mechanoelectric feedback mechanisms such as stretch-activated channels (25). A model of the coronary tree Torso models A model of the torso has been developed that includes the heart and lungs within the thoracic cavity and the layers of skin, fat, and muscles on the chest to compute the current flow out of the heart that gives rise to the measured EKG or more detailed body surface maps of skin potentials (5). One goal of this work is to solve the inverse problem of electrocardiology: how to compute activation patterns in the heart from measured potential maps on the body surface (40). Downloaded from http://physiologyonline.physiology.org/ by 10.220.33.6 on June 15, 2017 The next stage of development of cell models will need to take account of the spatial distribution of proteins within a cell and subcellular compartments, where second messengers (Ca2+, IP3, cAMP, etc.) are localized (47). The spatial distribution of the transverse-tubular system in cardiac myocytes has been measured with two-photon microscopy (51), and the spatial modeling of proton transport and buffering in the myocyte is also well advanced (55). Developing 3-D models at the cellular level will help to fill the large gap in spatial scales between proteins and intact cells. has also been developed (49, 50) and coupled with the ventricular mechanics model to examine regional energy demand/supply relations (48). Current work is linking myocardial mechanics to the fluid mechanics of blood flow in the ventricles and to the function of the heart valves. Future work will need to include models of the Purkinje network and the autonomic nervous system (39). Summary, Goals, and Future Directions Anatomically and biophysically based models of 4 of the 12 organ systems in the human body are now quite well developed at the organ and tissue levels (the cardiovascular, respiratory, digestive, and musculoskeletal systems). Others (the lymphatic system, the kidney and urinary system, the skin, the female reproductive system, and the special sense organs) are at an early stage of development, and the remainder (the endocrine, male reproductive, and brain and nervous systems) will be addressed over the next few years. Software to visualize anatomically based models and solve the coupled biophysical field equations on these models is available (see Table 1). All of the models can be readily adapted to an individual patient if appropriate clinical imaging data are available (14, 57). Many of the tools required to author, visualize, and run models at the cell level are either available now or soon will be (22, 24, 26, 32). In the near future these tools will become more integrated and more easily able to link through ontologies to bioinformatic databases (19, 20, 37). This will facilitate international collaboration between groups working in complementary areas—with the multiscale models acting as the glue to connect and integrate information (38, 42). Much of the framework for dealing with cell function is in place, and a substantial repository from published models has been created (http://www.cellml.org). An important future development of the CellML framework is to use the importing feature to assemble models from their individually described components. For example, a cardiac cell model may contain mathematical PHYSIOLOGY • Volume 20 • October 2005 • www.physiologyonline.org 323 REVIEWS We are grateful to many of our colleagues in the Bioengineering Institute at the University of Auckland, whose research work we have drawn on, and in particular to Shane Blackett, Matt Halstead, Catherine Lloyd, Andrew Miller, David Nickerson, Greg Sands, and Carey Stevens for contributions to the cmgui graphics library, ontologies, CellML model database, MozCellML code, CellML APIs, tissue image processing, and Mozilla/XUL user interfaces, respectively. We also gratefully acknowledge Professors Denis Noble, David Paterson, and Peter Kohl at Oxford University, and Professor Andrew McCulloch at the University of California, San Diego, all of whom are contributing to the Wellcome Trust Heart Physiome Project. Some of the work reported here has been funded by the New Zealand Government Tertiary Education Commission (TEC) through a Centre of Research Excellence grant to the Centre of Molecular Biodiscovery; the Wellcome Trust, through a grant on the Heart Physiome Project; the Health Research Council of New Zealand and the Royal Society (New Zealand) Marsden fund. References 1. Bassingthwaighte JB. Strategies for the physiome project. Ann Biomed Eng 28: 1043–1058, 2000. 2. Bassingthwaighte JB and Vinnakota KC. The computational integrated myocyte: a view into the heart. Ann NY Acad Sci 1015: 391–404, 2004. 3. Bock G and Goode JA. Integrative biological modeling in silico. In: ‘In Silico’ Simulation of Biological Processes, edited by Bock G and Goode JA. New York: Wiley, 2002. 4. Buist M, Cheng L, Yassi R, Bradshaw L, Richards W, and Pullan A. An anatomical model of the gastric system for producing bioelectric and biomagnetic fields. Physiol Meas 25: 849–861, 2004. 5. Buist M and Pullan AJ. Torso coupling techniques for the forward problem of electrocardiology. Ann Biomed Eng 30: 1299–1312, 2002. 6. Burrowes KS, Tawhai MH, and Hunter PJ. Modeling RBC and neutrophil distribution through an anatomically based pulmonary capillary network. Ann Biomed Eng 32: 585–595, 2004. 7. Capener CE and Sansom MS. MD simulations of a K channel model—sensitivity to changes in ions, waters and membrane environment. J Phys Chem B Condens Matter Mater Surf Interfaces Biophys 106: 4543–4551, 2002. 8. Capener CE, Shrivastava IH, Ranatunga ML, Forrest R, Smith GR, and Sansom MS. Homology modelling and molecular dynamics simulation studies of an inward rectifier potassium channel. Biophys J 78: 2929–2942, 2000. 9. Cook DL, Rosse C, and Mejino JVL. Evolution of a foundational model of physiology: symbolic representation for functional bioinformatics. In: MEDINFO 2004. Amsterdam: IOS, 2004, p. 336–340. 10. Crampin EJ, Halstead M, Hunter P, Nielsen P, Noble D, Smith N, and Tawhai M. Computational physiology and the Physiome Project. Exp Physiol 89: 1–26, 2004. 11. Crampin E, Smith N, and Hunter P. Multi-scale modelling and the IUPS physiome project. J Mol Histol 35: 707–714, 2004. 12. Cuellar AA, Lloyd CM, Nielsen PM, Bullivant DP, Nickerson DP, and Hunter PJ. An overview of CellML 1.1, a biological model description language. Simulation 79: 740–747, 2003. 13. Dokos S, Smaill BH, Young AA, and LeGrice IJ. Shear properties of passive ventricular myocardium. Am J Physiol Heart Circ Physiol 283: H2650—H2659, 2002. 14. Fernandez J, Mithraratne P, Thrupp S, Tawhai MH, and Hunter PJ. Anatomically based geometric modelling of the musculoskeletal system and other organs. Biomech Model Mechanobiol 2: 139–155, 2004. 15. Hedley W, Nelson MR, Bullivant D, and Nielsen P. A short introduction to CellML. Philos Trans R Soc Lond A Math Phys Sci 359: 1073–1089, 2001. 16. Hodgkin AL and Huxley AF. A quantitative description of membrane current and its application to conductance and excitation in nerve. J Physiol 117: 500–544, 1952. 17. Hooks DA, Tomlinson KA, Marsden SG, LeGrice IJ, Smaill BH, Pullan AJ, and Hunter PJ. Cardiac microstructure: implications for electrical propagation and defibrillation in the heart. Circ Res 91: 331–338, 2002. 18. Hucka M, Bolouri H, Finney A, Sauro HM, Doyle JC, Kitano H, Arkin AP, Bornstein BJ, Bray D, Cuellar A, Dronov S, Ginkel M, Gor V, Goryanin II, Hedley W, Hodgman TC, Hunter PJ, Juty NS, Kasberger JL, Kremling A, Kummer U, LeNovere N, Loew LM, Lucio D, Mendes P, Mjolsness ED, Nakayama Y, Nelson MR, Nielsen P, Sakurada T, Schaff JC, Shapiro BE, Shimizu TS, Spence HD, Stelling J, Takahashi K, Tomita M, Wagner J, and Wang J. The systems biology markup language (SBML): a medium for representation and exchange of biochemical network models. Bioinformatics 19: 524–531, 2003. 19. Hunter PJ. The IUPS Physiome Project: a framework for computational physiology. Prog Biophys Mol Biol 85: 551–569, 2004. 20. Hunter PJ and Borg TK. Integration from proteins to organs: the Physiome Project. Nat Rev Mol Cell Biol 4: 237–243, 2003. 21. Hunter PJ, McCulloch A, and ter Keurs HE. Modelling the mechanical properties of cardiac muscle. Prog Biophys Mol Biol 69: 289–331, 1998. 324 PHYSIOLOGY • Volume 20 • October 2005 • www.physiologyonline.org Downloaded from http://physiologyonline.physiology.org/ by 10.220.33.6 on June 15, 2017 descriptions of 30 different ion channels, receptors, transporters, pumps, exchangers, etc. The author of a new cell model will typically “tune” the parameters of these protein models to match measurements of the action potential profile and ion current responses to membrane voltage perturbations. The result is a plethora of models for different species and different parts of the heart (sinoatrial node, atrioventricular node, atria, Purkinje fiber, endothelial cell, mid-myocardial cell, epicardial cell, etc.); there are currently 37 such models in the CellML database. It would be highly preferable to define CellML models of the individual proteins based on patch-clamp data (with allowance for isoform variation) and to automatically import these models into a given cell type using the ontologies and a knowledge of the expression levels for each protein within a specified tissue. There is also an urgent requirement at the cell level for models that capture the 3-D structure of organelles and cytoplasm in the ~200 different human cell types. An important goal for the Physiome Project is also to use this modeling framework to help interpret clinical images for diagnostic purposes and to aid in the development of new medical devices. Another goal is to apply the anatomically and physiologically based models to virtual surgery, surgical training, and education. A longer-term goal is to help lower the cost of drug discovery by providing a rational multiscale and multiphysics modelingbased framework for dealing with the enormous complexity of physiological systems in the human body. The approach to computational physiology described here is at a very early stage of development but has the potential to help make physiological sense of the vast amount of cellular and molecular detail generated by biomedical scientists over the past 50 years. REVIEWS 22. Hunter PJ, Kohl P, and Noble D. Integrative models of the heart: achievements and limitations. Philos Trans R Soc Lond A Math Phys Sci 359: 1049–1054, 2001. 34. Nickerson DP, Smith NP, and Hunter PJ. A model of cardiac cellular electromechanics. Philos Trans R Soc Lond A Math Phys Sci 359: 1159–1172, 2001. 23. Hunter PJ, Pullan AJ, and Smaill BH. Modelling total heart function. Annu Rev Biomed Eng 5: 147–177, 2003. 35. Nielsen PMF, LeGrice IJ, Smaill BH, and Hunter PJ. Mathematical model of geometry and fibrous structure of the heart. Am J Physiol Heart Circ Physiol 260: H1365–H1378, 1991. 24. Hunter PJ, Robbins P, and Noble D. The IUPS Human Physiome Project. Pflügers Arch 445: 1–9, 2002. 25. Kohl P and Sachs F. Mechanoelectric feedback in cardiac cells. Philos Trans R Soc Lond A Math Phys Sci 359: 1173–1185, 2001. 26. Kohl P, Noble D, Winslow RL, and Hunter PJ. Computational modeling of biological system: tools and visions. Philos Trans R Soc Lond A Math Phys Sci 358: 579–610, 2000. 28. LeGrice IJ, Hunter PJ, Young AA, and Smaill BH. The architecture of the heart: a data-based model. Proc R Soc A 359: 1217–1232, 2001. 29. LeGrice IJ, Smaill BH, Chai LZ, Edgar SG, Gavin JB, and Hunter PJ. Laminar structure of the heart: ventricular myocyte arrangement and connective tissue architecture in the dog. Am J Physiol Heart Circ Physiol 269: H571–H582, 1995. 37. Noble D. The rise of computational biology. Nat Rev Mol Cell Biol 3: 460–463, 2002. 38. Noble D. Modeling the heart: from genes to cells to the whole organ. Science 295: 1678–1682, 2002. 39. Paterson D. Nitric oxide and the autonomic regulation of cardiac excitability. The G. L. Brown Prize Lecture. Exp Physiol 86: 1–12, 2001. 40. Pullan AJ, Cheng LK, Nash MP, Bradley CP, and Paterson DJ. Noninvasive electrical imaging of the heart: theory and model development. Ann Biomed Eng 29: 817–836, 2001. 41. Rosse C and Mejino JVL. A reference ontology for biomedical informatics: the foundational model of anatomy. J Biomed Inform 36: 478–500, 2003. 48. Smith NP, Nickerson DP, Crampin EJ, and Hunter PJ. Multi-scale computational modelling of the heart. Acta Numerica 13: 371–431, 2004. 49. Smith NP, Pullan AJ, and Hunter PJ. The generation of an anatomically accurate geometric coronary model. Ann Biomed Eng 28: 14–25, 2000. 50. Smith NP, Pullan AJ, and Hunter PJ. An anatomically based model of transient coronary blood flow in the heart. SIAM J Appl Math 62: 990–1018, 2002. 51. Soeller C and Cannell MB. Examination of the transverse tubular system in living cardiac rat myocytes by 2-photon microscopy and digital image-processing techniques. Circ Res 84: 266–275, 1999. 52. Stevens C and Hunter PJ. Sarcomere length changes in a 3D mathematical model of the pig ventricles. Prog Biophys Mol Biol 82: 229–241, 2003. 53. Tanskanen AJ, Greenstein JL, O’Rourke B, and Winslow RL. The role of stochastic and modal gating of cardiac L-type Ca2+ channels on early afterdepolarizations. Biophys J 88: 85–95, 2005. 42. Rudy Y. From genome to physiome: integrative models of cardiac excitation. Ann Biomed Eng 28: 945–950, 2000. 54. Tawhai MH, Hunter PJ, Tschirren J, Reinhardt JM, McLennan G, and Hoffman EA. CT-based geometry analysis and finite element models of the human and ovine bronchial tree. J Appl Physiol 97: 2310–2321, 2004. 31. McCulloch A. Modeling the human cardiome in silico. J Nucl Cardiol 7: 496–499, 2000. 43. Sansom MS, Shrivastava IH, Bright JN, Tate J, Capener CE, and Biggin PC. Potassium channels: structures, models, simulations. Biochim Biophys Acta 1565: 294–307, 2002. 55. Vaughan-Jones RD, Percy BE, Keener JP, and Spitzer KW. Intrinsic H+ ion mobility in the mammalian ventricular myocyte. J Physiol 541: 139–158, 2002. 32. McCulloch A, Bassingthwaighte JB, Hunter PJ, and Noble D. Computational biology of the heart: from structure to function. Prog Biophys Mol Biol 69: 151–572, 1998. 44. Saucerman JJ, Brunton LL, Michailova AP, and McCulloch AD. Modeling -adrenergic control of cardiac myocyte contractility in silico. J Biol Chem 278: 47997–48003, 2003. 56. Vaughan-Jones RD and Spitzer KW. Role of bicarbonate in the regulation of intracellular pH in the mammalian ventricular myocyte. Biochem Cell Biol 80: 579–596, 2002. 33. Nash MP and Hunter PJ. Computational mechanics of the heart. From tissue structure to ventricular function. J Elast 61: 113–141, 2000. 45. Saucerman JJ and McCulloch AD. Mechanistic systems models of cell signaling networks: a case study of myocyte adrenergic regulation. Prog Biophys Mol Biol 85: 261–278, 2004. 57. Young AA, Dokos S, Powell KA, Sturm B, McCulloch AD, Starling RC, McCarthy PM, and White RD. Regional heterogeneity of function in nonischemic dilated cardiomyopathy. Cardiovasc Res 49: 308–318, 2001. 30. Lloyd CM, Halstead MDB, and Nielsen PMF. CellML: its future, present and past. Prog Biophys Mol Biol 85: 433–450, 2004. 46. Saucerman JJ, Healy SN, Belik ME, Puglisi JL, and McCulloch AD. Proarrhythmic consequences of a KCNQ1 AKAP-binding domain mutation: computational models of whole cells and heterogeneous tissue. Circ Res 95: 1216–1224, 2004. PHYSIOLOGY • Volume 20 • October 2005 • www.physiologyonline.org 325 Downloaded from http://physiologyonline.physiology.org/ by 10.220.33.6 on June 15, 2017 27. LeGrice IJ, Hunter PJ, and Smaill BH. Laminar structure of the heart: a mathematical model. Am J Physiol Heart Circ Physiol 272: H2466–H2476, 1997. 36. Noble D. A modification of the Hodgkin-Huxley equations applicable to Purkinje fibre action and pace-maker potentials. J Physiol 160: 317–352, 1962. 47. Shannon TR and Bers DM. Integrated Ca2+ management in cardiac myocytes. Ann NY Acad Sci 1015: 28–38, 2004.