**** ****The goal of this manual is to teach biomodelers how to effectively use multi-scale, multi-cell simulation environment CompuCell3D to build, test, run and post-process simulations of biological phenomena occurring at single cell, tissue or even up to single organism levels. We first introduce basics of the Glazier-Graner-Hogeweg (GGH) model aka Cellular Potts Model (CPM) and then follow with essential information about how to use CompuCell3D and show simple examples of biological models implemented using CompuCell3D. Subsequently we will introduce more advanced simulation building techniques such as Python scripting and writing extension modules using C++. In everyday practice, however, the knowledge of C++ is not essential and C++ modules are usually developed by core CompuCell3D developers. However, to build sophisticated and biologically relevant models you will probably want to use Python scripting. Thus we strongly encourage readers to acquire at lease basic knowledge of Python. We don’t want to endorse any particular book but to guide users we might suggests names of the authors of the most popular books on Python programming: David Beazley, Mark Lutz, Mark Summerfield, Michael Dawson, Magnus Lie Hetland. ** ****I. Introduction** The last decade has seen fairly realistic simulations of single cells that can confirm or predict experimental findings. Because they are computationally expensive, they can simulate at most several cells at once. Even more detailed subcellular simulations can replicate some of the processes taking place inside individual cells. *E.g.*, Virtual Cell (`http://www.nrcam.uchc.edu `_) supports microscopic simulations of intracellular dynamics to produce detailed replicas of individual cells, but can only simulate single cells or small cell clusters. Simulations of tissues, organs and organisms present a somewhat different challenge: how to simplify and adapt single cell simulations to apply them efficiently to study, *in-silico*, ensembles of several million cells. To be useful, these simplified simulations should capture key cell-level behaviors, providing a phenomenological description of cell interactions without requiring prohibitively detailed molecular-level simulations of the internal state of each cell. While an understanding of cell biology, biochemistry, genetics, *etc.* is essential for building useful, predictive simulations, the hardest part of simulation building is identifying and quantitatively describing appropriate subsets of this knowledge. In the excitement of discovery, scientists often forget that modeling and simulation, by definition, require simplification of reality. One choice is to ignore cells completely, *e.g.*, Physiome *(****1****)* models tissues as continua with bulk mechanical properties and detailed molecular reaction networks, which is computationally efficient for describing dense tissues and non-cellular materials like bone, extracellular matrix (*ECM*), fluids, and diffusing chemicals *(****2****,****3****)*, but not for situations where cells reorganize or migrate. **Figure .**Detail of a typical two-dimensional GGH cell-lattice configuration. Each colored domain represents a single spatially-extended cell. The detail shows that each generalized cell is a set of cell-lattice sites (or *pixel*),, with a unique index, , here 4 or 7. The color denotes the cell type, . Multi-cell simulations are useful to interpolate between single-cell and continuum-tissue extremes because cells provide a natural level of abstraction for simulation of tissues, organs and organisms *(****4****)*. Treating cells phenomenologically reduces the millions of interactions of gene products to several behaviors: most cells can move, divide, die, differentiate, change shape, exert forces, secrete and absorb chemicals and electrical charges, and change their distribution of surface properties. The *Glazier-Graner-Hogeweg* (*GGH*) approach facilitates multiscale simulations by defining spatially-extended *generalized cells*, which can represent clusters of cells, single cells, sub-compartments of single cells or small subdomains of non-cellular materials. This flexible definition allows tuning of the level of detail in a simulation from intracellular to continuum without switching simulation framework to examine the effect of changing the level of detail on a macroscopic outcome, *e.g.*,**by switching from a coupled ordinary-differential-equation (*ODE*) *Reaction-Kinetics* (*RK*) model of gene regulation to a Boolean description or from a simulation that includes subcellular structures to one that neglects them. **II. GGH Applications** Because it uses a regular cell lattice and regular field lattices, GGH simulations are often faster than equivalent *Finite Element* (*FE*) simulations operating at the same spatial granularity and level of modeling detail, permitting simulation of tens to hundreds of thousands of cells on lattices of up to 10243 pixels on a single processor. This speed, combined with the ability to add biological mechanisms via terms in the effective energy, permit GGH modeling of a wide variety of situations, including: growth *(****5****-****9****)*, *(****10****-****12****)*, pigmentation *(****13****-****16****)*, *(****17****)*, *(****18****-****23****)*, the system *(****24****,****25****)*, colony growth*****(****26****,****27****)*, *myxobacteria******(****28****-****31****)*, cell differentiation *(****32****,****33****)*, *Dictyostelium discoideum* *(****34****-****37****)*, imulated evolution *(****38****-****43****)*,**developmental patterning *(****14****,****44****)*, extension *(****45****,****46****)*, formation *(****47****)*, *hydra* regeneration *(****48****,****49****)*, plant growth, patterning *(****50****,****51****)*, healing *(****47****,****52****,****53****)*, *(****54****-****57****)*, and limb-bud development*(****58****,****59****)*. **III. GGH Simulation Overview** All GGH simulations include a list of *objects*, a description of their *interactions* and *dynamics* and appropriate *initial conditions*. Objects in a GGH simulation are either generalized cells or *fields* in two dimensions (*2D*) or three dimensions (*3D*). Generalized cells are spatially-extended objects (Figure 1), which reside on a single *cell lattice* and may correspond to biological cells, sub-compartments of biological cells, or to portions of non-cellular materials, *e.g.*ECM, fluids, solids, *etc*.*(****8****,****48****,****60****-****72****)*. We denote a lattice site or *pixel* by a vector of integers, , the *cell index* of the generalized cell occupying pixel by and the *type* of the generalized cell by . Each generalized cell has a unique cell index and contains many pixels. Many generalized cells may share the same cell type. Generalized cells permit coarsening or refinement of simulations, by increasing or decreasing the number of lattice sites per cell, grouping multiple cells into clusters or subdividing cells into variable numbers of *subcells* (subcellular compartments). Compartmental simulation****permits detailed representation of phenomena like cell shape and polarity, force transduction, intracellular membranes and organelles and cell-shape changes. For details on the use of subcells, which we do not discuss in this chapter see *(****27****,****31****,****73****,****74****)*. Each generalized cell has an associated list of attributes, *e.g.*, *cell type*, *surface area* and *volume*, as well as more complex attributes describing a cell’s state, biochemical interaction networks, *etc.*. *Fields* are continuously-variable concentrations, each of which resides on its own lattice. Fields can represent chemical diffusants, non-diffusing ECM, *etc.*. Multiple fields can be combined to represent materials with textures, *e.g.*,**fibers. *Interaction descriptions* and *dynamics*define how GGH objects behave both biologically and physically. Generalized-cell behaviors and interactions are embodied primarily in the e*ffective energy*, which determines a generalized cell’s shape, motility, adhesion and response to extracellular signals. The effective energy mixes true energies, such as cell-cell adhesion with terms that mimic energies, *e.g.*, the response of a cell to a chemotactic gradient of a field *(****75****)*. Adding *constraints* to the effective energy allows description of many other cell properties, including osmotic pressure, membrane area, *etc*. *(****76****-****83****)*. The cell lattice evolves through attempts by generalized cells to move their boundaries in a caricature of cytoskeletally-driven cell motility. These movements, called *index-copy attempts*, change the effective energy, and we accept or reject each attempt with a probability that depends on the resulting *change of the effective energy*, *H*, according to an *acceptance function*. Nonequilibrium statistical physics then shows that the cell lattice evolves to locally minimize the total effective energy. The classical GGH implements a modified version of a classical stochastic Monte-Carlo pattern-evolution dynamics, called *Metropolis dynamics with Boltzmann acceptance* *(****84****,****85****)*. A *Monte Carlo Step*(*MCS*) consists of one index-copy attempt for each pixel in the cell lattice. *Auxiliary equations* describe cells’ absorption and secretion of chemical diffusants and extracellular materials (*i.e.*,**their interactions with fields), state changes within cells, mitosis, and cell death. These auxiliary equations can be complex, *e.g.*, detailed RK descriptions of complex regulatory pathways. Usually, state changes affect generalized-cell behaviors by changing parameters in the terms in the effective energy (*e.g.*, cell target volume or type or**the surface density of particular cell-adhesion molecules). *Fields* also evolve due to secretion, absorption, diffusion, reaction and decay according to *partial differential equations* (*PDE*s). While complex coupled-PDE models are possible, most simulations require only secretion, absorption, diffusion and decay, with all reactions described by ODEs running inside individual generalized cells. The movement of cells and variations in local diffusion constants (or diffusion tensors in anisotropic ECM) mean that diffusion occurs in an environment with moving boundary conditions and often with advection. These constraints rule out most sophisticated PDE solvers and have led to a general use of simple forward-Euler methods, which can tolerate them. The *initial condition* specifies the initial configurations of the cell lattice, fields, a list of cells and their internal states related to auxiliary equations and any other information required to completely describe the simulation. **III.A. Effective Energy** The core of GGH simulations is the *effective energy*, which describes cell behaviors and interactions. One of the most important effective-energy terms describes cell adhesion. If cells did not stick to each other and to extracellular materials, complex life would not exist *(****86****)*. Adhesion provides a mechanism for building complex structures, as well as for holding them together once they have formed. The many families of adhesion molecules (CAMs, cadherins, *etc.*) allow embryos to control the relative adhesivities of their various cell types to each other and to the noncellular ECM surrounding them, and thus to define complex architectures in terms of the cell configurations which minimize the adhesion energy. To represent variations in energy due to adhesion between cells of different types, we define a *boundary energy* that depends on , the *boundary energy per unit area* between two cells () of given types () at a *link*(the interface between two neighboring pixels): , \\\* MERGEFORMAT () where the sum is over all neighboring pairs of lattice sites and (note that the neighbor range may be greater than one), and the boundary-energy coefficients are symmetric, . \\\* MERGEFORMAT () In addition to boundary energy, most simulations include multiple constraints on cell behavior. The use of constraints to describe behaviors comes from the physics of classical mechanics. In the GGH context we write *constraint energies* in a general *elastic* form: . \\\* MERGEFORMAT () The constraint energy is zero if (the constraint is *satisfied*) and grows as *value* diverges from. The constraint is *elastic* because the exponent of 2 effectively creates an ideal spring pushing on the cells and driving them to satisfy the constraint. is the *spring constant*(a positive real number), which determines the *constraint strength*. Smaller values of allow the pattern to deviate more from the *equilibrium condition* (*i.e.*, the condition satisfying the constraint). Because the constraint energy decreases smoothly to a minimum when the constraint is satisfied, the energy-minimizing dynamics used in the GGH automatically drives any configuration towards one that satisfies the constraint. However, because of the stochastic simulation method, the cell lattice need not satisfy the constraint exactly at any given time, resulting in random fluctuations. In addition, multiple constraints may conflict, leading to configurations which only partially satisfy some constraints. Because biological cells have a given volume at any time, most GGH simulations employ a *volume constraint,* which restricts volume variations of generalized cells from their target volumes: , \\\* MERGEFORMAT () where for cell , denotes the *inverse compressibility* of the cell, is the number of pixels in the cell (its *volume*), and is the cell’s *target volume*. This constraint defines as the *pressure* inside the cell. A cell with has a positive internal pressure, while a cell with has a negative internal pressure. Since many cells have nearly fixed amounts of cell membrane, we often use a *surface- area constraint* of form: , \\\* MERGEFORMAT () where is the surface area of cell , is its target surface area, and is its *inverse membrane compressibility*. Adding the boundary energy and volume constraint terms together (equations ) and (4)), we obtain the basic *GGH effective energy*: \\\* MERGEFORMAT () **III.B. Dynamics** A GGH simulation consists of many attempts to copy cell indices between neighboring pixels. In CompuCell3D the default dynamical algorithm is *modified Metropolis dynamics*. During each index-copy attempt, we select a *target* pixel, , randomly from the cell lattice, and then randomly select one of its neighboring pixels, , as a *source* pixel. If they belong to the same generalized cell (*i.e.*, if ) we do not need copy index. Otherwise the cell containing the source pixel attempts to occupy the target pixel. Consequently, a successful index copy increases the volume of the *source* cell and decreases the volume of the *target* cell by one pixel. **Figure .****GGH representation of an index-copy attempt for two cells on a 2D square lattice – The “white” pixel (source) attempts to replace the “grey” pixel (target). The probability of accepting the index copy is given by equation*(7).* In the modified Metropolis algorithm we evaluate the change in the total effective energy due to the attempted index copy and accept the index-copy attempt with probability: , \\\* MERGEFORMAT () where is a parameter representing the *effective cell motility* (we can think of as the amplitude of cell-membrane fluctuations). Equation (7) is the *Boltzmann acceptance function*. Users can define other acceptance functions in CompuCell3D. The conversion between MCS and experimental time depends on the average values of. MCS and experimental time are proportional in biologically-meaningful situations *(****87****-****90****)*. **III.C. Algorithmic Implementation of Effective-Energy Calculations** Consider an effective energy consisting of boundary-energy and volume-constraint terms as in equation (6). After choosing the source () and destination () pixels (the cell index of the source will overwrite the target pixel if the index copy is accepted), we calculate the change in the effective energy that would result from the copy. We evaluate the change in the boundary energy and volume constraint as follows. First we visit the target pixel’s neighbors (). If the neighbor pixel belongs to a different generalized cell from the target pixel, *i.e.*, when (see equation )), we decrease by . If the neighbor belongs to a cell different from the source pixel () we increase by . The change in volume-constraint energy is evaluated according to: \\\* MERGEFORMAT () where and denote the volumes of the generalized cells containing the source and target pixels, respectively. In this example, we could calculate the change in the effective energy locally, *i.e.*, by visiting the neighbors of the target of the index copy. Most effective energies are quasi-local, allowing calculations of similar to those presented above. The locality of the effective energy is crucial to the utility of the GGH approach. If we had to calculate the effective energy for the entire cell lattice for each index-copy attempt, the algorithm would be prohibitively slow. **Figure .**Calculating changes in the boundary energy and the volume-constraint energy on a nearest-neighbor square lattice. For longer-range interactions we use the appropriate list of neighbors, as shown in 4 and 1. Longer-range interactions are less anisotropic but result in slower simulations. **Figure .****Locations of *n*th-nearest neighbors on a 2D square lattice and a 2D hexagonal lattice. 2D Square Lattice Neighbor Order Number of Neighbors Euclidian Distance 1 4 1 2 4 3 4 2 4 8 **Table .**Multiplicity and Euclidian distances of *n*th-nearest neighbors for 2D square and hexagonal lattices. The number of *n*th neighbors and their distances from the central pixel differ in a 3D lattice. CompuCell3D calculates distance between neighbors by setting the volume of a single pixel (whether in 2D or 3D) to 1. **IV. CompuCell3D** One advantage of the GGH model over alternative techniques is its mathematical simplicity. We can implement fairly easily a computer program that initializes the cell lattice and fields, performs index copies and displays the results. In the 15 years since the GGH model was developed, researchers have written numerous programs to run GGH simulations. Because all GGH implementations share the same logical structure and perform similar tasks, much of this coding effort has gone into rewriting code already developed by someone else. This redundancy leads to significant research overhead and unnecessary duplication of effort and makes model reproduction, sharing and validation needlessly cumbersome. To overcome these problems, we developed CompuCell3D as a framework for GGH simulations *(****91****,****92****)*. Unlike specialized research code, CompuCell3D is a *simulation* *environment* that allows researchers to rapidly build and run shareable GGH-based simulations. It greatly reduces the need to develop custom code and its adherence to open-source standards ensures that any such code is shareable. CompuCell3D supports non-programmers by providing visualization tools, an *eXtensible Markup Language* (*XML*) interface for defining simulations, and the ability to extend the framework through specialized modules. The C++ computational kernel of CompuCell3D is also accessible using the open-source scripting language Python, allowing users to create complex simulations without programming in lower-level languages such as C or C++. Unlike typical research code, changing a simulation does not require recompiling CompuCell3D. Users define simulations using *CompuCell3D XML* (*CC3DML*) *configuration files* and/or Python scripts. CompuCell3D reads and parses the CC3DML configuration file and uses it to define the basic simulation structure, then initializes appropriate Python services (if they are specified) and finally executes the underlying simulation algorithm. CompuCell3D is modular: each module carries out a defined task. CompuCell3D terminology calls modules associated with index copies or index-copy attempts *plugins*. Some plugins calculate changes in effective energy, while others (*lattice monitors*) react to accepted index copies, *e.g.*, by updating generalized cells’ surface areas, volumes or lists of neighbors. Plugins may depend on other plugins. For example, the Volume plugin (which calculates the volume-energy constraint in equation (4)) depends on VolumeTracker (a lattice monitor), which, as its name suggests, tracks changes in generalized cells’ volumes. When implicit plugin dependencies exist, CompuCell3D automatically loads and initializes dependent plugins. In addition to plugins, CompuCell3D defines modules called *steppables* which run either repeatedly after a defined intervals of Monte Carlo Steps or once at the beginning or end of the simulation. Steppables typically define initial conditions, alter cell states, update fields or output intermediate results. 5 shows the relations among index-copy attempts, Monte Carlo Steps, steppables and plugins. **Figure .**Flow chart of the GGH algorithm as implemented in CompuCell3D. CompuCell3D includes a *Graphical User Interface* (*GUI*) and visualization tool, *CompuCell Player* (also referred to as *Player*). From Player the user opens a CC3DML configuration file and/or Python file and hits the “Play” button to run the simulation. Player allows users to define multiple 2D or 3D visualizations of generalized cells, fields or various vector plots while the simulation is running and save them automatically for post-processing. **V. Building CC3DML-Based Simulations Using CompuCell3D** To show how to build simulations in CompuCell3D, the reminder of this chapter provides a series of examples of gradually increasing complexity. For each example we provide a brief explanation of the physical and/or biological background of the simulation and listings of the CC3DML configuration file and Python scripts, followed by a detailed explanation of their syntax and algorithms. We can specify many simulations using only a simple CC3DML configuration file. We begin with three examples using only CC3DML to define simulations. **V.A A Short Introduction to XML** XML is a text-based data-description language, which allows standardized representations of data. XML syntax consists of lists of *elements*, each either contained between opening () and closing () tags: *ElementText* or of form: We will denote the syntax as a *tag pair*. The opening tag of an XML element may contain additional *attributes* characterizing the element. The content of the XML element (*ElementText* in the above example) and the values of its attributes (*text1*, *attribute\_text1*, *attribute\_text2*) are strings of characters. Computer programs that read XML may interpret these strings as other data types such as integers, Booleans or floating point numbers. XML elements may be nested. The simple example below defines an element Cell with subelements (represented as nested XML elements) Nucleus and Membrane assigning the element Nucleus an attribute Size set to "10" and the element Membrane an attribute Area set to "20.5", and setting the value of the Membrane element to Expanding: Expanding Although XML parsers ignore indentation, all the listings presented in this chapter are block-indented for better readability. **V.B Grain-Growth Simulation** One of the simplest CompuCell3D simulations mimics crystalline grain growth or *annealing*. Most simple metals are composed of microcrystals, or *grains*, each of which has a particular crystalline-lattice orientation. The atoms at the surfaces of these grains have a higher energy than those in the bulk because of their missing neighbors. We can characterize this excess energy as a *boundary energy*. Atoms in convex regions of a grain's surface have a higher energy than those in concave regions, in particular than those in the concave face of an adjoining grain, because they have more missing neighbors. For this reason, an atom at a convex curved boundary can reduce its energy by “hopping” across the grain boundary to the concave side *(****62****)*. The movement of atoms moves the grain boundaries, lowering the net configuration energy through *annealing* or *coarsening*, with the net size of grains growing because of grain disappearance. Boundary motion may require thermal activation because atoms in the space between grains may have higher energy than atoms in grains. The effective energy driving grain growth is simply the boundary energy in equation ). In CompuCell3D, we can represent grains as generalized cells. CC3DML 1 defines our grain-growth simulation. 10000 5 Periodic Periodic 3 0 5 0 3 0 5 Grain **Listing .****CC3DML configuration file for 2D grain-growth simulation. Each CC3DML configuration file begins with the tag and ends with the tag. A CC3DML configuration file contains three sections in the following sequence: the *lattice section* (contained within the tag pair), the *plugins section*, and the *steppables section*. The lattice section defines global parameters for the simulation: cell-lattice and field-lattice dimensions (specified using the syntax ), the number of Monte Carlo Steps to run (defined within the tag pair) the effective cell motility (defined within the tag pair) and boundary conditions. The default boundary conditions are *no-flux*. However, in this simulation, we have changed them to be periodic along the *x* and *y* axes by assigning the value Periodic to the and tag pairs. The value set by the tag pair defines the range over which source pixels are selected for index-copy attempts (see 4 and 1). The plugins section lists the plugins the simulation will use. The syntax for all plugins which require parameter specification is: <*ParameterSpecification*/> The CellType plugin uses the parameter syntax to map verbose generalized-cell-type names to numeric cell TypeIds for all generalized-cell types. It does not participate directly in index copies, but is used by other plugins for cell-type-to-cell-index mapping. Even though the grain-growth simulation fills the entire cell lattice with cells of type Grain, the current implementation of CompuCell3D requires that all simulations define the Medium cell type with TypeId 0. Medium is a special cell type with unconstrained volume and surface area that fills all cell-lattice pixels unoccupied by cells of other types. The Contact plugin calculates changes in the boundary energy defined in equation ) for each index-copy attempt. The parameter syntax for the Contact plugin is: *EnergyValue* where *TypeName1* and *TypeName2* are the names of the cell types and *EnergyValue* is the boundary-energy coefficient,, between cells of *TypeName1* and *TypeName2* (see equation )). The tag pair specifies the interaction range of the boundary energy. The default value of this parameter is 1. The steppables section includes only the UniformInitializer steppable. All steppables have the following syntax: <*ParameterSpecification*/> The Frequency attribute is optional and by default is 1 MCS. CompuCell3D simulations require specification of initial condition. The simplest way to define the initial cell lattice is to use the built-in initializer steppables, which construct simple regions filled with generalized cells. The UniformInitializer steppable in the grain-growth simulation defines one or more rectangular (parallelepiped in 3D) regions filled with generalized cells of user selectable types and sizes. We enclose each region definition within a tag pair. We use the and tags to describe the boundaries of the region, The tag pair defines the size of the square (cubical in 3D) generalized cells and the tag pair creates space between neighboring cells. The tag pair lists the types of generalized cells. The grain-growth simulation uses only one cell type, Grain, but we can also initialize cells using types randomly chosen from the list, as in 2. 0 5 Condensing,NonCondensing **Listing .****CC3DML code excerpt using the UniformInitializer steppable to initialize a rectangular region filled with 5 x 5 pixel generalized cells with randomly-assigned cell types (either Condensing or NonCondensing). The coordinate values in BoxMax element must be one more than the coordinates of the corresponding corner of the region to be filled. So to fill a square of side 10 beginning with pixel location (5,5) we use the following region-boundary specification: Listing the same type multiple times results in a proportionally higher fraction of generalized cells of that type. For example, Condensing,Condensing,NonCondensing will allocate approximately 2/3 of the generalized cells to type Condensing and 1/3 to type NonCondensing. UniformInitializer allows specification of multiple regions. Each region is independent and can have its own cell sizes, types and cell spacing. If the regions overlap, later-specified regions overwrite earlier-specified ones. If region specification does not cover the entire lattice, uninitialized pixels have type Medium. 6 shows sample output generated by the grain-growth simulation. **Figure .****Snapshots of the cell-lattice configuration for the grain-growth simulation on a 100 x 100 pixel 3rd-neighbor square lattice, as specified in *1*. Boundary conditions are periodic in both directions. One advantage of GGH simulations compared to FE simulations is that going from 2D to 3D is easy. To run a 3D grain-growth simulation on a 100 x 100 x 100 lattice we only need to make the following replacements in 1: and, Grain growth simulations are particularly sensitive to lattice anisotropy, so running them on lower-anisotropy lattices is desirable. Longer-range lattices are less anisotropic but cause simulations to run slower. Fortunately a hexagonal lattice of a given range is less anisotropic than a square lattice of the same range. To run the grain-growth simulation on a hexagonal lattice, we add Hexagonal to the lattice section in 1 and change the two occurrences of: 31 7 shows snapshots for this simulation. **Figure .****Snapshots of the cell-lattice configuration for the grain-growth simulation on a 100 x 100 pixel 1st -neighbor hexagonal lattice as specified in *1* with substitutions described in the text. The *x* and *y* length units in an hexagonal lattice differ, resulting in differing *x* and *y* dimensions for a cell lattice with an equal number of pixels in the *x* and *y* directions. One inconvenience of the current implementation of CompuCell3D is that it does not automatically rescale parameter values when interaction range, lattice dimensionality or lattice type change. When changing these attributes, users must recalculate parameters to keep the underlying physics of the simulation the same. CompuCell3D dramatically reduces the amount of code necessary to build and run a simulation. The grain-growth simulation took about 25 lines of CC3DML instead of 1000 lines of C, C++ or Fortran. **V.C Cell-Sorting Simulation** Cell sorting is an experimentally-observed phenomenon in which cells with different adhesivities are randomly mixed and reaggregated. They can spontaneously sort to reestablish coherent homogenous domains *(****93****,****94****)*. Sorting is a key mechanism in embryonic development. The grain-growth simulation used only one type of generalized cell. Simulating sorting of two types of biological cell in an aggregate floating in solution is slightly more complex. 3 shows a simple cell-sorting simulation. It is similar to 1 with a few additional modules (shown in **bold**). The effective energy is that in equation (6). 10000 10 2 25 2.0 0 16 2 11 16 16 2 **** 0 5 40
Condensing,NonCondensing **Listing .****CC3DML configuration file simulating cell sorting between Condensing and NonCondensing cell types. Highlighted text indicates modules absent in *1*. Notice how little modification of the grain-growth CC3DML configuration file this simulation requires. The main change from 1 to the lattice section is that we omit the boundary condition specification and use default no-flux boundary conditions. In the CellType plugin we introduce the two cell types, Condensing and NonCondensing, in place of Grain. In addition we do not the fill lattice completely with Condensing and NonCondensing cells so the interactions with Medium become important. The boundary-energy matrix in the Contact plugin thus requires entries for the additional cell-type pairs. The hierarchy of boundary energies listed results in cell sorting. We also add the Volume plugin, which calculates the volume-constraint energy as given in equation (4). In this plugin the tag pair sets target volume for both Condensing cells and NonCondensing and the tag pair sets the constraint strength for both cell types. We will see later how to define volume-constraint parameters for each cell type or each cell individually. In the cell-sorting simulation we initialize the cell lattice using the BlobInitializer steppable which specifies circular (or spherical in 3D) regions filled with square (or cubical in 3D) cells of user-defined size and types. The syntax is very similar to that for UniformInitializer. Looking in detail at the syntax of BlobInitializer in 3, the tag pair defines the radius of a circular (or spherical) domain of cells in pixels. The
tag, with syntax
, defines the coordinates of the center of the domain. The remaining tags are the same as for UniformInitializer. As with UniformInitializer, we can define multiple regions. We can use both UniformInitializer and BlobInitializer in the same simulation. In the case of overlap, later-specified regions overwrite earlier ones. We show snapshots of the cell-sorting simulation in 8. The less cohesive NonCondensing cells engulf the more cohesive Condensing cells, which cluster and form a single central domain. By changing the boundary energies we can produce other cell-sorting patterns *(****95****,****96****)*. **Figure .****Snapshots of the cell-lattice configurations for the cell-sorting simulation in *3*. The boundary-energy hierarchy drives NonCondensing (light grey) cells to surround Condensing (dark grey) cells. The white background denotes surrounding Medium*.* **V.D Bacterium-and-Macrophage Simulation** In the two simulations we have presented so far, the cellular pattern develops without fields. Often, however, biological patterning mechanisms require us to introduce and evolve chemical fields and to have cells’ behaviors depend on the fields. To illustrate the use of fields, we model the *in vitro* behavior of bacteria and macrophages in blood. In the famous experimental movie taken in the 1950s by David Rogers at Vanderbilt University, the macrophage appears to chase the bacterium, which seems to run away from the macrophage. We can model both behaviors using cell secretion of diffusible chemical signals and movement of the cells in response to the chemical (*chemotaxis*): the bacterium secretes a signal (a *chemoattractant*) that attracts the macrophage and the macrophage secretes a signal (a *chemorepellant*) which repels the bacterium *(****97****)*. 4 shows the CC3DML configuration file for the bacterium-and-macrophage simulation. 100000 20 Hexagonal 0 150 8 150 15 8 0 0 150 150 150 150 30 150 150 2 ATTR 0.10 0.00005 Wall Red 200 REP 0.10 0.001 Wall Red 200 bacterium\_macrophage\_2D\_wall\_v3.pif **Listing .** CC3DML configuration file for the bacterium-and-macrophage simulation. The simulation has five generalized-cell types: Medium, Bacterium, Macrophage, Red blood cells and a surrounding Wall. It also has two diffusible fields, representing a chemoattractant, ATTR, and a chemorepellent, REP. Because the default boundary-energy between any generalized-cell type and the edge of the cell lattice is zero, we define a surrounding wall to prevent cells from sticking to the cell-lattice boundary. As in our previous simulations, we assign cell types using the CellType plugin. Note the new syntax in the line specifying the cell type making up the walls: The Freeze="" attribute excludes generalized cells of type Wall from participating in index copies, which makes the walls immobile. We replace the Volume plugin with VolumeFlex and add the plugin SurfaceFlex. These plugins allow independent assignment of target values and constraint strengths in the volume-constraint and surface-constraint energies (equations (4) and (5)). These plugins require a line for each generalized-cell type, specifying the type name and the target volume (or target surface area), and (or ) for that generalized-cell type, *e.g.*: We implement the actual bacterium-macrophage “chasing” mechanism using the Chemotaxis plugin, which specifies how a generalized cell of a given type responds to a field. The Chemotaxis plugin biases a cell’s motion up or down a field gradient by changing the calculated effective-energy change used in the acceptance function, equation (7). For a field : \\\* MERGEFORMAT () where is the chemical field at the index-copy target pixel, the field at the index-copy source pixel, and the strength and direction of chemotaxis. If and , then is negative, increasing the probability of accepting the index copy in equation (7). The net effect is that the cell moves up the field gradient with a velocity . If is negative, the opposite occurs, and the cell will move down the field gradient. Plugins with more sophisticated calculations (*e.g.*, including response saturation) are available within CompuCell3D (see the *CompuCell3D User Guide*).**** **** **Figure .****Connecting a field to GGH dynamics using a chemotaxis-energy term. The difference in the value of the field at the source, , and target, , pixels changes the of the index-copy**attempt. Here**and*,* so,**increasing the probability of accepting the index-copy attempt in equation *(7)*. In the Chemotaxis plugin we must identify the names of the fields, where the field information is stored, the list of the generalized-cell types that will respond to the fields, and the strength and direction of the response (Lambda =). The information for each field is specified using the syntax: In our current example, the first field, named ATTR, is stored in FlexibleDiffusionSolverFE. Macrophage cells are attracted to ATTR with . None of the other cell types responds to ATTR. Similarly, Bacterium cells are repelled by REP with . Keep in mind that fields are *not* created within the Chemotaxis plugin, which only specifies how different cell types respond to the fields. We define and store the fields elsewhere. Here, we use the FlexibeDiffusionSolverFE steppable as the source of our fields. The FlexibleDiffusionSolverFE steppable is the main CompuCell3D tool for defining diffusing fields, which evolve according to the diffusion equation: , \\\* MERGEFORMAT () where is the field concentration and , and denote the diffusion constant (in m2/s), decay constant (in s-1) and secretion rates (in concentration/s) of the field, respectively. , , and may vary with position and cell-lattice configuration. As in the Chemotaxis plugin, we may define the behaviors of multiple fields, enclosing each one within tag pairs. For each field defined within a tag pair, users****provide values for the name of the field (using the tag pair), the diffusion constant (using the tag pair) , and the decay constant (using the tag pair). Forward-Euler methods are numerically unstable for large diffusion constants, limiting the maximum nominal diffusion constant allowed in CompuCell3D simulations. However, by increasing the PDE-solver calling frequency, which reduces the effective time step, CompuCell3D can simulate arbitrarily large diffusion constants. For more information, see the *CompuCell3D User Guide*. Each optional **tag pair, with syntax: *cell\_type* prevents the field from diffusing into field-lattice pixels where the corresponding cell-lattice pixel, , is occupied by a cell, , of the specified type. In our case, chemical fields do not diffuse into the pixels occupied by Wall or Red cells. The optional tag pair defines a subsection which identifies cells types that secrete or absorb the field and the rates of secretion: *real\_rate1* *real\_rate2* A negative *rate* simulates absorption.**In the bacterium and macrophage simulation, Bacterium cells secrete ATTR and Macrophage cells secrete REP. We load the initial configuration for the bacterium-and-macrophage simulation using the PIFInitializer steppable. Many simulations require initial generalized-cell configurations that we cannot easily construct from primitive regions filled with cells using BlobInitializer and UniformInitializer. To allow maximum flexibility, CompuCell3D can read the initial cell-lattice configuration from *Pixel Initialization Files* (*PIFs*). A PIF is a text file that allows users to assign multiple rectangular (parallelepiped in 3D) pixel regions or single pixels to particular cells. Each line in a PIF has the syntax: *Cell\_id Cell\_type x\_low x\_high y\_low y\_high z\_low z\_high* where *Cell\_id* is a unique cell index. A PIF may have multiple, possibly non-adjacent, lines starting with the same *Cell\_id*; all lines with the same *Cell\_id* define pixels of the same generalized cell. The values *x\_low*, *x\_high*, *y\_low*, *y\_high*, *z\_low* and *z\_high* define rectangles (parallelepipeds in 3D) of pixels belonging to the cell. In the case of overlapping pixels, a later line overwrites pixels defined by earlier lines. The following line describes a 6 x 6-pixel square cell with cell index 0 and type Amoeba: 0 Amoeba 10 15 10 15 0 0 If we save this line to the file 'amoebae.pif', we can load it into a simulation using the following syntax: amoebae.pif 5 illustrates how to construct arbitrary shapes using a PIF. Here we define two cells with indices 0 and 1, and cell types Amoeba and Bacterium, respectively. The main body of each cell is a 6 x 6 square to which we attach additional pixels. 0 Amoeba 10 15 10 15 0 0 1 Bacterium 25 30 25 30 0 0 0 Amoeba 16 16 15 15 0 0 1 Bacterium 25 27 31 35 0 0 **Listing .** Simple PIF initializing two cells, one each of type Bacterium and Amoeba. All lines with the same cell index (first column) define a single cell. 10 shows the initial cell-lattice configuration specified in 5: **Figure .** Initial configuration of the cell lattice based on the PIF in *5*. In practice, because constructing complex PIFs by hand is cumbersome, we generally use custom-written scripts to generate the file directly, or convert images stored in graphical formats (*e.g.*, gif, jpeg, png) from experiments or other programs. 6 shows the PIF for the bacterium-and-macrophage simulation. 0 Red 10 20 10 20 0 0 1 Red 10 20 40 50 0 0 2 Red 10 20 70 80 0 0 3 Red 40 50 0 10 0 0 4 Red 40 50 30 40 0 0 5 Red 40 50 60 70 0 0 6 Red 40 50 90 95 0 0 7 Red 70 80 10 20 0 0 8 Red 70 80 40 50 0 0 9 Red 70 80 70 80 0 0 10 Wall 0 99 0 1 0 0 10 Wall 98 99 0 99 0 0 10 Wall 0 99 98 99 0 0 10 Wall 0 1 0 99 0 0 11 Bacterium 5 5 5 5 0 0 12 Macrophage 35 35 35 35 0 0 13 Bacterium 65 65 65 65 0 0 14 Bacterium 65 65 5 5 0 0 15 Bacterium 5 5 65 65 0 0 16 Macrophage 75 75 95 95 0 0 17 Red 24 28 10 20 0 0 18 Red 24 28 40 50 0 0 19 Red 24 28 70 80 0 0 20 Red 40 50 14 20 0 0 21 Red 40 50 44 50 0 0 22 Red 40 50 74 80 0 0 23 Red 54 59 90 95 0 0 24 Red 70 80 24 28 0 0 25 Red 70 80 54 59 0 0 26 Red 70 80 84 90 0 0 27 Macrophage 10 10 95 95 0 0 **Listing .** PIF defining the initial cell-lattice configuration for the bacterium-and-macrophage simulation. The file is stored as 'bacterium\_macrophage\_2D\_wall\_v3.pif'. In 4 we read the cell lattice configuration from the file 'bacterium\_macrophage\_2D\_wall\_v3.pif' using the lines: bacterium\_macrophage\_2D\_wall\_v3.pif 11 shows snapshots of the bacterium-and-macrophage simulation. By adjusting the properties and number of bacteria, macrophages and red blood cells and the diffusion properties of the chemical fields, we can build a surprisingly good reproduction of the experiment. **Figure .****Snapshots of the bacterium-and-macrophage simulation from *4* and the PIF in *6* saved in the file 'bacterium\_macrophage\_2D\_wall\_v3.pif'. The upper row shows the cell-lattice configuration with the Macrophages in grey, Bacteria in white, red blood cells in dark grey and Medium in blue. Middle row shows the concentration of the chemoattractant ATTR secreted by the *Bacteria*. The bottom row shows the concentration of the chemorepellant REPL secreted by the Macrophages. The bars at the bottom of the field images show the concentration scales (blue, low concentration, red, high concentration). **VI. Python Scripting** CC3DML is convenient for building simple simulations such as those we presented above. To describe more complex simulations, CompuCell3D allows users to write specialized, shareable modules in C/C++ (through the *CompuCell3D Application Programming Interface*, or *CC3D API*) or Python (through a Python-scripting interface). C and C++ modules have the advantage that they run at native speed. However, developing them requires knowledge of both C/C++ and the CC3D API, and their integration with CompuCell3D requires recompilation of the source code. Python module development is less complicated, since Python has simpler syntax than C/C++ and users can modify and extend a library of Python-module templates included with CompuCell3D. Moreover, Python modules do not require recompilation. Tasks performed by CompuCell3D modules either relate to index-copy attempts (plugins) or run either once, at the beginning or end of a simulation, or once every several MCS (steppables). Tasks run every index-copy attempt, like effective-energy-term calculations, are the most frequently-called tasks in a GGH simulation and writing them in Python may slow simulations. However, steppables and lattice monitors are good candidates for Python implementation and cause negligible performance degradation. Python implementations are suitable for most cell-parameter adjustments that depend on the state of the simulation, *e.g.*, simulating cell growth in response to a chemical, cell-type differentiation and changes in cell-cell adhesion. **VI.A A Short Introduction to Python** Python is an object-oriented scripting language with all the syntactic constructs present in any modern programming language. Python supports popular flow-control statements such as if-elif-else conditional instructions and for and while loops. Unlike C/C++, Python does not use ';' to end lines or '{' and '}' to define code blocks. Instead, Python relies on indentation to define blocks of code. if statements, for or while loops and their subsections are created by a ':' and increasing the level of indentation. The end of a block is indicated by a decrease in the level of indentation. Python uses the '=' operator for assignments and '==' for checking equality between objects. For example, in the following code: b=2 if b==2: a=10 for c in range(0,a): b=a+c print b we indent the body of the if statement and the body of the inner for loop. The for loop is executed inside the if statement. a=0 assigns the variable a a value of 10, while b==2 is true if b has a value of 2. The for loop assigns the variable c values 0 through a-1 and executes instructions inside the loop body. As an object-oriented language, Python supports *classes*, *inheritance* and *polymorphism*. Accessing *members* of *objects* uses the '.' operator. For example, to access the real part of a complex number, we use the following code: a=complex(2,3) a=1.5+0.5j print a.real Here, real is a member of the Python class complex, which represents complex numbers. If the object has composite subobjects, we use the '.' operator recursively: object.subobject.member\_of\_subobject Users may define Python objects without declaring their type. A single data structure such as a list or dictionary can store objects of multiple types. Python provides automatic memory management, which frees users from remembering to deallocate memory for objects that are no longer used. Long source code lines can be carried over to the following line using the '\\' character: very\_long\_variable\_name = \\ very\_long\_variable\_name \* very\_important\_constant Note that double underscore '\_\_' has a reserved meaning in Python and should not be confused with a single underscore '\_'. We will present additional Python features in the subsequent sections and explain step-by-step some basic concepts of Python programming (for more on Python, see *Learning Python*, by Mark Lutz *(****98****)*). For more information on Python scripting in CompuCell3D, see our *Python Tutorials* and *CompuCell3D User Guide* (available from the CompuCell3D website, *www.compucell3d.org*). **VI.B Building Python-Based CompuCell3D Simulations** Python scripting allows users to augment their CC3DML configuration files with Python scripts or to code their entire simulations in Python (in which case the Python script looks very similar to the CC3DML script it replaces). 7 shows the standard block of template code for running a Python script in conjunction with a CC3DML configuration file. import sys from os import environ from os import getcwd import string sys.path.append(environ["PYTHON\_MODULE\_PATH"]) import CompuCellSetup sim,simthread = CompuCellSetup.getCoreSimulationObjects() #Create extra player fields here or add attributes CompuCellSetup.initializeSimulationObjects(sim,simthread) #Add Python steppables here steppableRegistry=CompuCellSetup.getSteppableRegistry() CompuCellSetup.mainLoop(sim,simthread,steppableRegistry) **Listing**. Basic Python template to run a CompuCell3D simulation through a Python interpreter. Later examples will be based on this script. The import sys line provides access to standard functions and variables needed to manipulate the Python runtime environment. The next two lines, from os import environ from os import getcwd import environ and getcwd housekeeping functions into the current *namespace* (*i.e.*, current script) and are included in all our Python programs. In the next three lines, import string sys.path.append(environ["PYTHON\_MODULE\_PATH"]) import CompuCellSetup we import the string module, which contains convenience functions for performing operations on strings of characters, set the search path for Python modules and import the CompuCellSetup module, which provides a set of convenience functions that simplify initialization of CompuCell3D simulations. Next, we create and initialize the core CompuCell3D modules: sim,simthread = CompuCellSetup.getCoreSimulationObjects() CompuCellSetup.initializeSimulationObjects(sim,simthread) We then create a steppable *registry* (a Python *container* that stores steppables, *i.e.*, a list of all steppables that the Python code can access) and pass it to the function that runs the simulation: steppableRegistry=CompuCellSetup.getSteppableRegistry() CompuCellSetup.mainLoop(sim,simthread,steppableRegistry) In the next section, we extend this template to build a simple simulation. **VI.C Cell-Type-Oscillator Simulation** Suppose that we would like to add a caricature of oscillatory gene expression to our cell-sorting simulation (3) so that cells exchange types every 100 MCS. We will implement the changes of cell types using a Python steppable, since it occurs at intervals of 100 MCS. 8 shows the changes to the Python template in 7 that are necessary to create the desired type switching (changes are shown in **bold**). import sys from os import environ from os import getcwd import string sys.path.append(environ["PYTHON\_MODULE\_PATH"]) import CompuCellSetup sim,simthread = CompuCellSetup.getCoreSimulationObjects() **from PySteppables import \*** **class TypeSwitcherSteppable(SteppablePy):** def \_\_init\_\_(self,\_simulator,\_frequency=100): SteppablePy.\_\_init\_\_(self,\_frequency) self.simulator=\_simulator self.inventory=self.simulator.getPotts().getCellInventory() self.cellList=CellList(self.inventory) def step(self,mcs): for cell in self.cellList: if cell.type==1: cell.type=2 elif (cell.type==2): cell.type=1 else: print "Unknown type. In cellsort simulation there should\\ only be two types 1 and 2" #Create extra player fields here or add attributes CompuCellSetup.initializeSimulationObjects(sim,simthread) #Add Python steppables here steppableRegistry=CompuCellSetup.getSteppableRegistry() **typeSwitcherSteppable=TypeSwitcherSteppable(sim,100);** **steppableRegistry.registerSteppable(typeSwitcherSteppable)** CompuCellSetup.mainLoop(sim,simthread,steppableRegistry) **Listing .** Python script expanding the template code in *7* into a simple TypeSwitcherSteppable steppable. The code illustrates dynamic modification of cell parameters using a Python script. Lines added to *7* are shown in **bold**. **** A CompuCell3D steppable is a *class* (a type of *object*) that holds the parameters and functions necessary for carrying out a task. Every steppable defines at least 4 functions: \_\_init\_\_(self, \_simulator, \_frequency), start(self), step(self, mcs) and finish(self). CompuCell3D calls the start(self) function once at the beginning of the simulation before any index-copy attempts. It calls the step(self, mcs) function periodically after every \_frequency MCS. It calls the finish(self) function once at the end of the simulation. 8 does not have explicit start(self) or finish(self) functions. Instead, the class definition : class TypeSwitcherSteppable(SteppablePy): causes the TypeSwitcherSteppable to inherit components of the SteppablePy class. SteppablePy contains default definitions of the start(self), step(self,mcs) and finish(self) functions. Inheritance reduces the length of the user-written Python code and ensures that the TypeSwitcherSteppable object has all needed components. The line: from PySteppables import \* makes the content of 'PySteppables.py' file (or module) available in the current namespace. The PySteppables module includes the SteppablePy *base class*. The \_\_init\_\_ function is a *constructor* that accepts user-defined parameters and initializes a steppable object. Consider the \_\_init\_\_ function of the TypeSwitcherSteppable: def \_\_init\_\_(self,\_simulator,\_frequency=100): SteppablePy.\_\_init\_\_(self,\_frequency) self.simulator=\_simulator self.inventory=self.simulator.getPotts().getCellInventory() self.cellList=CellList(self.inventory) In the def line, we pass the necessary parameters: self (which is used in Python to access class variables from within the class), \_simulator (the main CompuCell3D kernel object which runs the simulation), and \_frequency (which tells steppableRegistry how often to run the steppable, here, every 100 MCS). Next we call the constructor for the inheritance class, SteppablePy, as required by Python. The following statement: self.simulator=\_simulator assigns to the class variable self.simulator a reference to \_simulator object, passed from the main script. We can think about Python reference as a pointer variable that stores the address of the object but not a copy of the object itself. The last two lines construct a list of all generalized cells in the simulation, a *cell inventory*, which allows us to visit all the cells with a simple for loop to perform various tasks. The cell inventory is a dynamic structure, *i.e.*, it updates automatically when cells are created or destroyed during a simulation. The section of the TypeSwitcherSteppable steppable which implements the cell-type switching is found in the step(self, mcs) function: def step(self,mcs): for cell in self.cellList: if cell.type==1: cell.type=2 elif (cell.type==2): cell.type=1 else: print "Unknown type" Here we use the cell inventory to iterate over all cells in the simulation and reassign their cell types between cell.type 1 and cell.type 2. If we encounter a cell.type that is neither 1 nor 2 (which we should not), we print an error message. Once we have created a steppable (*i.e.*, created an object of class TypeSwitcherSteppable) we must register it using registerSteppable function from steppableRegistry object: typeSwitcherSteppable=TypeSwitcherSteppable(sim,100); steppableRegistry.registerSteppable(typeSwitcherSteppable) CompuCell3D will not run unregistered steppables. As we will see, much of the script is not specific to this example. We will recycle it with slight changes in later examples. 12 shows snapshots of the cell-type-oscillator simulation. **Figure .****Results of the Python cell-type-oscillator simulation using the TypeSwitcherSteppable steppable implemented in *8* in conjunction with the CC3DML cell-sorting simulation in *3*. Cells exchange types and corresponding adhesivities and colors every 100 MCS; *i.e.*, between *t*=90 MCS and *t*=110 MCS and between *t*=1490 MCS and *t*=1510 MCS. We mentioned earlier that users can run simulations without a CC3DML configuration file. 9 shows the cell-type-oscillator simulation written entirely in Python, with changes to 8 shown in **bold**. **def configureSimulation(sim):** import CompuCell import CompuCellSetup ppd=CompuCell.PottsParseData() ppd.Steps(20000) ppd.Temperature(5) ppd.NeighborOrder(2) ppd.Dimensions(CompuCell.Dim3D(100,100,1)) ctpd=CompuCell.CellTypeParseData() ctpd.CellType("Medium",0) ctpd.CellType("Condensing",1) ctpd.CellType("NonCondensing",2) **cpd=CompuCell.ContactParseData()** cpd.Energy("Medium","Medium",0) cpd.Energy("NonCondensing","NonCondensing",16) cpd.Energy("Condensing","Condensing",2) cpd.Energy("NonCondensing","Condensing",11) cpd.Energy("NonCondensing","Medium",16) cpd.Energy("Condensing","Medium",16) vpd=CompuCell.VolumeParseData() vpd.LambdaVolume(1.0) vpd.TargetVolume(25.0) **bipd=CompuCell.BlobInitializerParseData()** region=bipd.Region() region.Center(CompuCell.Point3D(50,50,0)) **region.Radius(40)** region.Types("Condensing") region.Types("NonCondensing") region.Width(5) CompuCellSetup.registerPotts(sim,ppd) CompuCellSetup.registerPlugin(sim,ctpd) CompuCellSetup.registerPlugin(sim,cpd) **CompuCellSetup.registerPlugin(sim,vpd)** CompuCellSetup.registerSteppable(sim,bipd) import sys from os import environ from os import getcwd import string sys.path.append(environ["PYTHON\_MODULE\_PATH"]) import CompuCellSetup sim,simthread = CompuCellSetup.getCoreSimulationObjects() **configureSimulation(sim)** from PySteppables import \* class TypeSwitcherSteppable(SteppablePy): def \_\_init\_\_(self,\_simulator,\_frequency=100): SteppablePy.\_\_init\_\_(self,\_frequency) self.simulator=\_simulator self.inventory=self.simulator.getPotts().getCellInventory() self.cellList=CellList(self.inventory) def step(self,mcs): for cell in self.cellList: if cell.type==1: cell.type=2 elif (cell.type==2): cell.type=1 else: print "Unknown type. In cellsort simulation there should only be two types 1 and 2" #Create extra player fields here or add attributes CompuCellSetup.initializeSimulationObjects(sim,simthread) #Add Python steppables here steppableRegistry=CompuCellSetup.getSteppableRegistry() typeSwitcherSteppable=TypeSwitcherSteppable(sim,100); steppableRegistry.registerSteppable(typeSwitcherSteppable) CompuCellSetup.mainLoop(sim,simthread,steppableRegistry) CompuCellSetup.mainLoop(sim,simthread,steppableRegistry) **Listing .** Stand-alone Python cell-type-oscillator script containing an initial section that replaces the CC3DML configuration file from 3. Lines added to 8 are shown in **bold**. The configureSimulation function replaces the CC3DML file from 3. After importing CompuCell and CompuCellSetup, we have access to functions and modules that provide all the functionality necessary to code a simulation in Python. The general syntax for the opening lines of each block is: *snpd*=CompuCell.*SectionName*ParseData() where *SectionName* refers to the name of the section in a CC3DML configuration file and *snpd* is the name of the object of type *SectionName*ParseData. The rest of the block usually follows the syntax: *snpd*.*TagName*(*values*) where *TagName* corresponds to the name of the tag pair used to assign a value to a parameter in a CC3DML file. For values within subsections, the syntax is: *snpd*.*SubsectionName*().*TagName*(*values*) To input dimensions, we use the syntax: *snpd*.*TagName*(CompuCell.Dim3D(*x\_dim*,*y\_dim*,*z\_dim*)) where *x\_dim*, *y\_dim*, and *z\_dim* are the *x*, *y* and *z* dimensions. To input a set of (*x*,*y*,*z*) coordinates, we use the syntax: *snpd*.TagName(CompuCell.Point3D(*x\_coord*,*y\_coord*,*z\_coord*)) where *x\_coord*,*y\_coord*, and *z\_coord* are the *x*, *y*, and *z* coordinates. In the first block (PottsParseData), we input the cell-lattice parameter values using the syntax: ppd.*ParameterName*(*value*) where *ParameterName* matches a parameter name used in the CC3DML lattice section. Next we define the cell types using the syntax: ctpd.CellType("*cell\_type*",*cell\_id*) The next section assigns boundary energies between the cell types: cpd.Energy("cell\_type\_1","cell\_type\_2",contact\_energy) We specify the rest of the parameter values in a similar fashion, following the general syntax described above. The examples in 8 and 9 define the TypeSwitcherSteppable class within the main Python script. However, separating extension modules from the main script and using an import statement to refer to modules stored in external files is more practical. Using separate files ensures that each module can be used in multiple simulations without duplicating source code, and makes scripts more readable and editable. We will follow this convention in our remaining examples. **VI.D Two-Dimensional Foam-Flow Simulation** CompuCell3D can simulate simple physical experiments with foams. Indeed, GGH techniques grew out of foam-simulation techniques *(****73****)*. Our next example shows how to use CC3DML and Python scripts to simulate quasi-two-dimensional foam flow. The experimental apparatus (13) consists of a channel created by two parallel rectangular glass plates separated by 5 mm, with the gap between their long sides sealed and that between their short sides open. A foam generator injects small, uniform size bubbles at one short end, pushing older bubbles towards the open end of the channel, creating a foam flow. The top glass plate has a hole through which we inject air. Bubbles passing under this point grow because of the air injected into them, forming characteristic patterns (14) *(****99****)*. **Figure .****Schematic of experiment for studying quasi-2D foam flow. **Figure .****Detail of processed experimental image of flowing quasi-2D bubbles. Image size is 15 cm x 15 cm. Generalized cells will represent bubbles in this simulation. To simulate this experiment in CompuCell3D we need to write Python steppables that 1) create bubbles at one end of the channel, 2) inject air into the bubble which includes a given location (the identity of this bubble will change in time due to the flow), 3) remove bubbles at the open end of the channel. We will store the source code in a file called 'foamairSteppables.py'. We will also need a main Python script to call these steppables appropriately. We simulate bubble injection by creating generalized cells (bubbles) along the lattice edge corresponding to the left end of the channel (small-*x* values of the cell lattice). We simulate air injection into a bubble at the injection point, by identifying the bubble currently at the injection point and increasing its target volume at a fixed rate. Removing a bubble from the simulation simply requires assigning it a target volume of zero once it comes close to the right end of the channel (large-*x* values of the cell lattice). We first define a CC3DML configuration file for the foam-flow simulation (10). 10000 5 Hexagonal 5 5 5 3 **Listing .** CC3DML configuration file for the foam-flow simulation. This file initializes needed plugins but all of the interesting work is done in Python. The CC3DML configuration file is simple: it initializes the VolumeLocalFlex, CellType, Contact and CenterOfMass plugins. We do not use a cell-lattice-initializer steppable, because all bubbles are created as the simulation runs. We use VolumeLocalFlex because individual bubbles will change their target volumes during the simulation. We also include the CenterOfMass plugin to track the changing centroids of each bubble. The CenterOfMass plugin in CompuCell3D actually calculates , the centroid of the generalized cell multiplied by volume of the cell: , \\\* MERGEFORMAT () so the actual centroid of the bubble is: . \\\* MERGEFORMAT () The ability to track a generalized-cell’s centroid is useful if we need to pick a single reference point in the cell. In this example we will remove bubbles whose centroids have *x*-coordinate greater than a cutoff value. We will implement the Python script in four sections: 1) a main script (11), which runs every MCS and calls the steppables to (2) create bubbles at the left end of the cell lattice (BubbleNucleator, 12), (3) enlarge the target volume of the bubble at the injector site (AirInjector, 13), and (4) set the target volume of bubbles at the right end of the cell lattice to zero (BubbleCellRemover, 14). We store classes (2-4) in a separate file called 'foamairSteppables.py'. import sys from os import environ import string sys.path.append(environ["PYTHON\_MODULE\_PATH"]) import CompuCellSetup sim,simthread = CompuCellSetup.getCoreSimulationObjects() #Create extra player fields here CompuCellSetup.initializeSimulationObjects(sim,simthread) #Add Python steppables here steppableRegistry=CompuCellSetup.getSteppableRegistry() **from foamairSteppables import BubbleNucleator** **bubbleNucleator=BubbleNucleator(sim,20)** **bubbleNucleator.setNumberOfNewBubbles(1)** **bubbleNucleator.setInitialTargetVolume(25)** **bubbleNucleator.setInitialLambdaVolume(2.0)** **bubbleNucleator.setInitialCellType(1)** **steppableRegistry.registerSteppable(bubbleNucleator)** **** **from foamairSteppables import AirInjector** **airInjector=AirInjector(sim,40)** **airInjector.setVolumeIncrement(25)** **airInjector.setInjectionPoint(50,25,0)** **steppableRegistry.registerSteppable(airInjector)** **** **from foamairSteppables import BubbleCellRemover** **bubbleCellRemover=BubbleCellRemover(sim)** **bubbleCellRemover.setCutoffValue(170)** **steppableRegistry.registerSteppable(bubbleCellRemover)** **** CompuCellSetup.mainLoop(sim,simthread,steppableRegistry) **Listing .** Main Python Script for foam-flow simulation. Changes to the template (7) are shown in **bold**. The main script in 11 builds on the template Python code in 7; we show changes in **bold**. The line: from foamairSteppables import BubbleNucleator tells Python to look for the BubbleNucleator class in the file named 'foamairSteppables.py'. bubbleNucleator=BubbleNucleator(sim, 20) creates the steppable BubbleNucleator that will run every 20 MCS. The next few lines in this section pass the number of bubbles to create, which in our case is one: bubbleNucleator.setNumberOfNewBubbles(1) the initial for the new bubble, which is 25 pixels: bubbleNucleator.setInitialTargetVolume(25) the initial for the bubble: bubbleNucleator.setInitialLambdaVolume(2.0) and the bubble’s type.id: bubbleNucleator.setInitialCellType(1) Finally, we register the steppable: steppableRegistry.registerSteppable(bubbleNucleator) The next group of lines repeats the process for the AirInjector steppable, reading it from the file 'foamairSteppables.py': from foamairSteppables import AirInjector AirInjector is run every 40 MCS: airInjector=AirInjector(sim, 40) and increases by 25: airInjector.setVolumeIncrement(25) for the bubble occupying the pixel at the point (50, 25, 0) on the cell lattice: airInjector.setInjectionPoint(50,25,0) As before, the final line registers the steppable: steppableRegistry.registerSteppable(airInjector) **** The final new section reads the BubbleCellRemover steppable from the file 'foamairSteppables.py': from foamairSteppables import BubbleCellRemover and invokes the steppable, telling it to run every MCS; note that we have omitted the number after sim: bubbleCellRemover=BubbleCellRemover(sim) Next we set 170 as the *x*-coordinate at which we will destroy bubbles: bubbleCellRemover.setCutoffValue(170) and, finally, register BubbleCellRemover steppableRegistry.registerSteppable(bubbleCellRemover) We must also write Python code to define the three steppables BubbleNucleator, AirInjector, and BubbleCellRemover and save them in the file 'foamairSteppables.py'. 12 shows the code for the BubbleNucleator steppable. from CompuCell import Point3D from random import randint class BubbleNucleator(SteppablePy): def \_\_init\_\_(self,\_simulator,\_frequency=1): SteppablePy.\_\_init\_\_(self,\_frequency) self.simulator=\_simulator def start(self): self.Potts=self.simulator.getPotts() self.dim=self.Potts.getCellFieldG().getDim() def setNumberOfNewBubbles(self,\_numNewBubbles): self.numNewBubbles=int(\_numNewBubbles) def setInitialTargetVolume(self,\_initTargetVolume): self.initTargetVolume=\_initTargetVolume def setInitialLambdaVolume(self,\_initLambdaVolume): self.initLambdaVolume=\_initLambdaVolume def setInitialCellType(self,\_initCellType): self.initCellType=\_initCellType def createNewCell(self,pt): print "Nucleated bubble at ",pt cell=self.Potts.createCellG(pt) cell.targetVolume=self.initTargetVolume cell.type=self.initCellType cell.lambdaVolume=self.initLambdaVolume def nucleateBubble(self): pt=Point3D(0,0,0) pt.y=randint(0,self.dim.y-1) pt.x=3 self.createNewCell(pt) def step(self,mcs): for i in xrange(self.numNewBubbles): self.nucleateBubble() **Listing .** Python code for the BubbleNucleator steppable, saved in the file 'foamairSteppables.py'. This module creates bubbles at points with random *y* coordinates and *x* coordinate of 3. The first two lines import necessary modules, where the line: from CompuCell import Point3D allows us to access points on the simulation cell lattice, and the line: from random import randint allows us to generate random integers. In the constructor of the BubbleNucleator steppable class we assign to the variable self.simulator a reference to the simulator object from the CompuCell3D kernel. In the start(self) function, we assign a reference to the Potts object from the CompuCell3D kernel to the variable self.Potts: self.Potts=self.simulator.getPotts() and assign the dimensions of the cell lattice to self.dim: self.dim=self.Potts.getCellFieldG().getDim() In addition to the four essential steppable member functions (\_\_init\_\_(self, \_simulator, \_frequency), start(self), step(self, mcs) and finish(self)), BubbleNucleator includes several functions, some of which set parameters and some of which perform necessary tasks. The functions setNumberOfNewBubbles, setInitialTargetVolume and setInitialLambdaVolume accept the values passed from the main Python script in 11. The CreateNewCell function requires that we pass the coordinates of the point, pt, at which to create a new bubble: def CreateNewCell (self,pt): Then we use a built-in CompuCell3D function to add a new bubble at that location: cell=self.Potts.createCellG(pt) assigning the new cell a target volume : cell.targetVolume=self.initTargetVolume type, : cell.type=self.initCellType and compressibility : cell.lambdaVolume=initLambdaVolume based on the values passed to the BubbleNucleator steppable from the main script. The first three lines of the nucleateBubble function create a reference to a point on the cell lattice (pt=Point3D(0,0,0)), assign it a random *y*-coordinate between 0 and y\_dim-1: pt.y=randint(0,self.dim.y-1) and an *x*-coordinate of 3: pt.x=3 The line calls the createNewCell function and passes it the point (pt) at which to create the new bubble: self.createNewCell(pt) Finally, the step(self,mcs) function calls the nucleateBubble function self.numNewBubbles times per MCS. 13 shows the code for the AirInjector steppable. class AirInjector(SteppablePy): def \_\_init\_\_(self,\_simulator,\_frequency=1): SteppablePy.\_\_init\_\_(self,\_frequency) self.simulator=\_simulator self.Potts=self.simulator.getPotts() self.cellField=self.Potts.getCellFieldG() def start(self): pass def setInjectionPoint(self,\_x,\_y,\_z): self.injectionPoint=CompuCell.Point3D(int(\_x),int(\_y),int(\_z)) def setVolumeIncrement(self,\_increment): self.volumeIncrement=\_increment def step(self,mcs): if mcs <5000: return cell=self.cellField.get(self.injectionPoint) if cell: cell.targetVolume+=self.volumeIncrement **Listing .** Python code for the *AirInjector* steppable which simulates air injection into the bubble currently occupying the cell-lattice pixel at location (x,y,z). Air injection begins after 5000 MCS to allow the channel to partially fill with bubbles. The steppable is saved in file 'foamairSteppables.py'. The first three lines of the \_\_init\_\_(self,\_simulator,\_frequency) function are identical to the same lines in the BubbleNucleator steppable (12). The final line of the function: self.cellField=self.Potts.getCellFieldG() loads the cell-lattice parameters. The start(self) function in this steppable does not do anything: def start(self): pass The next two functions read the injectionPoint and volumeIncrement passed to the AirInjector steppable by the main Python script (11). The step function uses these values to identify the bubble at the injection site, self.injectionPoint: cell=self.cellField.get(self.injectionPoint) and then increment that bubble’s target volume by self.volumeIncrement: if cell: cell.targetVolume+=self.volumeIncrement Note the syntax: if cell: which we use to test whether a cell is Medium or not. Medium in CompuCell3D is assigned a NULL pointer, which, in Python, becomes a None object. Python evaluates the None object as False and other objects (in our case, bubbles) as True, so the task is only carried out on bubbles, not Medium. In the first two lines of the step(self,mcs) function, we tell the function not to perform its task until 5000 MCS have elapsed: if mcs <5000: return The 5000 MCS delay allows the simulation to establish a uniform flow of small bubbles throughout a large portion of the cell lattice. Finally, we define the BubbleCellRemover steppable (14). class BubbleCellRemover(SteppablePy): def \_\_init\_\_(self,\_simulator,\_frequency=1): SteppablePy.\_\_init\_\_(self,\_frequency) self.simulator=\_simulator self.inventory=self.simulator.getPotts().getCellInventory() self.cellList=CellList(self.inventory) def start(self): self.Potts=self.simulator.getPotts() self.dim=self.Potts.getCellFieldG().getDim() def setCutoffValue(self,\_cutoffValue): self.cutoffValue=\_cutoffValue def step(self,mcs): for cell in self.cellList: if cell: if int(cell.xCM/float(cell.volume))>self.cutoffValue: cell.targetVolume=0 cell.lambdaVolume=10000 **Listing .** Python code for the *BubbleCellRemover* steppable. This module removes cells once the x-coordinates of their centroids cutoffValue by setting their target volumes to zero and increasing their to 10000. Like the other steppables in the foam-flow simulation, we save it in the file 'foamairSteppables.py'. At each MCS we scan the cell inventory looking for cells whose centroid has an *x*-coordinate close to the right end of the lattice and remove these cells from the simulation by setting their target volumes to zero and increasing to 10000. The first two lines of the \_\_init\_\_ (self,\_simulator,\_frequency) function are identical to the corresponding lines in the BubbleNucleator and AirInjector steppables (12 and 13). In the third line of the function, we gain access to the generalized-cell inventory: self.inventory=self.simulator.getPotts().getCellInventory() and in the fourth line we make a list containing all of the generalized cells in the simulation: self.cellList=CellList(self.inventory) The start(self) function is identical to that of the BubbleNucleator steppable (12), and performs the same function. The next function: setCutoffValue(self,\_cutoffValue) reads the cutoffValue for the *x*-coordinate that we passed to BubbleCellRemover from the main Python script (11). Finally, the step(self, mcs) function iterates through the cell inventory. We first check to make sure that the cell is not Medium: if cell: For each non-Medium cell we test whether the *x*-coordinate of the cell’s centroid is greater than the cutoffValue: if int(cell.xCM/float(cell.volume))>self.cutoffValue: and, if it is, set that cell’s targetVolume,, to zero: cell.targetVolume=0 and its : cell.lambdaVolume=10000 Running the CC3DML file from 10 and the main Python script from 11 (which loads the steppables in 12, 13 and 14 from the file 'foamairSteppables.py') produces the snapshots shown in 15. **Figure .****Results of the foam-flow simulation on a 2D 3rd-neighbor hexagonal lattice. Simulation code is given in *10*,*11*, *12*, *13* and *14*. **VI.E. Diffusing-Field-Based Cell-Growth Simulation** One of the most frequent uses of Python scripting in CompuCell3D simulations is to modify cell behavior based on local field concentrations. To demonstrate this use, we incorporate stem-cell-like behavior into the cell-sorting simulation from 1. This extension requires including relatively sophisticated interactions between cells and diffusing chemical, *FGF* *(****100****)*. We simulate a situation where NonCondensing cells secrete FGF, which diffuses freely through the cell lattice and obeys: , \\\* MERGEFORMAT () where denotes the FGF concentration and Condensing cells respond to the field by growing at a constant rate proportional to the FGF concentration at their centroids: . \\\* MERGEFORMAT () When they reach a threshold volume, the Condensing cells undergo mitosis. One of the resulting daughter cells remains a Condensing cell, while the other daughter cell has an equal probability of becoming either another Condensing cell or a DifferentiatedCondensing**cell. DifferentiatedCondensing**cells do not divide. Each generalized cell in CompuCell3D has a default list of attributes, *e.g.* type, volume, surface (area), target volume, *etc.*. However, CompuCell3D allows users to add cell attributes during execution of simulations. *E.g.*, in the current simulation, we will record data on each cell division in a list attached to each cell. Generalized cell attributes can be added using either C++ or Python. However, attributes added using Python are not accessible from C++ modules. As in the foam-flow simulation, we divide the necessary simulation tasks among different Python modules (or classes) which we save in a file 'cellsort\_2D\_field\_modules.py' and call from the main Python script. We reuse elements of the CC3DML files we presented earlier to construct the CC3DML configuration file, presented in 15. 10000 10 2 0 16 2 11 16 16 2 2 11 16 2 FGF 0.10 0.00005 0.05 0 5 40
Condensing,NonCondensing **Listing .** CC3DML code for the diffusing-field-based cell-growth simulation. The CC3DML code is a slightly extended version of the cell-sorting code in 3 plus the FlexibleDiffusionSolverFE discussed in the bacterium-and-macrophage simulation (see 4). The initial cell-lattice does not contain any CondensingDifferentiated cells. These cells appear only as the result of mitosis. We use the VolumeLocalFlex plugin to allow the target volume to vary individually for each cell, allowing cell growth as discussed in the foam-flow simulation. We manage the volume-constraint parameters using a Python script. The CenterOfMass plugin provides a reference point in each cell at which we measure the FGF concentration. We then adjust the cell's target volume accordingly. To build this simulation in CompuCell3D we need to write several Python routines. We need: 1) A steppable, VolumeConstraintSteppable to initialize the volume-constraint parameters for each cell and to simulate cell growth by periodically increasing Condensing cells’ target volumes in proportion to the FGF concentration at their centroids. 2) A plugin, CellsortMitosis, that runs the CompuCell3D mitosis algorithm when any cell reaches a threshold volume and then adjusts the parameters of the resulting parent and daughter cells. This plugin also appends information about the time and type of cell division to a list attached to each cell. 3) A steppable, MitosisDataPrinterSteppable, that prints the cell-division information from the lists attached to each cell. 4) A class, MitosisData, which MitosisDataPrinterSteppable uses to extract and format the data it prints. 5) A main Python script to call the steppables and the CellsortMitosis plugin appropriately. We store the source code for routines 1)-4) in a separate file called 'cellsort\_2D\_field\_modules.py'. 16 shows the main Python script for the diffusing-field-based cell-growth simulation, with changes to the template (7) shown in **bold**. import sys from os import environ from os import getcwd import string sys.path.append(environ["PYTHON\_MODULE\_PATH"]) import CompuCellSetup sim,simthread = CompuCellSetup.getCoreSimulationObjects() #add additional attributes **pyAttributeAdder,listAdder=CompuCellSetup.attachListToCells(sim**) CompuCellSetup.initializeSimulationObjects(sim,simthread) #notice importing CompuCell to main script has to be #done after call to getCoreSimulationObjects() **import CompuCell changeWatcherRegistry=CompuCellSetup.getChangeWatcherRegistry(sim)** **stepperRegistry=CompuCellSetup.getStepperRegistry(sim)** **** **from cellsort\_2D\_field\_modules import CellsortMitosis** **cellsortMitosis=CellsortMitosis(sim,changeWatcherRegistry,\\** **stepperRegistry)** **cellsortMitosis.setDoublingVolume(50)** #Add Python steppables here steppableRegistry=CompuCellSetup.getSteppableRegistry() **from cellsort\_2D\_field\_modules import VolumeConstraintSteppable** **volumeConstraint=VolumeConstraintSteppable(sim)** **steppableRegistry.registerSteppable(volumeConstraint)** **** **from cellsort\_2D\_field\_modules import MitosisDataPrinterSteppable** **mitosisDataPrinterSteppable=MitosisDataPrinterSteppable(sim)** **steppableRegistry.registerSteppable(mitosisDataPrinterSteppable)** CompuCellSetup.mainLoop(sim,simthread,steppableRegistry) **Listing .** Main Python script for the diffusing-field-based cell-growth simulation. Changes to the template code (7) shown in **bold.** The first change to the template code (7) is: pyAttributeAdder,listAdder=CompuCellSetup.attachListToCells(sim) which instructs the CompuCell3D kernel to attach a Python-defined list to each cell when it creates it. This list serves as a generic container which can store any set of Python objects and hence any set of generalized-cell properties. In the current simulation, we use the list to store objects of the class MitosisData, which records the Monte Carlo Step at which each cell division involving the current cell or its parent, happened, as well as, the cell index and cell type of the parent and daughter cells. Because one of our Python modules is a lattice monitor, rather than a steppable, we need to create stepperRegistry and changeWatcherRegistry objects, which store the two types of lattice monitors: changeWatcherRegistry=CompuCellSetup.getChangeWatcherRegistry(sim) stepperRegistry=CompuCellSetup.getStepperRegistry(sim) The CellsortMitosis plugin is a lattice monitor because it acts in response to certain index-copy events; it is invoked whenever a cell's volume reaches the threshold volume for mitosis. The following lines create the CellsortMitosis lattice monitor and register it with the stepperRegistry and changeWatcherRegistry: from cellsort\_2D\_field\_modules import CellsortMitosis cellsortMitosis = CellsortMitosis(sim,changeWatcherRegistry,\\ stepperRegistry) Because the base class inherited by CellsortMitosis, unlike our steppables, handles registration internally, we do not have to register CellsortMitosis explicitly. We can now set the threshold volume at which Condensing cells divide: cellsortMitosis.setDoublingVolume(50) Next we import the VolumeConstraintSteppable steppable, which initializes cells’ target volumes and compressibilities at the beginning of the simulation and also implements chemical-dependent cell growth for Condensing cells, and register it: from cellsort\_2D\_field\_modules import VolumeConstraintSteppable volumeConstraint=VolumeConstraintSteppable(sim) steppableRegistry.registerSteppable(volumeConstraint) Finally, we import, create and register the MitosisDataPrinterSteppable steppable, which prints the content of MitosisData objects for cells that have divided: from cellsort\_2D\_field\_modules import MitosisDataPrinterSteppable mitosisDataPrinterSteppable=MitosisDataPrinterSteppable(sim) steppableRegistry.registerSteppable(mitosisDataPrinterSteppable) The number of MitosisData objects stored in each cell at any given Monte Carlo Step depends on cell type (NonCondensing cells do not divide, whereas Condensing cells can divide multiple times), and how often a given cell has divided. Moving on to the Python modules, we consider the VolumeConstraintSteppable steppable shown in 17. class VolumeConstraintSteppable(SteppablePy): def \_\_init\_\_(self,\_simulator,\_frequency=1): SteppablePy.\_\_init\_\_(self,\_frequency) self.simulator=\_simulator self.inventory=self.simulator.getPotts().getCellInventory() self.cellList=CellList(self.inventory) def start(self): for cell in self.cellList: cell.targetVolume=25 cell.lambdaVolume=2.0 def step(self,mcs): field=CompuCell.getConcentrationField(self.simulator,"FGF") comPt=CompuCell.Point3D() for cell in self.cellList: if cell.type==1: #Condensing cell comPt.x=int(round(cell.xCM/float(cell.volume))) comPt.y=int(round(cell.yCM/float(cell.volume))) comPt.z=int(round(cell.zCM/float(cell.volume))) concentration=field.get(comPt) # get concentration at comPt # and increase cell's target volume cell.targetVolume+=0.1\*concentration **Listing .** Python code for the VolumeConstraintSteppable, saved in the file 'cellsort\_2D\_field\_modules.py', for the diffusing-field-based cell-growth simulation. The VolumeConstraintSteppable*provides* dynamic volume constraint parameters for each cell, which depend on the cell type and the chemical field concentration at the cell’s centroid. The \_\_init\_\_ constructor looks very similar to the one in 14, with the difference that we pass \_frequency=1 to update the cell volumes once per MCS. We also request the field-lattice dimensions and values from CompuCell3D: self.dim=self.simulator.getPotts().getCellFieldG().getDim() and specify that we will work with a field named FGF: self.fieldName="FGF" The script contains two functions: one that initializes the cells’ volume-constraint parameters (start(self)) and one that updates them (step(self, mcs)). The start(self) function executes only once, at the beginning of the simulation. It iterates over each cell (for cell in self.cellList:) and assigns the initial cells’ targetVolume ( pixels) and lambdaVolume () parameters as the VolumeLocalFlex plugin requires. The first line of the step(self, mcs) function extracts a reference to the FGF concentration field defined using the FlexibleDiffusionSolverFE steppable in the CC3DML file (each field created in a CompuCell3D simulation is registered and accessible by both C++ and Python). The function then iterates over every cell in the simulation. If a cell is of cell.type 1 (Condensing – see the CC3DML configuration file, 15), we calculate its centroid: centerOfMassPoint.x=int(round(cell.xCM/float(cell.volume))) centerOfMassPoint.y=int(round(cell.yCM/float(cell.volume))) centerOfMassPoint.z=int(round(cell.zCM/float(cell.volume))) and retrieve the FGF concentration at that point: concentration=field.get(centerOfMassPoint) We then increase the target volume of the cell by 0.01 times that concentration: cell.targetVolume+=0.01\*concentration We must include the CenterOfMass plugin in the CC3DML code. Otherwise the centroid (cell.xCM, cell.yCM, cell.zCM) will have the default value (0,0,0). 18 shows the code for the CellsortMitosis plugin. The plugin divides the mitotic cell into two cells and adjusts both cells' attributes. It also initializes and appends MitosisData objects to the original cell's (self.parentCell) and daughter cell's (self.childCell) attribute lists. from random import random from PyPluginsExamples import MitosisPyPluginBase class CellsortMitosis(MitosisPyPluginBase): def \_\_init\_\_(self,\_simulator,\_changeWatcherRegistry,\\ \_stepperRegistry): MitosisPyPluginBase.\_\_init\_\_(self,\_simulator,\\ \_changeWatcherRegistry,\_stepperRegistry) def updateAttributes(self): self.parentCell.targetVolume=self.parentCell.volume/2.0 self.childCell.targetVolume=self.parentCell.targetVolume self.childCell.lambdaVolume=self.parentCell.lambdaVolume if (random()<0.5): self.childCell.type=self.parentCell.type else: self.childCell.type=3 ##record mitosis data in parent and daughter cells mcs=self.simulator.getStep() mitData=MitosisData(mcs,self.parentCell.id,self.parentCell.type,\\ self.childCell.id,self.childCell.type) #get a reference to lists storing Mitosis data parentCellList=CompuCell.getPyAttrib(self.parentCell) childCellList=CompuCell.getPyAttrib(self.childCell) parentCellList.append(mitData) childCellList.append(mitData) **Listing .** Python code for the CellsortMitosis plugin for the diffusing-field-based cell-growth simulation, saved in the file 'cellsort\_2D\_field\_modules.py'. The plugin handles division of cells when they reach a threshold volume. The second line of 18: from PyPluginsExamples import MitosisPyPluginBase lets us access the CompuCell3D base class MitosisPyPluginBase. CellsortMitosis inherits the content of the MitosisPyPluginBase class. MitosisPyPluginBase internally accesses the CompuCell3D-provided Mitosis plugin, which is written in C++, and handles all the technicalities of plugin initialization behind the scenes. The MitosisPyPluginBase class provides a simple-to-use interface to this plugin. To create a customized version of MitosisPyPluginBase, CellsortMitosis, we must call the constructor of MitosisPyPluginBase from the CellsortMitosis constructor: MitosisPyPluginBase.\_\_init\_\_(self,\_simulator,\\ \_changeWatcherRegistry,\_stepperRegistry) We also need to reimplement the function updateAttributes(self), which is called by MitosisPyPluginBase after mitosis takes place, to define the post-division cells’ parameters. The objects self.childCell and self.parentCell that appear in the function are initialized and managed by MitosisPyPluginBase. In the current simulation, after division we set for the parent and daughter cells to half of the of the parent just prior to cell division. is left unchanged for the parent cell and the same value is assigned to the daughter cell: self.parentCell.targetVolume=self.parentCell.volume/2.0 self.childCell.targetVolume=self.parentCell.targetVolume self.childCell.lambdaVolume=self.parentCell.lambdaVolume The cell type of one of the two daughter cells (childCell) is randomly chosen to be either Condensing (*i.e.*, the same as the parent type) or CondensingDifferentiated, which we have defined to be cell.type 3 (15): if (random()<0.5): self.childCell.type=self.parentCell.type else: self.childCell.type=3 The parent cell remains Condensing. We now add a description of this cell division to the lists attached to each cell. First we collect the data in a list called mitData: mcs=self.simulator.getStep() mitData=MitosisData(mcs,self.parentCell.id,self.parentCell.type,\\ self.childCell.id,self.childCell.type) then we access the lists attached to the two cells: parentCellList=CompuCell.getPyAttrib(self.parentCell) childCellList=CompuCell.getPyAttrib(self.childCell) and append the new mitosis data to these lists: parentCellList.append(mitData) childCellList.append(mitData) 19 shows the Python code for the MitosisData class, which stores the data on the cell division that we append to the cells’ attribute lists after each cell division. class MitosisData: def \_\_init\_\_(self,\_MCS,\_parentId,\_parentType,\_offspringId,\\ \_offspringType): self.MCS=\_MCS self.parentId=\_parentId self.parentType=\_parentType self.offspringId=\_offspringId self.offspringType=\_offspringType def \_\_str\_\_(self): return "Mitosis time="+str(self.MCS)+"\\ parentId="+str(self.parentId)+"\\ offspringId="+str(self.offspringId) **Listing .**Python code for the MitosisData class for the diffusing-field-based cell-growth simulation, saved in the file 'cellsort\_2D\_field\_modules.py'. MitosisData objects store information about cell divisions involving the parent and daughter cells. In the constructor of MitosisData, we read in the time (in MCS) of the division, along with the parent and daughter cell indices and types. The \_\_str\_\_(self) convenience function returns an ASCII string representation of the time and cell indices only, to allow the Python print command to print out this information. 20 shows the Python code for the MitosisDataPrinterSteppable steppable, which prints the mitosis data to the user's screen. class MitosisDataPrinterSteppable(SteppablePy): def \_\_init\_\_(self,\_simulator,\_frequency=100): SteppablePy.\_\_init\_\_(self,\_frequency) self.simulator=\_simulator self.inventory=self.simulator.getPotts().getCellInventory() self.cellList=CellList(self.inventory) def step(self,mcs): for cell in self.cellList: mitDataList=CompuCell.getPyAttrib(cell) if len(mitDataList) > 0: print "MITOSIS DATA FOR CELL ID",cell.id for mitData in mitDataList: print mitData **Listing .** The Python code for the MitosisDataPrinter steppable for the diffusing-field-based cell-growth simulation, saved in the file 'cellsort\_2D\_field\_modules.py'. The steppable prints the cell-division history for dividing cells (see 18). The constructor is identical to that for the VolumeConstraintSteppable steppable (17). Within the step(self,mcs) function, we iterate over each cell (for cell in self.cellList:) and access the Python list attached to the cell (mitDataList=CompuCell.getPyAttrib(cell)). If a given cell has undergone mitosis, then the list will have entries, and thus a nonzero length. If so, we print the MitosisData objects stored in the list: if len(mitDataList) > 0: print "MITOSIS DATA FOR CELL ID",cell.id for mitData in mitDataList: print mitData 16 and 17 show snapshots of the diffusing-field-based cell-growth simulation. Figure 18 shows a sample screen output of the cell-division history. **Figure .****Snapshots of the diffusing-field-based cell-growth simulation obtained by running the CC3DML file in *15* in conjunction with the Python file in *16*. As the simulation progresses, NonCondensing cells (light gray) secrete diffusing chemical, FGF, which causes Condensing (dark gray) cells to proliferate. Some Condensing cells differentiate to CondensingDifferentiated (white) cells. **Figure .****Snapshots of FGF concentration in the diffusing-field-based cell-growth simulation obtained by running the CC3DML file in *15* in conjunction with the Python files in *16*, *17***,***18*, *19***,***20*. The bars at the bottom of the field images show the concentration scales (blue, low concentration; red, high concentration). ** **Figure .**Sample output from the MitosisDataPrinterSteppable steppable in *20*. The diffusing-field-based cell-growth simulation includes concepts that extend easily to simulate biological phenomena that involve diffusants, cell growth and mitosis, *e.g.*, limb-bud development *(****58****,****59****)*, tumor growth *(****5****-****9****)* and *Drosophila* imaginal-disc development. **VII. Conclusion** In most cases, building a complex CompuCell3D simulation requires writing Python modules, a main Python script and a CC3DML configuration file. While the effort to write this code can be substantial, it is much less than that required to develop custom simulations in lower-level languages. Working from the substantial base of Python templates provided by CompuCell3D further streamlines simulation development. Python programs are fairly short, so simulations can be published in journal articles, greatly facilitating simulation validation, reuse and adaptation. Finally, CompuCell3D’s modular structure allows new Python modules to be reused from simulation to simulation. The CompuCell3D website, `*www.compucell3d.org* `_*,*allows users to archive their modules and make them accessible to other users. We hope the examples we have shown will convince readers to evaluate the suitability of GGH simulations using CompuCell3D for their research. All the code examples presented in this chapter are available from `*www.compucell3d.org* `_. They will be curated to ensure their correctness and compatibility with future versions of CompuCell3D. **VIII. Acknowledgements** We gratefully acknowledge support from the National Institutes of Health, National Institute of General Medical Sciences, grants 1R01 GM077138-01A1 and 1R01 GM076692-01, and the Office of Vice President for Research, the College of Arts and Sciences, the Pervasive Technologies Laboratories and the Biocomplexity Institute at Indiana University. Indiana University’s University Information Technology Services provided time on their BigRed clusters for simulation execution. Early versions of CompuCell and CompuCell3D were developed at the University of Notre Dame by J.A.G., Dr. Mark Alber and Dr. Jesus Izaguirre and collaborators with the support of National Science Foundation, Division of Integrative Biology, grant IBN-00836563. Since the primary home of CompuCell3D moved to Indiana University in 2004, the Notre Dame team have continued to provide important support for its development. **IX. XML Syntax of CompuCell3D modules** **IX.1. Potts Section** The first section of the .xml file defines the global parameters of the lattice and the simulation. 0 1000 5 1 Periodic Periodic 2 20 167473 statData.txt statDataSingleFlip This section appears at the beginning of the configuration file. Line declares the dimensions of the lattice to be 101 x 101 x 1, *i.e.*, the lattice is two-dimensional and extends in the xy plane. The basis of the lattice is 0 in each direction, so the 101 lattice sites in the x and y directions have indices ranging from 0 to 100. 1000 tells CompuCell how long the simulation lasts in MCS. After executing this number of steps, CompuCell can run simulation at zero temperature for an additional period. In our case it will run for 10 extra steps. FluctuationAmplitude parameter determines intrinsic fluctuation or motility of cell membrane. Fluctuation amplitude is a temperature parameter in classical GGH model formulation. We have decided tyo use FluctuationAmplitude term instead of temperature because using word “temperature” to describe intrinsic motility of cell membrane was quite confusing. In the above example, fluctuation amplitude applies to all cells in the simulation. To define fluctuation amplitude separately for each cell type we use the following syntax: When CompuCell3D encounters expanded definition of FluctuationAmplitude it will use it in place of a global definition – 5 To complete the picture CompUCell3D allows users to set fluctuation amplitude individually for each cell. Using Python scripting we write: for cell in self.cellList: if cell.type==1: cell.fluctAmpl=20 When determining which value of fluctuation amplitude to use, CompuCell first checks if fluctAmpl is non-negative. If this is the case it will use this value as fluctuation amplitude. Otherwise it will check if users defined fluctuation amplitude for cell types using expanded XML definition and if so it will use those values as fluctuation amplitudes. Lastly it will resort to globally defined fluctuation amplitude (Temperature). Thus, it is perfectly fine to use FluctuationAmplitude XML tags and set fluctAmpl for certain cells. In such a case CompuCell3D will use fluctAmpl for cells for which users defined it and for all other cells it will use values defined in the XML. In GGH model, the fluctuation amplitude is determined taking into account fluctuation amplitude of “source” (expanding) cell and “destination” (being overwritten) cell. Currently CompuCell3D supports 3 type functions used to calculate resultant fluctuation amplitude (those functions take as argument fluctuation amplitude of “source” and “destination” cells and return fluctuation amplitude that is used in calculation of pixel-copy acceptance). The 3 functions are Min, Max, and ArithmeticAverage and we can set them using the following option of the Potts section: Min By default we use Min function. Notice that if you use global fluctuation amplitude definition (Temperature) it does not really matter which function you use. The differences arise when “source” and “destination” cells have different fluctuation amplitudes. The above concepts are best illustrated by the following example: Demos/FluctuationAmplitude/FluctuationAmplitude.py\\ 10000 5 ArithmeticAverage\\ 2 Where in the XML section we define global fluctuation amplitude and we also use ArithmeticAverage function to determine resultant fluctuation amplitude for the pixel copy. In python script we will periodically set higher fluctuation amplitude for lattice quadrants so that when running the simulation we can see that cells belonging to different lattice quadrants have different membrane fluctuations: class FluctuationAmplitude(SteppableBasePy): def \_\_init\_\_(self,\_simulator,\_frequency=1): SteppableBasePy.\_\_init\_\_(self,\_simulator,\_frequency) self.quarters=[[0,0,50,50],[0,50,50,100],\\ [50,50,100,100],[50,0,100,50]] self.steppableCallCounter=0 def step(self, mcs): quarterIndex=self.steppableCallCounter % 4 quarter=self.quarters[quarterIndex] for cell in self.cellList: if cell.xCOM>=quarter[0] and cell.yCOM>=quarter[1] and\\ cell.xCOM 0 1000 5 1 Periodic Periodic 2 20 167473 statData.txt statDataSingleFlip *This section appears at the beginning of the configuration file. Line****declares the dimensions of the lattice to be 101 x 101 x 1, i.e., the lattice is two-dimensional and extends in the xy plane. The basis of the lattice is 0 in each direction, so the 101 lattice sites in the x and y directions have indices ranging from 0 to 100.**1000**tells CompuCell how long the simulation lasts in MCS. After executing this number of steps, CompuCell can run simulation at zero temperature for an additional period. In our case it will run for**10**extra steps. Setting the temperature is as easyas writing**5**.* *We can also set temperature (or in other words cell motility) individually for each cell type. The syntax to do this is following:* *You may use it in the Potts section in place of****.* Now, as you remember from the discussion about the difference between spin-flip attempts and MCS we can specify how many spin flips should be attempted in every MCS. We specify this number indirectly by specifying the Flip2DimRatio - 1, which tells CompuCell that it should make 1 x number of lattice sites attempts per MCS – in our case one MCS is 101x101x1 spin-flip attempts. To set 2.5x101x101x1 spin flip attempts per MCS you would write 2.5. The next line specifies the neighbor order. The higher neighbor order the longer the Euclidian distance from a given pixel. In previous versions of CompuCell3D we have been using or (in Contact energy plugins) flag to accomplish same task. Since now CompuCell3D supports two kinds of latices it would be inconvenient to change distances. It is much easier to think in terms n-th nearest neighbors . For the backwards compatibility we still support old flags but we discourage its use, especially that in the future we might support more than just two lattice types. Using nearest neighbor interactions may cause artifacts due to lattice anisotropy. The longer the interaction range, the more isotropic the simulation and the slower it runs. In addition, if the interaction range is comparable to the cell size, you may generate unexpected effects, since non-adjacent cells will contact each other. On hex lattice those problems seem to be less seveare and there 1st or 2nd nearest neighbor usually are sufficient. The Potts section also contains tags called and .These tags impose boundary conditions on the lattice. In this case the x and y axes are **periodic** (Periodic) so that *e.g.* the pixel with x=0, y=1, z=1 will neighbor the pixel with x=100, y=1, z=1. If you do not specify boundary conditions CompuCell will assume them to be of type **no-flux**, *i.e.* lattice will not be extended. The conditions are independent in each direction, so you can specify any combination of boundary conditions you like. DebugOutputFrequency is used to tell CompuCell3D how often it should output text information about the status of the simulation. This tag is optional. RandomSeed is used to initialize random number generator. If you do not do this all simulations will use same sequence of random numbers. Something you may want to avoid in the real simulations but is very useful while debugging your models. EnergyFunctionCalculator is another option of Potts object that allows users to output statistical data from the simulation for further analysis. The OutputFileName tag is used to specify the name of the file to which CompuCell3D will write average changes in energies returned by each plugins with corresponding standard deviations for those MCS whose values are divisible by the Frequency argument. Here it will write these data every 10 MCS. A second line with OutputCoreFileNameSpinFlips tag is used to tell CompuCell3D to output energy change for every plugin, every spin flip for MCS' divisible by the frequency. Option GatherResults=”” will ensure that there is only one file written for accepted (OutputAccepted), rejected (OutputRejected)and accepted and rejected (OutputTotal) spin flips. If you will not specify GatherResults CompuCell3D will output separate files for different MCS's and depending on the Frequency you may end up with many files in your directory. One option of the Potts section that we have not used here is the ability to customize acceptance function for Metropolis algorithm: -0.1 1.2 This ensures that spin flips attempts that increase the energy of the system are accepted with probability where  and *k* are specified by Offset and KBoltzman tags respectively. By default =0 and *k=1*. As an alternative to exponential acceptance function you may use a simplified version which is essentially 1 order expansion of the exponential: To be able to use this function all you need to do is to add the following line in the Pots section: FirstOrderExpansion **IX.1.1 Lattice Type** Early versions of CompuCell3D allowed users to use only square lattice. Most recent versions however, allow the simulation to be run on hexagonal lattice as well. To enable hexagonal lattice you need to put Hexagonal in the Potts section of the XML configuration file. There are few things to be aware of. When using hexagonal lattice. Obviously your pixels are hexagons (2D) or rhombic dodecahedrons (3D) but what is more important is that surface or perimeter of the pixel (depending whether in 2D or 3D) is different than in the case of sqaure pixel. The way CompuCell3D hex lattice implementation was done was that the volume of the pixel was constrained to be 1 regardless of the lattice type. Second, there is one to one correspondence between pixels of the square lattice and pixels of the hex lattice. Consequently we can come up with transformation equations which give positions of hex pixels as a function of square lattice pixel position: Based on the above facts one can work out how unit length and unit surface transform to the hex lattice. The conversion factors are given below: For the 2D case, assuming that each pixel has unit volume, we get: where denotes length of the hexagon and denotes a distance between centers of the hexagons. Notice that unit surface in 2D is simply a length of the hexagon side and surface area of the hexagon with side 'a' is: In 3D we can derive the corresponding unit quantities starting with the formulae for Volume and surface of rhombic dodecahedron (12 hedra) where 'a' denotes length of dodecahedron edge. Constraining the volume to be one we get and thus unit surface is given by: and unit length by: **IX.2. Plugins Section** In this section we overview XML syntax for all the plugins available in CompuCell3D. Plugins are either energy functions, lattice monitors or store user assigned data that CompuCell3D uses internally to configure simulation before it is run. **IX.2.1. CellType Plugin** An example of the plugin that stores user assigned data that is used to configure simulation before it is run is a CellType Plugin. This plugin is responsible for defining cell types and storing cell type information. It is a basic plugin used by virtually every CompuCell simulation. The syntax is straight forward as can be seen in the example below: Here we have defined three cell types that will be present in the simulation: Medium,Fluid,Wall. Notice that we assign a number – TypeId – to every cell type. It is strongly recommended that TypeId’s are consecutive positive integers (e.g. 0,1,2,3...). Medium is traditionally given TypeId=0 but this is not a requirement. However every CC3D simulation mut define CellType Plugin and include at least Medium specification. Notice that in the example above cell type “Wall” has extra attribute Freeze=””. This attribute tells CompuCell that cells of “frozen” type will not be altered by spin flips. Freezing certain cell types is a very useful technique in constructing different geometries for simulations or for restricting ways in which cells can move. In the example below we have frozen cell types wall to create tube geometry for fluid flow studies. **IX.2.2. Simple Volume and Surface Constraints** One of the most commonly used energy term in the GGH Hamiltonian is a term that restricts variation of single cell volume. Its simplest form can be coded as show below: 25 2.0 By analogy we may define a term which will put similar constraint regarding the surface of the cell: 20 1.5 These two plugins inform CompuCell that the Hamiltonian will have two additional terms associated with volume and surface conservation. That is when spin flip is attempted one cell will increase its volume and another cell will decrease. Thus overall energy of the system may or will change. Volume constraint essentially ensures that cells maintain the volume which close (this depends on thermal fluctuations) to target volume . The role of surface plugin is analogous to volume, that is to “preserve” surface. Note that surface plugin is commented out in the example above. Energy terms for volume and surface constraints have the form: **Remark:** **Notice that flipping a single spin may cause surface change in more that two cells – this is especially true in 3D.** **IX.2.3.VolumeTracker and SurfaceTracker plugins** **** These two plugins monitor lattice and update volume and surface of the cells once spin flip occurs. In most cases users will not call those plugins directly. They will be called automatically when either Volume (calls Volume Tracker) or Surface (calls Surface Tracker) or CenterOfMass (calls VolumeTracker) plugins are requested. However one should be aware that in some situations, for example when doing foam coarsening simulation as presented in the introduction, when neither Volume or Surface plugins are called, one may still want to track changes ion surface or volume of cells . In such situations one can explicitely invoke VolumeTracker or Surface Tracker plugin with the following syntax: **** **IX.2.4. VolumeFlex Plugin** VolumeFlex plugin is more sophisticated version of Plugin. While Volume Plugin treats all cell types the same i.e. they all have the same target volume and lambda coefficient, VolumeFlex plugin allows you to assign different lambda and different target volume to different cell types. The syntax for this plugin is straightforward and essentially mimics the example below. Notice that in the example above cell types Wall and Ground have target volume and coefficient lambda set to 0 – very unusual. That's because in this particular those cells are were frozen so the parameters specified for these cells do not matter. In fact it is safe to remove specifications for these cell types, but just for the illustration purposes we left them. Using VolumeFlex Plugin you can effectively freeze certain cell types. All you need to do is to put very high lambda coefficient for the cell type you wish to freeze. You have to be careful though , because if initial volume of the cell of a given type is different from target volume for this cell type the cells will either shrink or expand to match target volume (this is out of control and you should avoid it), and only after this initial volume adjustment will they remain frozen . That is provided LambdaVolume is high enough. In any case, we do not recommend this way of freezing cells because it is difficult to use, and also not efficient in terms of speed of simulation run. **IX.2.5. SurfaceFlex Plugin** SurfaceFlex plugin is more sophisticated version of Plugin. Everything that was said with respect to plugin applies to . For syntax see example below: **IX.2.6. VolumeLocalFlex Plugin** VolumeLocalFlex Plugin is very similar to plugin. You specify both lambda coefficient and target volume, but as opposed to Volume Plugin the energy is calculated using target volume and lambda volume that are specified individually for each cell. In the course of simulation you can change this target volume depending on e.g. concentration of FGF in the particular cell. This way you can specify which cells grow faster, which slower based on a state of the simulation. This plugin requires you to develop a module (plugin or steppable) which will alter target volume for each cell. You can do it either in C++ or even better in Python. Example syntax: **IX.2.7. SurfaceLocalFlex Plugin** This plugin is analogous to VolumeLocalFlex but operates on cell surface. Example syntax: **** **IX.2.8. NeighborTracker Plugin** This plugin, as its name suggests, tracks neighbors of every cell. In addition it calculates common contact area between cell and its neighbors. We consider a neighbor this cell that has at least one common pixel side with a given cell. This means that cells that touch each other either “by edge” or by “corner” are not considered neighbors. See the drawing below: 5 5 5 4 4 5 5 5 4 4 5 5 4 4 4 1 1 2 2 2 1 1 2 2 2 **Figure 19**. Cells 5,4,1 are considered neighbors as they have non-zero common surface area. Same applies to pair of cells 4 ,2 and to 1 and 2. However, cells 2 and 5 are not neighbors because they touch each other “by corner”. Notice that cell 5 has 8 pixels cell 4 , 7 pixels, cell 1 4 pixels and cell 2 6 pixels. Example syntax: This plugin is used as a helper module by other plugins and steppables e.g. Elasticity and AdvectionDiffusionSolver use NeighborTracker plugin. **IX.2.9. Chemotaxis** Chemotaxis plugin , as its name suggests is used to simulate chemotaxis of cells. For every spin flip this plugin calculates change of energy associated with pixel move. There are several methods to define a change in energy due to chemotaxis. By default we define a chemotaxis using the following formula: where , denote chemical concentration at the spin-flip-source and spin-flip-destination pixel. respectively. We also support a slight modification of the above formula in the Chemotaxis plugin where is non-zero only if the cell located at after the spin flip is non-medium. to enable such mode users need to include tag in the body of XML plugin. Let's look at the syntax by studying the example usage of the Chemotaxis plugin: The body of the chemotaxis plugin description contains sections called ChemicalField. In this section you tell CompuCell3D which module contains chemical field that you wish to use for chemotaxis. In our case it is FlexibleDiffusionSolverFE. Next you need to specify the name of the field - FGF in our case. Next you specify lambda for each cell type so that cells of different type may respond differently to a given chemical. In particular types not listed will not respond to chemotaxis at all. Older versions of CompuCell3D allowed for different syntaxes as well. Despite the fact that those syntaxes are still supported for backward compatibility reasons, we discourage their use, because, they are somewhat confusing. Ocassionally you may want to use different formula for the chemotaxis than the one presented above. Current CompCell3D allows you to use the following definitions of change in chemotaxis energy (Saturation and SaturationLinear respectively ): or where '*s*' denotes saturation constant. To use first of the above formulas all you need to do is to let CompuCell3D know the value of the saturation coefficient: Notice that this only requires small change in line where you previously specified only lambda. To use second of the above formulas use SaturationLinearCoef instead of SaturationCoef: Sometimes it is desirable to have chemotaxis between only certain types of cells and not between other pairs of types. To deal with this situation it is enough to augment ChemotaxisByType element with the following attribute: This will cause that the change in chemotaxis energy will be non-zero only for those spin flip attempts that will try to slip Amoeba and Medium pixels. The definitions of chemotaxis presented so far do not allow specification of chemotaxis parameters individually for each cell. To do this we will use Python scripting. We still need to specify in the XML which fields are important from chamotaxis stand point. Only fields listed in the XML will be used to calculate chemotaxis energy: … … In the above excerpt from the XML configuration file we see that cells of type Macrophage will chemotax in response to ATTR gradient. Using Python scripting we can modify chemotaxing properties of individual cells as follows: class ChemotaxisSteering(SteppableBasePy): def \_\_init\_\_(self,\_simulator,\_frequency=100): SteppableBasePy.\_\_init\_\_(self,\_simulator,\_frequency) def start(self): for cell in self.cellList: if cell.type==2: cd=self.chemotaxisPlugin.addChemotaxisData(cell,"ATTR") cd.setLambda(20.0) # cd.initializeChemotactTowardsVectorTypes("Bacterium,Medium") cd.assignChemotactTowardsVectorTypes([0,1]) break def step(self,mcs): for cell in self.cellList: if cell.type==2: cd=self.chemotaxisPlugin.getChemotaxisData(cell,"ATTR") if cd: l=cd.getLambda()-3 cd.setLambda(l) break In the start function for first encountered cell of type Macrophage (type==2) we insert ChemotaxisData object (it determines chemotaxing properties) and initialize  parameter to 20. We also initialize vector of cell types towards which Macrophage cell will chemotax (it will chemotax towards Medium and Bacterium cells). Notice the break statement inside the if statement, inside the loop. It ensures that only first encountered Macrophage cell will have chemotaxing properties altered. In the step function we decrease lambda chemotaxis by 3 units every 100 MCS. In effect we turn a cell from chemotaxing up ATTR gradient to being chemorepelled. In the above example we have more than one macrophage but only one of them has altered chemotaxing properties. The other macrophages have chemotaxing properties set itn eh XML section. CompuCell3D first checks if local definitions of chemotaxis are available (i.e. for individual cells) and if so it uses those. Otherwise it will use definitions from from the XML. The ChemotaxisData structure has additional functions which allo to set chemotaxis formula used. For example we may type: def start(self): for cell in self.cellList: if cell.type==2: cd=self.chemotaxisPlugin.addChemotaxisData(cell,"ATTR") cd.setLambda(20.0) **cd.setSaturationCoef(200.0)** # cd.initializeChemotactTowardsVectorTypes("Bacterium,Medium") cd.assignChemotactTowardsVectorTypes([0,1]) break to activate Saturation formula. To activate SaturationLinear formula we would use: **cd.setSaturationLinearCoef(2.0)** **CAUTION:** when you use chemotaxis plugin you have to make sure that fields that you refer to and module that contains this fields are declared in the xml file. Otherwise you will most likely cause either program crash (which is not as bad as it sounds) or unpredicted behavior (much worse scenario, although unlikely as we made sure that in the case of undefined symbols, CompuCell3D exits) **IX.2.10. ExternalPotential plugin** Chemotaxis plugin is used to cause directional cell movement. Another way to achieve directional movement is to use ExternalPotential plugin. This plugin is responsible for imposing a directed pressure (or rather force) on cells. It is used mainly in fluid flow studies with periodic boundary conditions along these coordinates along which force acts. If NoFlux boundary conditions are set instead , the cells will be squeezed. This is the example usage of this plugin: Lambda is a vector quantity and determines components of force along three axes. In this case we apply force along x. We can also apply external potential to specific cell types: Where in ExternalPotentialParameters we specity which cell type is subject to external potential (Lambda is specified using x,y,z attributes). We can also apply external potential to individual cells. In that case, in the XML section we only need to specify: and in the Python file we change lambdaVecX, lambdaVecY, lambdaVecZ, which are properties of cell. For example in Python we could write: cell.lambdaVecX=-10 Calculations done by ExternalPotential Plugin are by default based on direction of pixel copy (similarly as in chemotaxis plugin). One can however force CC3D to do calculations based on movement of center of mass of cell. To use algorithm based on center of mass movement we use the following XML syntax: CenterOfMassBased **Remark:**Note that in the pixel-based algorithm the typical value of pixel displacement used in calculations is of the order of 1 (pixel) whereas typical displacement of center of mass of cell due to single pixel copy is of the order of 1/cell volume (pixels) – ~ 0.1 pixel. This implies that to achieve compatible behavior of cells when using center of mass algorithm we need to multiply lambda’s by appropriate factor, typicall of the order of 10. **IX.2.11. CellOrientation Plugin** Similarly as ExternalPotential plugin this plugin gives preference to those pixel copies whose direction aligns with polarization vector (which is a property of each cell): , where *(i)* denotes cell at site *i*, is polarization vector for cell at site *i* and pixel copy vector. Because two cell participate in the pixel copy process the net energy change is simply a sum of above expressions: one for growing cell and one for shrinking cell. To set lambda we have two options: use global setting in the XML: 0.5 Or set  individually for each cell and manage values of  from Python. In this case we use the following XML syntax: or equivalently the shorter version: If we manage  values in Python we would use the following syntax to acces and modify values of lambda: self.cellOrientationPlugin.getLambdaCellOrientation(cell) self.cellOrientationPlugin.setLambdaCellOrientation(cell,0.5) Calculations done by CellOrientation Plugin are by default based on direction of pixel copy (similarly as in chemotaxis plugin). One can however force CC3D to do calculations based on movement of center of mass of cell. To use algorithm based on center of mass movement we use the following XML syntax: CenterOfMassBased See remark in External potential description about rescaling of parameters when changing algorithm to Center Of Mass–based. **IX.2.12. PolarizationVector Plugin** PolarizationVector plugin is a simple plugin whose only task is to ensure that each cell in CompuCell3D simulation has as its attribute 3-component vector of floating point numbers. This plugin is normally used in together with CellOrientation but it also can be reused in other applications, assuming that we do not use CellOrientation plugin at the same time. The XML syntax is very simple: To access or modify polarization vector requires use of Python scripting. self.polarizationPlugin.getPolarizationVector(cell) or to change values of the polarization vector: self.polarizationPlugin.getPolarizationVector(cell,0.1,0.2,0.3) **IX.2.13. CenterOfMass Plugin** This plugin monitors changes n the lattice and updates centroids of the cell: where *i* denotes pixels belonging to a given cell. To obtain coordinates of a center of mass f a given cell you need to divide centroids by cell volume: This plugin is aware of boundary conditions and centroids are calculated properly regardless which boundary conditions are used. The XML syntax is very simple: **IX.2.12. Contact Energy** **** Energy calculations for the foam simulation are based on the boundary or contact energy between cells (or surface tension, if you prefer). Together with volume constraint contact energy is one of the most commonly used energy terms in the GGH Hamiltonian. In essence it describes how cells "stick" to each other. The explicit formula for the energy is: , where *i* and *j* label two neighboring lattice sites ,'s denote cell Ids,'s denote cell types . In the case of foam simulation the total energy of the foam is simply the total boundary length times the surface tension (here defined to be 2*J*). Once again, in the above formula, you need to differentiate between cell types and cell Ids. This formula shows that cell types and cell Ids are not the same. The Contact plugin in the .xml file, defines the energy per unit area of contact between cells of different types () and the interaction range (NeighborOrder) of the contact: 3 0 0 2 In this case, the interaction range is 2, thus only up to second nearest neighbor pixels of a pixel undergoing a change or closer will be used to calculate contact energy change. Foam cells have contact energy per unit area of 3 and Foam and Medium as well as Medium and Medium have contact energy of 0 per unit area. **IX.2.13. ContactLocalProduct Plugin** This plugin calculates contact energy based on local (i.e. per cell) cadhering expression levels. This plugin has to be used in conjunction with a steppable that assigns cadherin expression levels to the cell. Such steppables are usually written in Python – see ContactLocalProductExample in Demos directory. We use the following formulas to calculate energy for this plugin: By default. is a function of cadherins and can be either a simple product , a product of squared expression levels or a . In the case of the second formula plays the role of “regular” contact energy between cell and medium. The syntax of this plugin is as follows : 0 -16 -16 -2 2.75 -1 Quadratic 0.0 2 Users need to specify ContactSpecificity () between different cell types ContactFunctionType (by default it is set to Linear - but other allowed key words are Quadratic - and Min - ). EnergyOffset can be set to user specified value using above syntax. NeighborOrder has the same meaning as for “regular” Contact plugin. Alternatively one can write customized function of the two cadherins and use it instead of the 3 choices given above. To do this, simply use the following syntax: 0 -16 -16 -2 2.75 -1 Quadratic 0.0 2 J1 J2 sin(J1\*J2) Here we define variable names for cadherins in interacting cells (J1 denotes cadherin for one of the cells and cell2 denotes cadherin for another cell). Then in the Expression tag we give mathematical expression involving the two cadherin levels. The expression syntax has to follow syntax of the muParser - `http://muparser.sourceforge.net/mup\_features.html#idDef2 <%5Cl%20%22idDef2%22>`_. **IX.2.14. AdhesionFlex Plugin** Adhesion Flex is a generalization of ContactLocalProduct plugin. It allows setting individual adhesivity properties for each cell. Users can use either XML syntax or Python scripting to initialize adhesion molecule density for each cell. In addition, Medium can also carry its own adhesion molecules. We use the following formula to calculate Contact energy in AdhesionFlex plugin: where indexes *i*, *j* label pixels, denotes contact energy between cell types and , exactly as in “regular” contact plugin and indexes *m*,*n* label cadherins in cells composed f pixels *i* and *j* respectively. *F* denotes user-defined function of *N**m* and *N**n*. Altohugh this may look a bit complex, the basic idea is simple: each cell has certain number of cadherins on its surface. When cell touch each other the resultant energy is simpy a “product” -- of every cadherin from one cell with every cadherin from another cell.The XML syntax for this plugin is given below: ** min(Molecule1,Molecule2)** -1.0 2.0 -10.0 -10.0 2 *k**mn* matrix is specified within the AdhesionInteractionMatrix tag – the elements are listed using BindingParameter tags. The AdhesionMoleculeDensity tag specifies initial concentration of adhesion molecules. Even if you are going to modify those from Python (in the start function of the steppable) you are still required to specify the names of adhesion molecules and associate them with appropriate cell types. Failure to do so may result in simulation crash or undefined behaviors. The user-defined function *F* is specified using Formula tag where the arguments of the function are called Molecule1 and Molecule2. The syntax has to follow syntax of the muParser - `http://muparser.sourceforge.net/mup\_features.html#idDef2 <%5Cl%20%22idDef2%22>`_ . CompuCell3D example – *Demos/AdhesionFlex* - demonstrates how to manipulate concentration of adhesion molecules: self.adhesionFlexPlugin.getAdhesionMoleculeDensity(cell,"NCad") allows to access adhesion molecule concentration using its name (as given in the XML above using AdhesionMoleculeDensity tag). self.adhesionFlexPlugin.getAdhesionMoleculeDensityByIndex(cell,1) allows to access adhesion molecule concentration using its index in the adhesion molecule density vector. The order of the adhesion molecule densities in the vector is the same as the order in which they were declared in the XML above - AdhesionMoleculeDensity tags. self.adhesionFlexPlugin.getAdhesionMoleculeDensityVector(cell) allows access to entire adhesion molecule density vector. Each of these functions has its corresponding function whith operates on Medium. In this case we do not give cell as first argument: self.adhesionFlexPlugin.getMediumAdhesionMoleculeDensity(“Int”) self.adhesionFlexPlugin.getMediumAdhesionMoleculeDensityByIndex (0) self.adhesionFlexPlugin.getMediumAdhesionMoleculeDensityVector(cell) To change the value of the adhesion molecule density we use set functions: self.adhesionFlexPlugin.setAdhesionMoleculeDensity(cell,"NCad",0.1) self.adhesionFlexPlugin.setAdhesionMoleculeDensityByIndex(cell,1,1.02) self.adhesionFlexPlugin.setAdhesionMoleculeDensityVector(cell,\\ [3.4,2.1,12.1]) Notice that in this las function we passed entire Python list as the argument. CC3D will check if the number of entries in this vector is the same as the number of entries in the currently used vector. If so the values from the passed vector will be copied, otherwise they will be ignored. **IMPORTANT:** during mitosis we create new cell (childCell) and the adhesion molecule vector of this cell will have no components. However in order for simulation to continue we have to initialize this vector with number of cadherins appropriate to childCell type. We know that setAdhesionMoleculeDensityVector is not appropriate for this task so we have to use: self.adhesionFlexPlugin.assignNewAdhesionMoleculeDensityVector(cell,\\ [3.4,2.1,12.1]) which will ensure that the content of passed vector is copied entirely into cell’s vector (making size adjustments as necessry). **** **IMPORTANT: You have to make sure that the number of newly assigned adhesion molecules is exactly the same as the number of adhesion molecules declared for the cell of this particular type.** All of get functions has corresponding set function which operates on Medium: self.adhesionFlexPlugin.setMediumAdhesionMoleculeDensity("NCam",2.8) self.adhesionFlexPlugin.setMediumAdhesionMoleculeDensityByIndex(2,16.8) self.adhesionFlexPlugin.setMediumAdhesionMoleculeDensityVector(\\ [1.4,3.1,18.1]) self.adhesionFlexPlugin.assignNewMediumAdhesionMoleculeDensityVector(\\ [1.4,3.1,18.1]) **IX.2.15. ContactMultiCad Plugin** **** ContactMultiCad plugin is a modified version of ContactLocalProduct plugin. In this case users can use several cadherins and describe how they translate into contact energy. The energy formula used by this plugin is given below: where indexes *i*, *j* label pixels, denotes contact energy between cell types and , exactly as in “regular” contact plugin and indexes m,n label cadherins in cells composed f pixels *i* and *j* respectively. The syntax for this plugin is as follows: 0 0 0 0 0 -10 -12 -1 0.0 2 Entries of the type 0 have the same meaning as in “regular” contact energy. Specificity parameters specification are enclosed between tags and . The names NCad0 and Ncad1 are arbitrary. However the matrix will be ordered according to lexographic order of Cadherin names. For that reason we recommend that you name cadherins in such a way that makes it easy what the order will be. As in the example above using *NameNumber* (e.g. NCad0, NCad1) makes it easy to figure out what the order will be (NCad0 will get index 0 and NCad1 will get index 1). This is important because cadherins will be set in Python and if you won't keep track of the ordering of the specificity you might wrongly assign cadherins in Python and get unexpected results. In the example the order of cadherins is clear based on the definition of cadherin specificity parameters. **IX.2.15. MolecularContact** This plugin is analogous to ContactLocalProduct and allows users to specify functional form of adhesion molecules interactions using Python syntax. It is in beta state and for this reason we are not discussing it in more detail and currently suggest to use Either AdhesionFlex or ContactLocal product plugins. **IX.2.15. ContactCompartment** This plugin is a generalization of the contact energy plugin for the case of compartmental cell models. where *i* and *j* denote pixels ,  denotes, as before, a cell type of a cell with  cluster id and  cell id. In compartmental cell models a cell is a collection of subcells. Each subcell has a unique id (cell id). In addition to that each subcell will have additional attribute, a cluster id that determines to which cluster of subcells a given subcell belongs. (think of a cluster as a cell with nonhomogenous cytoskeleton) The idea here is to have different contact energies between subcells belonging to the same cluster and different energies for cells belonging to different clusters. Technically subcells of a cluster are “regular” CompuCell3D cells. By giving them an extra attribute cluster id we can introduce a concept of compartmental cells. In our convention *0,0)* denotes medium Introduction of cluster id and cell id are essential for the definition of. As you can see from above there are two hierarchies of contact energies – external and internal. The energies depend on cell types as in the case “regular” Contact plugin. Now, however, depending whether pixels for which we calculate contact energies belong to the same cluster or not we will use internal or external contact energies respectively. **IX.2.16. LengthConstraint Plugin** This plugin imposes elongation constraint on the cell. Effectively it “measures” a cell along its “axis of elongation” and ensures that cell length along the elongation axis is close to target length. For detailed description of this algorithm in 2D see Roeland Merks' paper “Cell elongation is a key to in silico replication of in vitro vasculogenesis and subsequent remodeling” Developmental Biology **289**(2006) 44-54). This plugin is usually used in conjunction with Connectivity Plugin. The syntax is as follows: LambdaLength determines the degree of cell length oscillation around TargetLength parameter. The higher LambdaLength the less freedom a cell will have to deviate from TargetLength. In the 3D case we use the following syntax: Notice new attribute called MinorTargetLength. In 3D it is not sufficient to constrain the "length" of the cell you also need to constrain "width" of the cell along axis perpendicular to the major axis of the cell. This "width" is referred to as MinorTargetLength. For 2D simulations we have also an option to use LengthConstraintLocalFlex plugin which calculate elongation constraints based on local parameters (i.e. on a per cell basis). The syntax is as follows: The parameters are assigned using Python – see *Demos\\elongationLocalFlexTest*example. **Remark:** For 3D simulations we can only define elongation parameters pn a per cell type basis. We will fix this limitation in the next release. **Remark:**When using target length plugins (either global , as shown here, or local as we will show in the subsequent subsection) it is important to use connectivity constraint. This constrain will check if a given pixel copy can break cell connectivity. If so, it will add large energy penalty (defined by a user) to change of energy effectively prohibiting such pixel copy. In the case of 2D on square lattice checking cell connectivity can be done locally and thus is very fast. Unfortunately on hex lattice and in 3D on either lattice we don’t have an algorithm of performing such check locally and therefore we do it globally using breadth first search algorithm and comparing volumes of cells calculated this way with actual volume of the cell. If they agree we conclude that cell connectivity is preserved. However the computational cost of running such algorithm, can be quite high. Therefore if one does need extremely elongated cells (it is when connectivity algorithm has to do a lot of work) one may neglect connectivity constraint and use Length constrain only. For slight cells elongations the connectivity should be preserved however, occasionally cells may fragment. **IX.2.17. Connectivity Plugins** The basic Connectivity plugin works **only in 2D and only on square lattice** and is used to ensure that cells are connected or in other words to prevent separation of the cell into pieces. The detailed algorithm for this plugin is described in Roeland Merks' paper “Cell elongation is a key to *in-silico* replication of in vitro vasculogenesis and subsequent remodeling” Developmental Biology **289** (2006) 44-54). There was one modification of the algorithm as compared to the paper. Namely, to ensure proper connectivity we had to reject all spin flips that resulted in more that two collisions. (see the paper for detailed explanation what this means). The syntax of the plugin is straightforward: 100000 **** Penalty denotes energy that will be added to overall change of energy if attempted spin flip would violate connectivity constraints. If the penalty is positive and much larger than the absolute value of other energy changes in the simulation this has the effect of preventing a spin flip from occurring. A more general type of connectivity constraint is implemented in ConnectivityGlobal plugin. In this case we calculate volume of a cell using breadth first search algorithm and compare it with actual volume of the cell. If they agree we conclude that cell connectivity is preserved. This plugin works both in 2D and 3D and on either type of lattice. However the computational cost of running such algorithm, can be quite high so it is best to limit this plugin to cell types for which connectivity of cell is really essential: 1000000000 In certain types of simulation it may happen that at some point cells change cell types. If a cell that was not subject to connectivity constraint, changes type to the cell that is constrained by global connectivity and this cell is fragmented before type change this situation normally would result in simulation freeze. However CompuCell3D, first before applying constraint it will check if the cell is fragmented. If it is, there is no constraint. Global connectivity constraint is only applied when cell is non-fragmented. The numerical value of Penalty in the XML syntax above does not really matter as long as it is greater than 0. CompuCell3D guarantees that cells for which penalty is greater than 0 will remain connected. We will modify global connectivity plugin to allow application of connectivity constraints to individual cells. Quite often in the simulation we don't need to impose connectivity constraint on all cells or on all cells of given type. Usually only select cell types or select cells are elongated and therefore need connectivity constraint. In such a case we use ConnectivityLocalFlex plugin and assign connectivity constraints to particular cells in Python In XML we only declare: In Python we manipulate/access connectivity parameters for individual cells using the following syntax: self.connectivityLocalFlexPlugin.setConnectivityStrength(cell,20.7) self.connectivityLocalFlexPlugin.getConnectivityStrength(cell) See also example in *Demos\\elongationLocalFlexTest.* **IX.2.18. Mitosis Plugin** Mitosis plugin carries out cell division into two cells once the parent cell reaches critical volume (DoublingVolume). The two cells after mitosis will have approximately the same volume although it cannot be guaranteed in general case if the parent cell is fragmented. One major problem with Mitosis plugin is that after mitosis the attributes of the offspring cell might not be initialized properly. By default cell type of the offspring cell will be the same as cell type of parent and they will also share target volume . All other parameters for the new cell remain uninitialized. **Remark:** For this reason we stringly recommend using Mitosis plugin through Python interface as there users can quite easily customize what happens to parent and offspring cells after mitosis. An example of the use of Mitosis plugin through Python scripting is provided in CompuCell3D’s Python Scripting Manual. The syntax of the “standard” mitosis plugin is the following: 50 Every time a cell reaches DoublingVolume it will undergo the mitosis and the offspring cell will inherit type and target volume of the parent. If this simple behavior is unsatisfactory consider use Python scripting to implement proper mitotic divisions of cells. **IX.2.19. Secretion Plugin** In earlier version os of CC3D secretion was part of PDE solvers. We still support this mode of model description however, starting in 3.5.0 we developed separate plugin which handles secretion only. Via secretion plugin we can simulate cell ular secretion of various chemicals. The secretion plugin allows users to specify various secretion modes in the XML file - XML syntax is practically identical to the SecretionData syntax of PDE solvers. In addition to this Secretion plugin allows users to maniupulate secretion properties of individual cells from Python level. To account for possibility of PDE solver being called multiple times during each MCS, the Secretion plugin can be called multiple times in each MCS as well. We leave it up to user the rescaling of secretion constants when using multiple secretion calls in each MCS. **Note**:Secretion for individual cells invoked via Python will be called only once per MCS. Typical XML xyntax for Secretion plugin is presented below: 200 300 500 By default ExtraTimesPerMC is set to 0 - meaning no extra calls to Secretion plugin per MCS. Typical use of secretion from Python is dempnstrated best in the example below: class SecretionSteppable(**SecretionBasePy**): def \_\_init\_\_(self,\_simulator,\_frequency=1): **SecretionBasePy**.\_\_init\_\_(self,\_simulator, \_frequency) def step(self,mcs): attrSecretor=self.getFieldSecretor("ATTR") for cell in self.cellList: if cell.type==3: attrSecretor.secreteInsideCell(cell,300) attrSecretor.secreteInsideCellAtBoundary(cell,300) attrSecretor.secreteOutsideCellAtBoundary(cell,500) attrSecretor.secreteInsideCellAtCOM(cell,300) **Remark**: Instead of using SteppableBasePy class we are using SecretionBasePy class. The reason for this is that in order for secretion plugin with secretion modes accessible from Python to behave exactly as previous versions of PDE solvers (where secretion was done first followed by “diffusion” step) we have to ensure that secretion steppable implemented in Python is called before each Monte Carlo Step, which implies that it will be also called before “diffusing” function of the PDE solvers. SecretionBasePy sets extra flag which ensures that steppable which inherits from SecretionBasePy is called before MCS (and before all “regular’ Python steppables). There is no magic to SecretionBasePy - if you still want to use SteppableBasePy as a base class for secretion (or for that matter SteppablePy) do so, but remember that you need to set flag: self.runBeforeMCS=1 to ensure that your new stoppable will run before each MCS. See example below for alternative implementation of SecretionSteppable using SteppableBasePy as a base class: class SecretionSteppable(**SteppableBasePy**): def \_\_init\_\_(self,\_simulator,\_frequency=1): **SteppableBasePy**.\_\_init\_\_(self,\_simulator, \_frequency) **self.runBeforeMCS=1** def step(self,mcs): attrSecretor=self.getFieldSecretor("ATTR") for cell in self.cellList: if cell.type==3: attrSecretor.secreteInsideCell(cell,300) attrSecretor.secreteInsideCellAtBoundary(cell,300) attrSecretor.secreteOutsideCellAtBoundary(cell,500) attrSecretor.secreteInsideCellAtCOM(cell,300) The secretion of individual cells is handled through Field Secretor objects. Field Secretor concenpt is quite convenient because the amoun of Python coding is quite small. To secrete chemical (this is now done for individual cell) we first create field secretor object, attrSecretor=self.getFieldSecretor("ATTR"), which allows us to secrete into field called ATTR. **Remark:** Make sure that fields into which you will be secreting chemicals exist. They are usually fields defined in PDE solvers. When using secretion plugin you do not need to specify SecretionData section for the PDE solvers Then we pick a cell and using field secretor we simulate secretion of chemical ATTR by a cell: attrSecretor.secreteInsideCell(cell,300) Currently we support 4 secretion modes for individual cells: #. 1. secreteInsideCell – this is equivalent to secretion in every pixel belonging to a cell #. 2. secreteInsideCellAtBoundary – secretion takes place in the pixels belonging to the cell boundary #. 3. secreteInsideCellAtBoundary – secretion takes place in pixels which are outide the cell but in contact with cell boundary pixels #. 4. secreteInsideCellAtCOM – secretion at the center of mass of the cell As you may infer from above modes 1, 2 and 3 require tracking of pixels belonging to cell and pixels belonging to cell boundary. If you are not using modes 1-3 you may disable pipxel tracking by including and/or tags – as shown in the example below: 200 300 500 **IX.2.20. PDESolverCaller Plugin** PDE solvers in CompuCell3D are implemented as steppables . This means that by default they are called every MCS. In many cases this is insufficient. For example if diffusion constant is large, then explicit finite difference method will become unstable and the numerical solution will have no sense. To fix this problem one could call PDE solver many times during single MCS. This is precisely the task taken care of by PDESolverCaller plugin. The syntax is straightforward: All you need to do is to give the name of the steppable that implements a given PDE solver and pass let CompCell3D know how many extra times per MCS this solver is to be called (here FlexibleDiffusionSolverFE was 8 extra times per MCS). **IX.2.21. Elasticity Plugin and ElasticityTracker Plugin** This plugin is responsible for handling the following energy term: where is a distance between center of masses of cells *i* and *j* and is a target length corresponding to . The syntax of this plugin is the following 200.0 6 In this case and are the same for all participating cells types. By adding extra attribute to the above plugin: 200.0 6 we tell CompuCell3D to use and defined on per pair of cells basis. The initialization of and usually takes place in Python script and users must make sure that and or else one can get unexpected results. We provide example python and xml files that demo the use of plasticity plugin. Users have to specify which cell types participate in the plasticity calculations. This is done by including ElasticityTracker plugin **before** Elasticity plugin in the xml file. The syntax is very clear: Body1 Body2 Body3 All is required is a list of participating cell types. Here cells of type Body1, Body2 and Body3 will be taken into account for elasticity energy calculation purposes. The way in which CompuCell3D determines which cells are to be included in the elasticity energy calculations is by examining which cells are in contact with each other before simulation begins . If the types of cells touching each other are listed in the list of IncudeTypes of ElasticityTracker then such cells are being taken into account when calculating elastic constraint. Cells which initially are not touching will not participate in calculations even if their type is included in the list of “ElasticityTracker”. However, in some cases it is desirable to add elasticity pair even for cells that do not touch each other or do it once simulation has started. To do this ElasticityTracker plugin defines two function : assignElasticityPair(\_cell1 , \_cell2) removeElasticityPair(\_cell1 , \_cell2) where \_cell1 and \_cell2 denote pointers to cell objects. These functions add or remove two cell links to or from elastic constraint. Typically they are called from Python level. **IX.2.22. FocalPointPlasticity Plugin** Similarly as Elasticity plugin, FocalPointPlasticity pust constrains the distance between cells’ center of masses. The main difference is that the list of “focal point plasticity neighbors” can change as the simulation goes and user specifies the maximum number of “focal point plasticity neighbors” a given cell can have. Let’s look at relatively simple XML syntax of FocalPointPlasticityPlugin (see *Demos/FocalPointPlasticity* example and we will show more complex examples later): 10.0 -50.0 7 20.0 2 10.0 -50.0 7 20.0 2 1 Parameters section describes properties of links between cells. MaxNumberOfJunctions, ActivationEnergy, MaxDistance and NeighborOrder are responsible for establishing connections between cells. CC3D constantly monitors pixel copies and during pixel copy between two neighboring cells/subcells it checks if those cells are already participating in focal point plasticity constraint. If they are not, CC3D will check if connection can be made (e.g. Condensing cells can have up to two connections with Condensing cells and up to 2 connections with NonCondensing cells – see first line of Parameters section and MaxNumberOfJunctions tag). The NeighborOrder parameter determines the pixel vicinity of the pixel that is about to be overwritten which CC3D will scan in search of the new link between cells. NeighborOrder 1 (which is default value if you do not specify this parameter) means that only nearest pixel neighbors will be visited. The ActivationEnergy parameter is added to overall energy in order to increase the odds of pixel copy which would lead to new connection. Once cells are linked the energy calculation is carried out in a very similar way as for the Elasticity plugin: where is a distance between center of masses of cells *i* and *j* and is a target length corresponding to . *ij* and *L**ij* between different cell types are determined using Lambda and TargetDistance tags. The MaxDistance determines the distance between cells’ center of masses when the link between those cells break. When the link breaks, then in order for the two cells to reconnect they would need to come in contact (in order to reconnect). However it is usually more likely that there will be other cells in the vicinity of separated cells so it is more likely to establish new link than restore broken one. The above example was one of the simplest examples of use of FocalPointPlasticity. A more complicated one involves compartmental cells. In this case each cell has separate “internal” list of links between cells belonging to the same cluster and another list between cells belonging to different clusters. The energy contributions from both lists are summed up and everything that we have said when discussing example above applies to compartmental cells. Sample syntax of the FocalPointPlasticity plugin which includes compartmental cells is shown below. We use InternalParameters tag/section to describe links between cells of the same cluster (see *Demos/FocalPointPlasticity* example): 10.0 -50.0 7 20.0 1 10.0 -50.0 7 20.0 1 10.0 -50.0 7 20.0 1 10.0 -50.0 7 20.0 1 1 Sometimes it is necessary to modify link parameters individually for every cell pair. In this case we would manipulate FocalPointPlasticity links using Python scripting. Example *Demos/FocalPointPlasticityCompartments* demonstrates exactly this situation. Still, you need to include XML section as the one shown above for compartmental cells, because we need to tell CC3D how to link cells. The only notable difference is that in the XML we have to include tag to signal that we will set link parameters (Lambda, TaretDistance, MaxDistance) individually for each cell pair: 10.0 -50.0 7 20.0 1 …….. Python steppable where we manipulate cell-cell focal point plasticity link properties is shown below: class FocalPointPlasticityCompartmentsParams(SteppablePy): def \_\_init\_\_(self,\_simulator,\_frequency=10): SteppablePy.\_\_init\_\_(self,\_frequency) self.simulator=\_simulator self.focalPointPlasticityPlugin=CompuCell.\\ getFocalPointPlasticityPlugin() self.inventory=self.simulator.getPotts().\\ getCellInventory() self.cellList=CellList(self.inventory) def step(self,mcs): for cell in self.cellList: for fppd in InternalFocalPointPlasticityDataList\\ (self.focalPointPlasticityPlugin,cell): self.focalPointPlasticityPlugin.\\ setInternalFocalPointPlasticityParameters\\ (cell,fppd.neighborAddress,0.0,0.0,0.0) The syntax to change focal point plasticity parameters (or as here internal parameters) is as follows: setFocalPointPlasticityParameters(cell1, cell2, lambda,\\ targetDistance, maxDistance) setInternalFocalPointPlasticityParameters(cell1, cell2, lambda,\\ targetDistance, maxDistance) Similarly to inspect current values of the focal point plasticity parameters we would use the following Python construct: for cell in self.cellList: for fppd in InternalFocalPointPlasticityDataList\\ (self.focalPointPlasticityPlugin,cell): print "fppd.neighborId",fppd.neighborAddress.id " lambda=",fppd.lambdaDistance For non-internal parameters we simply use FocalPointPlasticityDataList instead of InternalFocalPointPlasticityDataList . Examples *Demos/FocalPointPlasticity…* show in relatively simple way how to use FocalPointPlasticity plugin. Those examples also contain useful comments. When using FocalPointPlasticity Plugin from mitosis module one might need to break or create focal point plasticity links. To do so FocalPointPlasticity Plugin provides 4 convenience functions which can be invoked from the Python level: deleteFocalPointPlasticityLink(cell1,cell2) deleteInternalFocalPointPlasticityLink(cell1,cell2) createFocalPointPlasticityLink(\\ cell1,cell2,lambda,targetDistance,maxDistance) createInternalFocalPointPlasticityLink(\\ cell1,cell2,lambda,targetDistance,maxDistance) **IX.2.23.Curvature Plugin** This plugin implements energy term for compartmental cells. It is based on “A New Mechanism for Collective Migration in *Myxococcus xanthus*”, J. Starruß, Th. Bley, L. Søgaard-Andersen and A. Deutsch, *Journal of Statistical Physics*, DOI: **10.1007/s10955-007-9298-9**, (2007). For a “long” compartmental cell composed of many subcells it imposes constraint on curvature of cells. The syntax is slightly complex: 100.0 -50.0 100.0 -50.0 100.0 -50.0 The InternalTypeSpecificParameter tells Curvature Plugin how many neighbors a cell of given type will have. In this case, numbers which make sense are 1 and 2. The middle segment will have 2 connection and head and tail segments will have only one connection with neighboring segmens (subcells). The connections are established dymamically. The way it happens is that during simulation CC3D constantly monitors pixel copies and during pixel copy between two neighboring cells/subcells it checks if those cells are already “connected” using curvature constraint. If they are not, CC3D will check if connection can be made (e.g. Center cells can have up to two connections and Top and Bottom only one connection). Usually establishing connections takes place at the beginning if the simulation and often happens within first Monte Carlo Step (depending on actual initial configuration, of course, but if segments touch each other connections are established almost immediately). The ActivationEnergy parameter is added to overall energy in order to increase the odds of pixel copy which would lead to new connection. Lambda tag/parameter determines “the strength” of curvature constraint. The higher the Lambda the more “stiff” cells will be i.e. they will tend to align along straight line. **IX.2.24.PlayerSettings Plugin** This plugin allows users to specify or configure Player settings directly from XML, without s single click. Some users might prefer this way of setting configuring Player. In addition to this if users want to run two different simulations at the same time on the same machine but with different , say, cell colors, then doing it with “regular” Player configuration file might be tricky. The solution is to use PlayerSetting Plugin. The syntax of this plugin is as follows: As can be seen from above syntax all the keywords correspond to an action in the Player. Project2D sets up the values of the projection on the Player steering bar. Here we set the player to start 2D display in the xz projection with y coordinate set to 50 . Borders and contours properties are handled using Border and Contour elements. Specifying cell colors is done using Cell element. VisualControl element allows users to specify zoom factor and screen update and screenshot frequencies. Notice, screen update frequency migh not work properly when using Python script. In this case CompuCell will use whatever screen update frequency was stored in the config file (by default 1). We may also change things such as screen update frequency or screenshot frequency and choose whether or not to close the player after the simulation. To start Player in the 3D view instead of adding tag we add section: The camera settings stored here position 3D camera. The best way to get those settings is to run a simulation, add 2D screenshot using camera button, switch to 3D and position camera (using mouse) however you like and subsequently add another screenshot using camera button, save screenshot descrition file (File->Save Screenshot Description…) and open up in text editor newly saved screenshot description file (with .sdfml extension) and copy camera setting from there into PlayerSettings section. An example of using Player settings is shown in *Demos\\cellsort\_2D\\cellsort\_2D\_PlayerSettings.xml*. Although the set of allowed changes of player settings is fairly small at the moment we believe that the options that users have right now are quite sufficient for configuring the Player from the XML or python level. We will continue adding new options though. **IX.2.25.BoundaryPixelTracker Plugin** This plugin allows storing list of boundary pixels for each cell. The syntax is as follows: 1 This plugin is also used by other plugins as a helper module. **IX.2.26.GlobalBoundaryPixelTracker** This plugin tracks boundary pixels of all the cells including medium It is used in a Boundary Walker algorithm where instead of blindly picking pixel copy candidate we pick it from the set of pixels comprising boundaries of non frozen cells. In situations when lattice is large and there are not that many cells it makes sense to use BoundaryWalker algorithm to limit number of "wrong" pixel picks when perfming pixel copy attempts. Take a look at the following example: 10 10000 5 1 2 BoundaryWalker Periodic 2 Here we are using BoundaryWalkeralgorithm (Pottssection) and subsequently we list GlobalBoundaryTracker plugin where we set neighbor order to match that in the Potts section. The neighbor order determines how "thick" the overall boundary of cells will be. The higher this number the more pixels will belong to the boundary. **IX.2.27. PixelTracker Plugin** This plugin allows storing list of all pixels belonging to a given cell. The syntax is as follows: This plugin is also used by other plugins (e.g. Mitosis) as a helper module. **IX.2.28.MomentOfInertiaplugin** This plugin updates tensor of inertia for every cell. Internally it uses parallel axis theorem to calculate most up-to-date tensor of inertia. It can be called directly: However, most commonly it is called indirectly by other plugins like Elongation plugin. MomentOfInertia plugin gives users access (via Python scripting) to current lengths of cell’s semiaxes. Examples in Demos/MomentOfInertiademonstrate how to get lengths of semiaxes: axes=self.momentOfInertiaPlugin.getSemiaxes(cell) axes is a 3-component vector with 0thelement being length of minor axis, 1st– length of median axis (which is set to 0 in 2D) and 2ndelement indicating the length of major semiaxis. **Important:** Because calculating lengths of semiaxes involves quite a few of floating point operations it may happen (usually on hexagonal lattice) that for cells composed of 1, 2, or 3 pixels one moment the square of one of the semiaxes may end up being slightly negative leadind to NaN(not a number)length. This is due to roundoff error and whenever CC3D detects very small absolute value of square of the length of semiaxes (10-6) it sets length of this semiaxes to 0.0 regardless whether the squared value is positive or negative. However it is a good practice to test whether the length of semiaxis is sane by doing simple if (here we show how to test for a NaN) jf length!=length: print “length is NaN”: else: print “length is a proper floating point number” **IX.2.29.SimpleClockplugin** This plugin adds an integer as a cell attribute: **IX.2.30.ConvergentExtensionplugin** This is very specialized plugin which currently is in Tier 2 plugins in terms of support. IT implements energy term described in “Simulating Convergent Extension by Way of Anisotropic Differemtial Adhesion,” *Zajac* M, *Jones* GL, and *Glazier* JA, Journal of Theoretical Biology **222** (2), 2003. CC3D’s ConvergentExtension plugin is a somewhat simplified version of energy term described in the paper. This plugin uses the following syntax: 0.99 0.99 2 The Alphatag represents numerical value of parameter from the paper. **IX.3. Steppable Section** Steppables are CompuCell modules that are called every Monte Carlo Step (MCS). More precisely, they are called after all the spin attempts in a given MCS have been carried out. Steppables may have various functions like for example solving PDE's, checking if critical concentration threshold have been met, updating target volume or target surface given the concentration of come growth factor, initializing cell field, writing numerical results to a file etc. In summary Steppables perform all functions that need to be done every MCS. In the reminder of this section we will present steppables currently available in the CompuCell and describe their usage. **IX.3.1 UniformInitializer Steppable** This steppable lays out pattern of cells on the lattice. It allows users to specify rectangular regions of field with square (or cube in 3D) cells of user defined types (or random types). Cells can be touching each other or can be separated by a gap. The syntax of the plugin is as follows: 0 5 psm Above we have defined a 2D rectangular box filled with 5x5 cells touching each other (Gap=0) and having type psm. Notice that if you want to initialize 2D box in xz plane as above then y\_min and y\_max have to be 0 and 1 respectively. Users can include as many regions as they want. The regions can overlap each other. Simply cells that are overwritten will either disappear or be truncated. Additionally users can initialize region with random cell types chosen from provided list of cell types: 0 5 psm,ncad,ncam When user specifies more than one cell type between tags (notice, the types have to be separated with ',' and there should be no spaces) then cells for this region will be initialized with types chosen randomly from the provided list (here the choices would be psm, ncad, ncam). **Remark:** If one of the type names is repeated inside element this type will get greater weighting means probability of assigning this type to a cell will be greater. So for example psm,ncad,ncam,ncam,ncam ncam will assigned to a cell with probability 3/5 and psm and ncad with probability 1/5. **IX.3.2. BlobInitializer Steppable** This steppable is used to lay out circular blob of cells on the lattice. This plugin does not have yet the flexibility of UniformInitializer but this will change in the future release. Original syntax of this plugin looks as follows: 0 5 yes 40 The blob is centered in the middle of th lattice and has radius given by parameter all cells are initially squares (or cubes in 3D) - determines the length of the cube or square side and determines space between squares or cubes. tag and value yes is used to initialize cells randomly with type id being either 1 or 2. Otherwise all cells will have type id 1. This can be easily modified in Python . The most recent syntax for this plugin gives users additional flexibility in initializing cell field using BlobFieldInitializer: 0 5 40
Condensing,NonCondensing Similarly as for the UniformFieldInitializer users can define many regions each of which is a blob of a particular center point , radius and list of cell types that will be assigned to cells forming the blob. **IX.3.3. PIF Initializer** To initialize the configuration of the simulation lattice you can can write your own **lattice initialization file**. Our experience suggests that you will probably have to write your own initialization files rather than relying on built-in initializers. The reason is simple: the built-in initializers implement very simple cell layouts, and if you want to study more complicated cell arrangements, the built-in initializers will not be very helpful. Therefore we encourage you to learn how to prepare lattice initialization files. Again, file definition is not complicated and we will explain every step. The lattice initialization file tells CompuCell3D how to lay out assign the simulation lattice pixels to cells. **** The **Potts Initial File** (***PIF***) is a simple file format that we created for easy specification of initial cell positions. The PIF consists of multiple lines of the following format: cell# celltype x1 x2 y1 y2 z1 z2 Where cell# is the unique integer index of a cell, celltype is a string representing the cell's initial type, and x1 and x2 specify a *range* of x-coordinates contained in the cell (similarly y1 and y2 specify a range of y-coordinates and z1 and z2 specify a range of z-coordinates). Thus each line assigns a rectangular volume to a cell. If a cell is not perfectly rectangular, multiple lines can be used to build up the cell out of rectangular sub-volumes (just by reusing the cell# and celltype). A PIF can be provided to CompuCell3D by including the steppable object **PIFInitializer.** Let's look at a PIF example for foams: 0 Medium 0 101 0 101 0 0 1 Foam 13 25 0 5 0 0 2 Foam 25 39 0 5 0 0 3 Foam 39 46 0 5 0 0 4 Foam 46 57 0 5 0 0 5 Foam 57 65 0 5 0 0 6 Foam 65 76 0 5 0 0 7 Foam 76 89 0 5 0 0 These lines define a background of Medium which fills the whole lattice and is then overwritten by seven rectangular cells of type Foam numbered 1 through 7. Notice that these cells lie in the xy plane (z1=0 z2=0 implies that cells have thickness =1) so this example is a two-dimensional initialization. You can write the PIF file manually, but using a script or program that will write PIF file for you in the language of your choice (Perl, Python, Matlab, Mathematica, C, C++, Java or any other programming language) will save a great deal of typing. You may also use tools like PIFTracer which allow you to "paint" the lattice by tracing regions of the experimental pictures. Notice, that for compartmental cell model the format of the PIF file is different: Include Clusters cluster # cell# celltype x1 x2 y1 y2 z1 z2 For example: Include Clusters 1 1 Side1 23 25 47 56 10 14 1 2 Center 26 30 50 54 10 14 1 3 Side2 31 33 47 56 10 14 1 4 Top 26 30 55 59 10 14 1 5 Bottom 26 30 45 49 10 14 2 6 Side1 35 37 47 56 10 14 2 7 Center 38 42 50 54 10 14 2 8 Side2 43 45 47 56 10 14 2 9 Top 38 42 55 59 10 14 2 10 Bottom 38 42 45 49 10 14 **IX.3.4. PIFDumper Steppable** This steppable does opposite to PIFIitializer – it writes PIF file of current lattice configuration. The syntax similar to the syntax of PIFInitializer: line Notice that we used Frequency attribute of steppable to ensure that PIF files are written every 100 MCS. Without it they would be written every MCS. The file names will have the following format: PIFName.MCS.pif In our case they would be line.0.pif, line.100.pif, line.200.pif etc... This plugin is actually quite useful. For example, if we want to start simulation from a more configuration of cells (not rectangular cells as this is the case when we use Uniform or Blob initializers). In such a case we would run a simulation with a PIFDumper included and once the cell configuration reaches desired shape we would stop and use PIF file corresponding to this state. Once we have PIF initial configuration we may run many simulation starting from the same, realistic initial condition. **IX.3.5. Mitosis Steppabe.** This steppable is described in great detail in Python tutorial but because of its importance we are including a copy of that description here. In developmental simulations we often need to simulate cells which grow and divide. In earlier versions of CompuCell3D we had to write quite complicated plugin to do that which was quite cumbersome and unintuitive (see example 9). The only advantage of the plugin was that exactly after the pixel copy which had triggered mitosis condition CompuCell3D called cell division function immediately. This guaranteed that any cell which was supposed divide at any instance in the simulation, actually did. However, because state of the simulation is normally observed after completion of full a Monte Carlo Step, and not in the middle of MCS it makes actually more sense to implement Mitosis as a steppable. Let us examine the simplest simulation which involves mitosis. We start with a single cell and grow it. When cell reaches critical (doubling) volume it undergoes Mitosis. We check if the cell has reached doubling volume at the end of each MCS. The folder containing this simulation is *examples\_PythonTutorial/steppableBasedMitosis*. The mitosis algorithm is implemented in *examples\_PythonTutorial/steppableBasedMitosis/steppableBasedMitosisSteppables.py* File: *examples\_PythonTutorial/steppableBasedMitosis/steppableBasedMitosisSteppables.py* from PySteppables import \* from PySteppablesExamples import MitosisSteppableBase import CompuCell import sys class VolumeParamSteppable(SteppablePy): def \_\_init\_\_(self,\_simulator,\_frequency=1): SteppablePy.\_\_init\_\_(self,\_frequency) self.simulator=\_simulator self.inventory=self.simulator.getPotts().getCellInventory() self.cellList=CellList(self.inventory) def start(self): for cell in self.cellList: cell.targetVolume=25 cell.lambdaVolume=2.0 def step(self,mcs): for cell in self.cellList: **cell.targetVolume+=1** class **MitosisSteppable(MitosisSteppableBase)**: def \_\_init\_\_(self,\_simulator,\_frequency=1): MitosisSteppableBase.\_\_init\_\_(self,\_simulator, \_frequency) def step(self,mcs): **cells\_to\_divide=[]** for cell in self.cellList: if cell.volume>50: **# mitosis condition** cells\_to\_divide.append(cell) **for cell in cells\_to\_divide: self.divideCellRandomOrientation(cell) **def updateAttributes(self): parentCell=self.mitosisSteppable.parentCell childCell=self.mitosisSteppable.childCell childCell.targetVolume=parentCell.targetVolume childCell.lambdaVolume=parentCell.lambdaVolume if parentCell.type==1: childCell.type=2 else: childCell.type=1 Two steppables: VolumeParamSteppableand MitosisSteppableare the essence of the above simulation. The first steppable initializes volume constraint for all the cells present at T=0 MCS (only one cell) and then every 10 MCS (see the frequency with which VolumeParamSteppable in initialized to run - *examples\_PythonTutorial/steppableBasedMitosis/steppableBasedMitosis.py*) it increases target volume of cells, effectively causing cells to grow. The second steppable checks every 10 MCS (we can, of course, run it every MCS) if cell has reached doubling volume of 50. If so such cell is added to the list cells\_to\_divide which subsequently is iterated and all the cells in it divide. **Remark:** It is important to divide cells outside the loop where we iterate over entire cell inventory. If we keep dividing cells in this loop we are adding elements to the list over which we iterate over and this might have unwanted side effects. The solution is to use use list of cells to divide as we did in the example. Notice that we call self.divideCellRandomOrientation(cell)function to divide cells. Other modes of division are available as well and they are shown in *examples\_PythonTutorial/steppableBasedMitosis/steppableBasedMitosisSteppables.py* as commented line with appropriate explanation. Notice MitosisSteppable inherits MitosisSteppableBase class (defined in *PySteppablesExamples.py*).It is is the base class which ensures that after we call any of the cell dividing function (e.g. divideCellRandomOrientation) CompuCell3D will automatically call updatAttributes function as well. updateAttributes function is very important and we must call it in order to ensure integrity and sanity of the simulation. During mitosis new cell is created (accessed in Python as childCell – defined in MitosisSteppableBase - self.mitosisSteppable.childCell) and as such this cell is uninitialized. It does have default attributes of a cell such as volume, surface (if we decide to use surface constraint or SurfaceTracker plugin) but all other parameters of such cell are set to default values. In our simulation we have been setting targetVolume and lambdaVolume individually for each cell. After mitosis childCell will need those parameters to be set as well. To make things more interesting, in our simulation we decided to change type of cell to be different than type of parent cell. In more complex simulations where cells have more attributes which are used in the simulation, we have to make sure that in the updateAttributes function childCell and its attributes get properly initialized. It is also very common practice to change attributes of parentCell after mitosis as well to account for the fact that parentCell is not the original parentCell from before the mitosis. **Important:** If you specify orientation vector for the mitosis the actual division will take place along the line/plane **perpendicular to this vector**. **Important:**the name of the function where we update attributes after mitosis has to be exactly updateAtttributes. If it is called differently CC3D will not call it automatically. We can of course call such function by hand, immediately we do the mitosis but this is not very elegant solution. Now we will discuss how to use PDE solvers in ComuCell3D. Most of the PDE solvers solve PDE with diffusive terms. Let's take a look at them **IX.3.5. AdvectionDiffusionSolver.** This steppable solves advection diffusion equation on a cell field as opposed to grid. Of course, the inaccuracies are bigger than in the case of PDE being solved on the grid but on the other hand solving the PDE on a cell field means that we associate cocentration with a given cell (not just with a lattice point). This means that as cells move so does the concentration. In other words we get advection for free. The mathematical treatment of this kind of approximation was spelled out in Phys. Rev. E 72, 041909 (2005) paper by D.Dan et al. The equation solved by this steppable is of the type: where denotes concentration , is diffusion constant, decay constant, is velocity field. In addition to just solving advection-diffusion equation this module allows users to specify secretion rates of the cells as well as different secretion modes. More about it in a moment. First let's see how one uses AdvectionDiffusionSolver: This is example syntax: FGF 0.05 0.003 flowFieldConcentration2D.txt Wall 0.5 0.3 Inside AdvectionDiffusionSolver you need to define sections that describe a field on which the steppable is to operate. In our case we declare just one diffusion field. Inside the diffusion field we specify sections describing diffusion and secretion. Let's take a look at DiffusionData section first: FGF 0.05 0.003 flowFieldConcentration2D.txt Wall We give a name (FGF) to the diffusion field – this is required as we will refer to this field in other modules. Next we specify diffusion constant and decay constant. **Caution:** We use Forward Euler Method to solve these equations. This is not a stable method for solving diffusion equation and we do not perform stability checks. If you enter too high diffusion constant for example you may end up with unstable (wrong) solution. Always test your parameters to make sure you are not in the unstable region. ConcentrationFileName is an optional tag and lets you specify a text file that contains values of concentration for every pixel. The value of concentratio of the last pixel read for a given cell becomes an overall value of concentration for a cell. That is if cell has, say 8 pixels, and you specify different concentration at every pixel, then cell concentration will be the last one read from the file. **Concentration file format** is as follows: *x y z c* where x,y,z, denote coordinate of the pixel. c is the value of the concentration. **Example:** 0 0 0 1.2 0 0 1 1.4 ... You may also specify cells which will not participate in the diffusion. You do it using tag. In this example you do not let any FGF diffuse into Wall cells. You may of course use as many as necessary tags . In addition to diffusion parameters we may specify how secretion should proceed. SecretionData section contains all the necessary information to tell CompuCell how to handle secretion. Let's study the example: 0.5 0.3 Here we have a definition two major secretion modes. Line: 0.5 ensures that every cell of type Fluid will get 0.5 increase in concentration every MCS. Line: 0.3 means that cells of type Fluid will get additional 0.3 increase in concentration but only when they touch cell of type Wall. This mode of secretion is called SecretionOnContact. **IX.3.6. FlexibleDiffusionSolver** This steppable is one of the basic and most important modules in CompuCell3D simulations. As the name suggests it is responsible for solving diffusion equation but in addition to this it also handles chemical secretion which maybe thought of as being part of general diffusion equation. where *k* is a decay constant of concentration *c* and *D* is the diffusion constant. The term called *secretion* has the meaning as described below. The principles of operations are analogous as in the case of AdvectionDiffusionSolver so most of has been said there applies to FlexibleDiffusionSolve. Also the syntax is very similar. Let's see an example FGF8 0.1 0.002 5 0.1 1.0 Bacteria 0.1 FGF 0.02 0.001 0.01 0.1 Bacteria 0.1 0.1 We can see new xml tags and . Their values determine the correspondence between MCS and actual time and between lattice spacing and actual spacing size. In this example for the first diffusion field one MCS corresponds to 0.1 units of actual time and lattice spacing is equal 1 unit of actual length. What is happening here is that the diffusion constant gets multiplied by: DeltaT/(DeltaX\* DeltaX) provided the decay constant is set to 0. If the decay constant is not zero DeltaT appears additionally in the term (in the explicit numerical approximation of the diffusion equation solution) containing decay constant so in this case it is more than simple diffusion constant rescaling. DeltaT and DeltaX settings are closely related to ExtraTimesPerMCS setting which allows calling of diffusion (and only diffusion) more than once per MCS. The number of extra calls per MCS is specified by the user on a per-field basis using ExtraTimesPerMCS tag. **IMPORTANT**: When using ExtraTimesPerMCS secretion functions will called only once per MCS. This is different than using PDESolverCaller where entire module is called multiple times (this include diffusion and secretion for all fields). **Instabilities of the Forward Euler Method** Most of the PDE soplvers in CC3D use Forward Euler exmplicit numerical scheme. This method is unstable for large diffusioni constant. As a matter of fact using D=0.25 with pulse initial condition will lead to instabilities in 2D. To deal with this you would normally use implicit solvers however due to moving boundary conditions that we have to deal with in CC3D simulations, memory requirements, perofmance and the fact that most diffusion constants encountered in biology are quite low (unfortunately this is not for all chemicals e.g. oxygen ) we decided to use explicit scheme. If you have to use large diffusion constants with explicit solvers you need to do rescaling: #. 1) Set D, t, x according to your model 2) If you will need to call solver multiple times per MCS. 3) Set to N-1 where: and SecretionData sections are analogous to those defined in AdvectionDiffusionSolver. here however, the secretion is done done on per-pixel basis (as opposed to per cell basis for AdvectionDiffusionSolver). For example when we use the following xml statement 0.1 this means that every pixel that belongs to cells of type Amoebae will get boost in concentration by 0.1. That is the secretion proceeds uniformly in the whole body of a cell. Alternative secretion mode would be the one described by the following line: 0.1 Here the secretion will take place in medium and only in those pixels belonging to Medium that touch directly Amoeba. More secretion schemes will be added in the future. The FlexibleDiffusionSolver is also capable of solving simple coupled diffusion type PDE of the form: where are coupling coefficients. To code the above equations in xml CompuCell3D syntax you need to use the following syntax: c 0.1 0.002 0.1 1.0 Bacteria 0.1 d 0.02 0.001 0.01 0.1 Bacteria 0.1 f 0.02 0.001 0.01 0.1 Bacteria 0.1 As one can see the only addition that is required to couple diffusion equations has simple syntax: **IX.3.7. FastDiffusionSolver2D** FastDiffusionSolver2DFE steppable is a simplified version of the FlexibleDiffusionSolverFE steppable. It runs several times faster that flexible solver but lacks some of its features. Typical syntax is shown below: FGF 0.010 0.003 2 Wall Demos/diffusion/diffusion\_2D\_fast\_box.pulse.txt In particular for fast solver you cannot specify cells into which diffusion is prohibited. However, you may specify cell types where diffusant decay is prohibited For exmplanation how ExtraTimesPerMCS works see section on FlexibleDiffusionSolverFE. **IX.3.8. KernelDiffusionSolver** This diffusion solver has the advantage over previous solvers that it can handle large diffusion constants. It is also stable. However, it does not accept options like or . It also requires periodic boundary conditions. Simply put KernelDiffusionSolver solves diffusion equation With fixed, periodic boundary conditions on the edges of the lattice. This is different from FlexibleDiffusionSolver where the boundary conditions evolve. You also need to choose a proper Kernel range (K) according to the value of diffusion constant. Usually when K2 e-(K^2 / (4D) ) is small (this is the main part of the integrand), the approximation convergers to the exact value. The syntax for this solver is as follows: 4 FGF 1.0 0.000 Demos/diffusion/diffusion\_2D.pulse.txt Inside tag one may also use option to For example: 4 2 FGF 1.0 0.000 Demos/diffusion/diffusion\_2D.pulse.txt **IX.3.9. ReactionDiffusionSolver** The reaction diffusion solver solves the following system of N reaction diffusion equations: Let's consider a simple example of such system: It can be coded as follows: F 0.010 Demos/diffusion/diffusion\_2D.pulse.txt -0.01\*H H 0.0 0.01\*F Notice how we implement functions *f* from the general system of reaction diffusion equations. We simply use tag and there we type arithmetic expression involving field names (tags ). In addition to this we may include in those expression word CellType. For example: 0.01\*F\*CellType This means that function *f* will depend also on CellType . CellType hodls the value of the type of the cell at particular location - x, y, z - of the lattice. The inclusion of the cell type might be useful if you want to use additional terms which may change depending of the cell type. Then all you have to do is to either use if statements inside or form equivalent mathematical expression using functions allowed by muParser (`http://muparser.sourceforge.net/mup\_features.html#idDef2 <%5Cl%20%22idDef2%22>`_) For example, let's assume that additional term for second equation is the following: In such a case additional term would be coded as follows: if (CellType==1,0.01\*F,0.15\*F) Notice that we have used here muParser function called if. The syntax of it is as follows: if(condition, expression if condition true, \\ expression if condition false) One thing to remember is that computing time of the additional term depends on the level of complexity of this term. Thus it is not the best idea to code very complex expressions using muParser. **IX.3.10. BoxWatcher Steppable** This steppable can potentially speed-up your simulation. Every MCS (or every Frequency MCS) it determines maximum and minimum coordinates of cells and then imposes slightly bigger box around cells and ensures that in the subsequent MCS spin flip attempts take place only inside this box containing cells (plus some amount of medium on the sides). Thus instead of sweeping entire lattice and attempting random spin flips CompuCell3D will only spend time trying flips inside the box. Depending on the simulation the performance gains are up to approx. 30%. The steppable will work best if you have simulation with cells localized in one region of the lattice with lots of empty space. The steppable will adjust box every MCS (or every Frequency MCS) according to evolving cellular pattern. The syntax is as follows: 5 5 5 All that is required is to specify amount of extra space (expressed in units of pixels) that needs to be added to the a tight box i.e. the box whose sides just touch most peripheral cells' pixels. **IX.4. Additional Plugins and Modules** Besides the modules that were introduced above CompuCell3D contains other modules which were developed to solve particular problem. For example module called DictyFieldInitializer is used to prepare initial cell configuration for the simulation of *Dictyostelium discoideum* morphogenesis based on the paper by P.Hogeweg and N.Savill **Modelling morphogenesis: from single cells to crawling slugs. J. theor. Biol. 184, 229-235.** Such modules have limited area of applicability and are mostly used in a single simulation. For this reason we will not describe them in more detail here. Interested user may consult CompuCell3D manual 3.2.0 where all such modules were described. It is our goal however to eliminate a need to write customized modules as much as possible. For example, DictyFieldInitializer can be easily replaced by using UniformInitializer and defining several regions there. Similarly Reaction diffusion solver for this simulation can be replaced by a general Reaction Diffusion solver described above. While we might run into performance issues when using general as opposed to customized, the flexibility and portability associated with using general use modules are worth extra run time. **X. References** 1. Bassingthwaighte, J. B. (2000) for the Physiome project*.* *Annals of* *Biomedical Engineering* **28**, 1043-1058. 2. Merks, R. M. H., Newman, S. A., and Glazier, J. A. (2004) oriented modeling of *in vitro*capillary development. *Lecture Notes in Computer Science* **3305**, 425-434. 3. Turing, A. M. (1953) Chemical Basis of Morphogenesis.*Philosophical Transactions of the Royal Society B***237**, 37-72. 4. Merks, R. M. H. and Glazier, J. A. (2005) cell-centered approach to developmental biology*.* *Physica A* **352**, 113-130. 5. Dormann, S. and Deutsch, A. (2002) of self-organized avascular tumor growth with a hybrid cellular automaton. *In Silico Biology* **2**, 1-14. 6. dos Reis, A. N., Mombach, J. C. M., Walter, M., and de Avila, L. F. (2003) interplay between cell adhesion and environment rigidity in the morphology of tumors*.* *Physica A* **322**, 546-554. 7. Drasdo, D. and Hohme, S. (2003) based approaches to birth and death in avascular tumors. *Mathematical and Computer Modelling* **37**, 1163-1175. 8. Holm, E. A., Glazier, J. A., Srolovitz, D. J., and Grest*,* G. S.**(1991) Effects of Lattice Anisotropy and Temperature on Domain Growth in the Two-Dimensional Potts Model. *Physical Review A* 43, 2662-2669. 9. Turner, S. and Sherratt, J. A. (2002) Intercellular adhesion and cancer invasion: A discrete simulation using the extended Potts model. *Journal of Theoretical Biology* **216**, 85-100. 10. Drasdo, D. and Forgacs, G. (2000) the interplay of generic and genetic mechanisms in cleavage, blastulation, and gastrulation*.* *Developmental Dynamics* **219**, 182-191. 11. Drasdo, D., Kree, R., and McCaskill, J. S. (1995) Carlo approach to tissue-cell populations. *Physical Review E* **52**, 6635-6657. 12. Longo, D., Peirce, S. M., Skalak, T. C., Davidson, L., Marsden, M., and Dzamba, B. (2004) Mcomputer simulation of morphogenesis: blastocoel roof thinning and matrix assembly in *Xenopus laevis*. *Developmental Biology* **271**, 210-222. 13. Collier, J. R., Monk, N. A. M., Maini, P. K., and Lewis, J. H. (1996) Pattern formation by lateral inhibition with feedback: A mathematical model of Delta-Notch intercellular signaling. *Journal of Theoretical Biology* **183**, 429-446. 14. Honda, H. and Mochizuki, A. (2002) Formation and maintenance of distinctive cell patterns by coexpression of membrane-bound ligands and their receptors. *Developmental Dynamics* **223**, 180-192. 15. Moreira, J. and Deutsch, A. (2005) Pigment pattern formation in zebrafish during late larval stages: A model based on local interactions. *Developmental Dynamics* **232**, 33-42. 16. Wearing, H. J., Owen, M. R., and Sherratt, J. A. (2000) Mathematical modelling of juxtacrine patterning. *Bulletin of Mathematical Biology* **62**, 293-320. 17. Zhdanov, V. P. and Kasemo, B. (2004) of the growth of neurospheres. *Europhysics Letters* **68**, 134-140. 18. Ambrosi, D., Gamba, A., and Serini, G. (2005) directional persistence and chemotaxis in vascular morphogenesis. *Bulletin of Mathematical Biology* **67**, 195-195. 19. Gamba, A., Ambrosi, D., Coniglio, A., de Candia, A., di Talia, S., Giraudo, E., Serini, G., Preziosi, L., and Bussolino, F. (2003) , morphogenesis, and Burgers dynamics in blood vessels formation*.* *Physical Review Letters* **90**, 118101. 20. Novak, B., Toth, A., Csikasz-Nagy, A., Gyorffy, B., Tyson, J. A., and Nasmyth, K. (1999) Finishing the cell cycle. *Journal of Theoretical Biology* **199**, 223-233. 21. Peirce, S. M., van Gieson, E. J., and Skalak, T. C. (2004) simulation predicts microvascular patterning and *in silico*tissue assembly*.* *FASEB Journal* **18**, 731-733. 22. Merks, R. M. H., Brodsky, S. V., Goligorksy, M. S., Newman, S. A., and Glazier, J. A. (2006) Cell elongation is key to *in silico* replication of *in vitro* vasculogenesis and subsequent remodeling.*Developmental Biology* **289**, 44-54. 23. Merks, R. M. H. and Glazier, J. A. (2005) Contact-inhibited chemotactic motility can drive both vasculogenesis and sprouting angiogenesis. *q-bio/0505033.* 24. Kesmir, C. and de Boer., R. J. (2003) spatial model of germinal center reactions: cellular adhesion based sorting of B cells results in efficient affinity maturation. *Journal of Theoretical Biology* **222**, 9-22. 25. Meyer-Hermann, M., Deutsch, A., and Or-Guil, M. (2001) probability and dynamical properties of germinal center reactions. *Journal of Theoretical Biology* **210**, 265-285. 26. Nguyen, B., Upadhyaya, A. van Oudenaarden, A., and Brenner, M. P. (2004) Elastic instability in growing yeast colonies. *Biophysical Journal* **86**, 2740-2747. 27. Walther, T., Reinsch, H., Grosse, A., Ostermann, K., Deutsch, A., and Bley, T. (2004) Mathematical modeling of regulatory mechanisms in yeast colony development. *Journal of Theoretical Biology* **229**, 327-338. 28. Borner, U., Deutsch, A., Reichenbach, H., and Bar, M. (2002) patterns in aggregates of *myxobacteria*arise from cell-cell collisions. *Physical Review Letters* **89**, 078101. 29. Bussemaker, H. J., Deutsch, A., and Geigant, E. (1997) field analysis of a dynamical phase transition in a cellular automaton model for collective motion. *Physical Review Letters* **78**, 5018-5021. 30. Dormann, S., Deutsch, A., and Lawniczak, A. T. (2001) analysis of Turing-like pattern formation in cellular automaton models. *Future Generation Computer Systems* **17**, 901-909. 31. Börner, U., Deutsch, A., Reichenbach, H., and Bär, M. (2002) Rippling patterns in aggregates of myxobacteria arise from cell-cell collisions. *Physical Review Letters* **89,** 078101. 32. Zhdanov, V. P. and Kasemo, B. (2004) of the growth and differentiation of stem cells on a heterogeneous scaffold*.* *Physical Chemistry Chemical Physics* **6**, 4347-4350. 33. Knewitz, M. A. and Mombach, J. C. (2006) Computer simulation of the influence of cellular adhesion on the morphology of the interface between tissues of proliferating and quiescent cells. *Computers in Biology and Medicine* **36**, 59-69. 34. Marée, A. F. M. and Hogeweg, P. (2001) amoeboids self-organize into a fruiting body: Multicellular coordination in Dictyostelium discoideum*.* *Proceedings of the National Academy of Sciences of the USA* **98**, 3879-3883. 35. Marée, A. F. M. and Hogeweg, P. (2002) Dictyostelium discoideum morphogenesis: the culmination. *Bulletin of Mathematical Biology* **64**, 327-353. 36. Marée, A. F. M., Panfilov, A. V., and Hogeweg, P. (1999) and thermotaxis of *Dictyostelium discoideum*slugs, a model study. *Journal of Theoretical Biology* **199**, 297-309. 37. Savill, N. J. and Hogeweg, P. (1997) morphogenesis: From single cells to crawling slugs. *Journal of Theoretical Biology* **184**, 229-235. 38. Hogeweg, P. (2000) mechanisms of morphogenesis: on the interplay between differential adhesion and cell differentiation. *Journal of Theoretical Biology***203**, 317-333. 39. Johnston, D. A. (1998) animals*.* *Journal of Physics A* **31**, 9405-9417. 40. Groenenboom, M. A. and Hogeweg, P. (2002) Space and the persistence of male-killing endosymbionts in insect populations. *Proceedings in Biological Sciences* **269**, 2509-2518. 41. Groenenboom, M. A., Maree, A. F., and Hogeweg, P. (2005) The RNA silencing pathway: the bits and pieces that matter. *PLoS Computational Biology***1**, 155-165. 42. Kesmir, C., van Noort, V., de Boer, R. J., and Hogeweg, P. (2003) Bioinformatic analysis of functional differences between the immunoproteasome and the constitutive proteasome. *Immunogenetics* **55**, 437-449. 43. Pagie, L. and Hogeweg, P. (2000) Individual- and population-based diversity in restriction-modification systems. *Bulletin of Mathematical Biology* **62**, 759-774. 44. Silva, H. S. and Martins, M. L. (2003) A cellular automata model for cell differentiation. *Physica A* **322**, 555-566. 45. Zajac, M., Jones, G. L., and Glazier, J. A. (2000) of convergent extension in animal morphogenesis. *Physical Review Letters* **85**, 2022-2025. 46. Zajac, M., Jones, G. L., and Glazier, J. A. (2003) convergent extension by way of anisotropic differential adhesion. *Journal of Theoretical Biology* **222**, 247-259. 47. Savill, N. J. and Sherratt, J. A. (2003) of epidermal stem cell clusters by Notch-mediated lateral induction. *Developmental Biology* **258**, 141-153. 48. Mombach, J. C. M, de Almeida, R. M. C., Thomas, G. L., Upadhyaya, A., and Glazier, J. A. (2001) Bursts and cavity formation in Hydra cells aggregates: experiments and simulations*. Physica A* **297**, 495-508. 49. Rieu, J. P., Upadhyaya, A., Glazier, J. A., Ouchi, N. B. and Sawada, Y. (2000) Diffusion and deformations of single hydra cells in cellular aggregates. *Biophysical Journal* **79**, 1903-1914. 50. Mochizuki, A. (2002) Pattern formation of the cone mosaic in the zebrafish retina: A cell rearrangement model. *Journal of Theoretical Biology* **215**, 345-361. 51. Takesue, A., Mochizuki, A., and Iwasa, Y. (1998) Cell-differentiation rules that generate regular mosaic patterns: Modelling motivated by cone mosaic formation in fish retina. *Journal of Theoretical Biology* **194**, 575-586. 52. Dallon, J., Sherratt, J., Maini, P. K., and Ferguson, M. (2000) Biological implications of a discrete mathematical model for collagen deposition and alignment in dermal wound repair. *IMA Journal of Mathematics Applied in Medicine and Biology* **17**, 379-393. 53. Maini, P. K., Olsen, L., and Sherratt, J. A. (2002) Mathematical models for cell-matrix interactions during dermal wound healing. *International Journal of Bifurcations and Chaos* **12**, 2021-2029. 54. Kreft, J. U., Picioreanu, C., Wimpenny, J. W. T., and van Loosdrecht, M. C. M. (2001) Individual-based modelling of biofilms. *Microbiology* **147**, 2897-2912. 55. Picioreanu, C., van Loosdrecht, M. C. M., and Heijnen, J. J. (2001) Two-dimensional model of biofilm detachment caused by internal stress from liquid flow. *Biotechnology and Bioengineering* **72**, 205-218. 56. van Loosdrecht, M. C. M., Heijnen, J. J., Eberl, H., Kreft, J., and Picioreanu, C. (2002) Mathematical modelling of biofilm structures. *Antonie Van Leeuwenhoek International Journal of General and Molecular Microbiology* **81**, 245-256. 57. Popławski, N. J., Shirinifard, A., Swat, M., and Glazier, J. A. (2008) Simulations of single-species bacterial-biofilm growth using the Glazier-Graner-Hogeweg model and the CompuCell3D modeling environment. *Mathematical Biosciences and Engineering* **5**, 355-388. 58. Chaturvedi, R., Huang, C., Izaguirre, J. A., Newman, S. A., Glazier, J. A., Alber, M. S. (2004) hybrid discrete-continuum model for 3-D skeletogenesis of the vertebrate limb*. Lecture Notes in Computer Science* **3305**, 543-552. 59. Popławski, N. J., Swat, M., Gens, J. S., and Glazier, J. A. (2007) Adhesion between cells, diffusion of growth factors, and elasticity of the AER produce the paddle shape of the chick limb. *Physica A* **373**, 521-532. 60. Glazier, J. A. and Weaire, D. (1992) The Kinetics of Cellular Patterns*.* *Journal of Physics: Condensed Matter* 4, 1867-1896. 61. Glazier, J. A. (1993) Grain Growth in Three Dimensions Depends on Grain Topology*.* *Physical Review Letters* 70, 2170-2173. 62. Glazier, J. A., Grest, G. S., and Anderson, M. P. (1990) Ideal Two-Dimensional Grain Growth. In *Simulation and Theory of Evolving Microstructures*, M. P. Anderson and A. D. Rollett, editors. The Minerals, Metals and Materials Society, Warrendale, PA, pp. 41-54. 63. Glazier, J. A., Anderson, M. P., and Grest, G. S. (1990) Coarsening in the Two-Dimensional Soap Froth and the Large-Q Potts Model: A Detailed Comparison*.* *Philosophical Magazine B***62**, -637*.* 64. Grest, G. S., Glazier, J. A., Anderson, M. P., Holm, E. A., and Srolovitz, D. J. (1992) Coarsening in Two-Dimensional Soap Froths and the Large-Q Potts Model. *Materials Research Society Symposium* 237, 101-112. 65. Jiang, Y. and Glazier, J. A. (1996) Extended Large-Q Potts Model Simulation of Foam Drainage. *Philosophical Magazine Letters* 74, 119-128. 66. Jiang, Y., Levine, H., and Glazier, J. A. (1998) Possible Cooperation of Differential Adhesion and Chemotaxis in Mound Formation of*Dictyostelium*. *Biophysical Journal* 75, 2615-2625. 67. Jiang, Y., Mombach, J. C. M., and Glazier, J. A. (1995) Grain Growth from Homogeneous Initial Conditions: Anomalous Grain Growth and Special Scaling States. *Physical Review E* 52, 3333-3336. 68. Jiang, Y., Swart, P. J., Saxena, A., Asipauskas, M., and Glazier, J. A. (1999) Hysteresis and Avalanches in Two-Dimensional Foam Rheology Simulations. *Physical Review E*59, 5819-5832. 69. Ling, S., Anderson, M. P., Grest, G. S., and Glazier, J. A. (1992) Comparison of Soap Froth and Simulation of Large-Q Potts Model. *Materials Science Forum* 94-96, 39-47. 70. Mombach, J. C. M. (2000) Universality of the threshold in the dynamics of biological cell sorting. *Physica A* **276**, 391-400. 71. Weaire, D. and Glazier, J. A. (1992) Modelling Grain Growth and Soap Froth Coarsening: Past, Present and Future. *Materials Science Forum* 94-96, 27-39. 72. Weaire, D., Bolton, F., Molho, P., and Glazier, J. A. (1991) Investigation of an Elementary Model for Magnetic Froth. *Journal of Physics: Condensed Matter*3, 2101-2113. 73. Glazer, J. A., Balter, A., Popławski, N. (2007) Magnetization to Morphogenesis: A Brief History of the Glazier-Graner-Hogeweg Model. In *Single-Cell-Based Models in Biology and Medicine*. Anderson, A. R. A., Chaplain, M. A. J., and Rejniak, K. A., editors. Birkhauser Verlag Basel, Switzerland. pp. 79-106. 74. Walther, T., Reinsch, H., Ostermann, K., Deutsch, A. and Bley, T. (2005) Coordinated growth of yeast colonies: experimental and mathematical analysis of possible regulatory mechanisms. *Engineering Life Sciences* **5**, 115-133. 75. Keller, E. F. and Segel., L. A. (1971) for chemotaxis*.* *Journal of Theoretical Biology* **30**, 225-234. 76. Glazier, J. A. and Upadhyaya, A. (1998) First Steps Towards a Comprehensive Model of Tissues, or: A Physicist Looks at Development. In *Dynamical Networks in Physics and Biology: At the Frontier of Physics and Biology*, D. Beysens and G. Forgacs editors. EDP Sciences/Springer Verlag, Berlin, pp. 149-160. 77. Glazier, J. A. and Graner, F. (1993) of the differential adhesion driven rearrangement of biological cells. *Physical Review E* **47**, 2128-2154. 78. Glazier, J. A. (1993) Cellular Patterns. *Bussei Kenkyu* 58, 608-612. 79. Glazier, J. A. (1996) Thermodynamics of Cell Sorting*. Bussei Kenkyu* 65, 691-700. 80. Glazier, J. A., Raphael, R. C., Graner, F., and Sawada, Y. (1995) The Energetics of Cell Sorting in Three Dimensions. In *Interplay of Genetic and Physical Processes in the Development of Biological Form*, D. Beysens, G. Forgacs, F. Gaill, editors. World Scientific Publishing Company, Singapore, pp. 54-66. 81. Graner, F. and Glazier, J. A. (1992) of biological cell sorting using a 2-dimensional extended Potts model. *Physical Review Letters* **69**, 2013-2016. 82. Mombach, J. C. M and Glazier, J. A. (1996) Single Cell Motion in Aggregates of Embryonic Cells. *Physical Review Letters* 76, 3032-3035. 83. Mombach, J. C. M., Glazier, J. A., Raphael, R. C., and Zajac, M. (1995) comparison between differential adhesion models and cell sorting in the presence and absence of fluctuations*.* *Physical Review Letters* **75**, 2244-2247. 84. Cipra, B. A. (1987) Introduction to the Ising-Model. *American Mathematical Monthly* **94**, 937-959. 85. Metropolis, N., Rosenbluth, A., Rosenbluth, M. N., Teller, A. H., and Teller, E. (1953) of state calculations by fast computing machines*.* *Journal of Chemical Physics* **21**, 1087-1092. 86. , G. and Newman, S. A. (2005). *Biological Physics of the Developing Embryo*. Cambridge Univ. Press, Cambridge. 87. Alber, M. S., Kiskowski, M. A., Glazier, J. A., and Jiang, Y. On cellular automation approaches to modeling biological cells*.* In*Mathematical Systems Theory in Biology, Communication and Finance*. J. Rosenthal, and D. S. Gilliam, editors. Springer-Verlag, New York, pp. 1-40. 88. Alber, M. S., Jiang, Y., and Kiskowski, M. A. (2004) gas cellular automation model for rippling and aggregation in *myxobacteria*. *Physica D* **191**, 343-358. 89. Novak, B., Toth, A., Csikasz-Nagy, A., Gyorffy, B., Tyson, J. A., and Nasmyth, K. (1999) Finishing the cell cycle*. Journal of Theoretical Biology* **199**, 223-233. 90. Upadhyaya, A., Rieu, J. P., Glazier, J. A., and Sawada, Y. (2001) Anomalous Diffusion in Two-Dimensional Hydra Cell Aggregates. *Physica A* 293, 549-558. 91. Cickovski, T., Aras, K., Alber, M. S., Izaguirre, J. A., Swat, M., Glazier, J. A., Merks, R. M. H., Glimm, T., Hentschel, H. G. E., Newman, S. A. (2007) From genes to organisms via the cell: a problem-solving environment for multicellular development. *Computers in Science and Engineering***9**, 50-60. 92. Izaguirre, J.A., Chaturvedi, R., Huang, C., Cickovski, T., Coffland, J., Thomas, G., Forgacs, G., Alber, M., Hentschel, G., Newman, S. A., and Glazier, J. A. (2004) CompuCell, a multi-model framework for simulation of morphogenesis. *Bioinformatics* **20**, 1129-1137. 93. Armstrong, P. B. and Armstrong, M. T. (1984) A role for fibronectin in cell sorting out. *Journal of Cell Science* **69**, 179-197. 94. Armstrong, P. B. and Parenti, D. (1972) Cell sorting in the presence of cytochalasin B. *Journal of Cell Science* **55**, 542-553. 95. Glazier, J. A. and Graner, F. (1993) Simulation of the differential adhesion driven rearrangement of biological cells. *Physical Review E* **47**, 2128-2154. 96. Glazier, J. A. and Graner, F. (1992) Simulation of biological cell sorting using a two-dimensional extended Potts model. *Physical Review Letters* **69**, 2013-2016. 97. Ward, P. A., Lepow, I. H., and Newman, L. J. (1968) Bacterial factors chemotactic for polymorphonuclear leukocytes. *American Journal of Pathology* **52**, 725-736. 98. Lutz, M. (1999) *Learning Python*. Sebastopol, CA: O’Reilly & Associates, Inc. 99. Balter, A. I., Glazier, J. A., and Perry, R. (2008) Probing soap-film friction with two-phase foam flow. *Philosophical Magazine, submitted.* 100. Dvorak, P., Dvorakova, D., and Hampl, A. (2006) Fibroblast growth factor signaling in embryonic and cancer stem cells. *FEBS Letters* **580**, 2869-2287. **Appendix** ***1. Calculating Inertia Tensor in CompuCell3D.*** For each cell the inertia tensor is defined as follows: where index '*i*' denotes *i*-th pixel of a given cell and *x**i**, y**i**,z**i* are coordinates of that pixel in a given coordinate frame. where index '*i*' denotes *i*-th pixel of a given cell and *x**i**, y**i**,z**i* are coordinates of that pixel in a given coordinate frame. In Figure 21 we show one possible coordinate frame in which one can calculate inertia tensor. If the coordinate frame is fixed calculating components of inertia tensor for cell gaining or losing one pixel is quite easy. We will be adding and subtracting terms like or . However, in CompuCell3D we are mostly interested in knowing tensor of inertia of a cell with respect to *xyz*coordinate frame with origin at the center of mass (*COM*) of a given cell as shown in Fig 21. Now, to calculate such tensor we cannot simply add or subtract terms like or to account for lost or gained pixel. If a cell gains or loses a pixel its COM coordinates change. If so then all the *x**i**, y**i**,z**i* coordinates that appear in the inertia tensor expression will have different value. Thus for each change in cell shape (gain or loss of pixel) we would have to recalculate inertia tensor from scratch. This would be quite time consuming and would require us to keep track of all the pixels belonging to a given cell. It turns out however that there is a better way of keeping track of inertia tensor for cells. We will be using parallel axis theorem to do the calculations. Paralel axis theorem states that if ICOM is a moment of inertia with respect to axis passing through center of mass then we can calculate moment of inertia with respect to any parallel axis to the one passin through the COM by using the following formula: where denotes moment of inertia with respect to *x* axis passing through center of mass, is a moment of inertia with respect to axis parallel to the *x* axis passing through center of mass, *d* is the distance between the axes and *M* is mass of the cell. Let us now draw a picture of a cell gaining one pixel: Now using parallel axis theorem we can write expression for the moment of inertia after cell gains one pixel the following that: where as before denotes moment of inertia of a cell with new pixel with respect to *x* axis passing through center of mass, is a moment of inertia with respect to axis parallel to the *x* axis passing through center of mass, *d**new* is the distance between the axes and *V+1* is volume of the cell after it gained one pixel. Now let us rewrite above equation by adding ad subtracting *Vd**2* term: Therefore we have found an expression for moment of inertia passing through the center of mass of the cell with additional pixel. Note that this expression involves moment of inertia but for the old cell (*i.e*. the original cell, not the one with extra pixel). When we add new pixel we know its coordinates and we can also easily calculate *d**new* .Thus when we need to calculate the moment of intertia for new cell instead of performing summation as given in the definition of the inertia tensor we can use much simpler expression. This was diagonal term of the inertia tensor. What about off-diagonal terms? Let us write explicitely expression for *I**xy* : where denotes *x* COM position of the cell, similarly denotes *y* COM position of cell and *V* denotes cell volume. In the above formula we have used the fact that and similarly for the *y* coordinate. Now, for the new cell with additional pixel we have the following relation: where we have added and subtracted to be able to form on the right hand side of the expression for . As it was the case for diagonal element, calculating off-diagonal of the inertia tensor involves and positions of center of mass of the cell before and after gaining new pixel. All those quantities are either known a priori () or can be easily calculated (center of mass position after gaining one pixel). Therefore we have shown how we can calculate tensor of inertia for a given cell with respect to a coordinate frame with origin at cell's center of mass, without evaluating full sums. Such "local" calculations greatly speed up simulations ***2.Calculating shape constraint of a cell – elongation term*** The shape of single cell immersed in medium and not subject to too drastic surface or surface constraints will be spherical (circular in 2D). However in certain situation we may want to use cells which are elongated along one of their body axes. To facilitate this we can place constraint on principal lengths of cell. In 2D it is sufficient to constrain one of the principal lenghths of cell how ever in 3D we need to constrain 2 out of 3 principal lengths. Our first task is to diagonalize inertia tensor (i.e. find a coordinate frame transformation which brings inertia tensor to a giagonal form) **2.1. Diagonalizing inertia tensor** We will consider here more difficult 3D case. The 2D case is described in detail in M.Zajac, G.L.jones, J,A,Glazier "*Simulating convergent extension by way of anisotropic differential adhesion"* Journal of Theoretical Biology **222** (2003) 247–259. In order to diagonalize inertia tensor we need to solve eigenvalue equation: or in full form The eigenvalue equation will be in the form of 3rd order polynomial. The roots of it are guaranteed to be real. The polynomial itself can be found either by explicit derivation, using symbolic calculation or simply in Wikipedia ( `http://en.wikipedia.org/wiki/Eigenvalue\_algorithm `_ ) so in our case the eigenvalue equation takes the form: This equation can be solved analytically, again we may use Wikipedia ( `http://en.wikipedia.org/wiki/Cubic\_function `_ ) Now, the eigenvalues found that way are principal moments of inertia of a cell. That is they are components of inertia tensor in a coordinate frame rotated in such a way that off-diagonal elements of inertia tensor are 0: In our cell shape constraint we will want to obtain ellipsoidal cells. Therefore the target tensor of inertia for the cell should be tensor if inertia for ellipsoid: where *a,b,c* are parameters describing the surface of an ellipsoid: In other words *a,b,c* are half lengths of principal axes (they are analogues of circle's radius) Now we can determine semi axes lengths in terms of principal moments of inertia by inverting the following set of equations: Once we have calculated semiaxes lengths in terms of moments of inertia we can plug –in actual numbers for moment of inertia (the ones for actual cell) and obtain lengths of semiexes. Next we apply quadratic constraint on largest (semimajor) and smallest (seminimor axes). This is what elongation plugin does. ***3 Forward Euler method for solving PDE's in CompuCell3D.*** In CompuCell3D most of the solvers uses explicit schemes (Forward Euler method) to obtain PDE solutions. Thus for the diffusion equation we have: In a discretetized form we may write: where to save space we used shorthand notation: and similarly for other coordinates. After rearranging terms we get the following expression: where the sum over index '*i*' goes over neighbors of point (*x,y,z*) and the neighbors will have the following concentrations:,,…, . ***4. Calculating center of mass when using periodic boundary conditions.*** When you are running calculation with periodic boundary condition you may end up with situation like in the figure below: Clearly what happens is that simply connected cell is wrapped around the lattice edge so part of it is in the region of high values of x coordinate and the other is in the region where x coordinates have low values. Consequently a naïve calculation of center of mass position according to : would result in being somewhere in the middle of the lattice and abviously outside the cell.A better procedure could be as follows: Before calculating center of mass when new pixel is added or lost we "shift" a cell and new pixel (gained or lost )to the middle of the lattice do calculations "in the middle of the lattice" and shift back. Now if after shifting back it turns out that center of mass of a cell lies outside lattice position it in the center of mass by applygin a shift equal to the length of the lattice and whose direction should be such that the center of mass of the cell ends up inside the lattice (there is only one such shift and it might be be equal to zero vector). This is how we do it using mathematical formulas: First we define shift vector as a vector difference between vector pointing to center of mass of the lattice and vector pointing to (approximately) the middle of the lattice. Next we shift cell to the middle of the lattice using : where denotes center of mass position of a cell after shifting but before adding or subtracting a pixel. Next we take into account the new pixel (either gained or lost) and calculate center of mass position (for the shifted cell): Above we have assumed that we are adding one pixel. Now all that we need to do is to shift back by same vector that brought cell to (approximately) center of the lattice. We are almost done. We still have to check if is inside the lattice. If this is not the case we need to shift it back to the lattice but now we are allowed to use only a vector whose components are multiples of lattice dimensions (and we can safely restrict to +1 and -1 multiples of the lattice dimmensions) . For example we may have: where , , are dimensions of the lattice. There is no cheating here. In the lattice with periodic boundary conditions you are allowed to shift point coordinates a vector whose components are multiples of lattice dimensions. All we need to do is to examine new center of mass position and form suitable vector . ***5. Dividing cluster cells*** While dividing non-clustered cells is straightforward, doing the same for clustered cells is more challenging. To divide non-cluster cell using directional mitosis algorithm we construct a line or a plane passing through center of mass of a cell and pixels of the cell (we are using PixelTracker plugin with mitosis) on one side of the line/plane end up in child cell and the rest stays in parent cell. The orientation of the line/plane can be either specified by the user or we can use CC3D built-in feature to calculate calculate orientation tion of principal axes and divide either along minor or major axis. With compartmental cells, things get more complicated because: 1) Compartmental cells are composed of many subcells. 2) There can be different topologies of clusters. Some clusters may look “snake-like” and some might be compactly packed blobs of subcells. The algorithm which we implemented in CC3D works in the following way: #. 1) We first construct a set of pixels containing every pixel belonging to a cluster cell. You may think of it as of a single “regular” cell. #. 2) We store volumes of compartments so that we know how big compartments shold be after mitosis (they will be half of original volume) #. 3) We calculate center of mass of entire cluster and calculate vector offsets between center of mass of a cluster and center of mass of particulat compartments as on the figure below: **Figure 25**. Vectors and show offsets between center of mass of a cluster and center of mass particular compartments. #. 4) We pick division line/plane and for parents and child cells we offsets between cluster center of mass (after mitosis) and center of masses of clusters. We do it according to the formula: where denotes offset after mitosis from center of mass of child (parent) clusters, is orientation vector before mitosis (see picture above) and is a normalized vector perpendicular to division line/plane. If we try to divide the cluster along dashed line as on the picture below **Figure 26**. Division of cell along dashed line. Notice the orientation of the offsets after the mitosis for child and parent cell will be and as expected because both parent and child cells will retain their heights but after mitosis will become twice narrower (cell with grey outer compartments is a parent cell): **Figure 27.** Child and parent (the one with grey outer compartments) cells after mitosis. The formula given above is heuristic. It gives fairly simple way of assigning pixels of child/parent clusters to cellular compartments. It is not perfect but the idea is to get approximate shape of the cell after the mitosis and as simulation runs cell shape will readjust based on constraints such as adhesion of focal point plasticity. Before continuing with the mitosis we check if center of masses of compartments belong to child/parent clusters. If the center of masses are outside their target pixels we abandon mitosis and wait for readjustment of cell shape at which point mitosis algorithm will pass this sanity check. For certain “exotic” shapes of cluster shapes presented mitosis algorithm may not work well or at all . In this case we would have to write specialized mitosis algorithm. #. 5) We divide clusters and knowing offsets from child/parent cluster center of mass we assign pixels to particular compartments. The assignment is based on the distance of particular pixel to center of masses of clusters. Pixel is assigned to particular compartment if its distance to the center of mass of the compartment is the smallest as compared to distances between centroids of other compartments. If given compartment has reached its target volume and other compartmets are underpopulated we would assign pixels to other compartments based on the closest distance criterion. Altohugh this method may result in some deviation from perfect 50-50 division of compartment volume in most cases after few MCS cells will readjust their volume. **Figure 28**. CC3D example of compartmental cell division. See also examples\_PythonTutorial/clusterMitosis. ***7. Command line options of CompuCell3D*** Although most users run CC3D using Player GUI sometimes it is very convenient to run CC3D using command line options. CC3D allows to invoke Player directly from command line which is convenient because if saves several clicks and if you run many simulations this might be quite convenient. **Remark:** On Windows we use .bat extension for run scripts and on Linux/OSX it is .sh. Otherwise all the material in this section applies to all the platforms. **7.1. CompuCell3D Player Command Line Options** The command line options for running simulation with the player are as follows: compucell3d.bat [options] Options are: -i - users specify simulation file they want to run. It can be either CC3DML (XML) configuration file or Python script. -s - name of the file containing description of screenshots to be taken with the simulation. Usually this file is prepared using Player by switching to different views, clickin camera button and saving screenshot description file from the Player File menu. -o - allows users to specify where screenshots will be written. Overrides default settings. --noOutput - instructs CC3D not to store any screenshots. Overrides Player settings. --exitWhenDone - instructs CC3D to exit at the end of simulation. Overrides Player settings. -h, --help - prints command line usage on the screen Example command may look like: compucell3d.bat –i Demos\\cellsort\_2D\\cellsort\_2D.xml –s screenshotDescription.sdfml –o Demos\\cellsort\_2D\\screenshot The frequency of the screenshots is read using Player settings so if you need to adjust it please use either GUI directly or change it using PlayerSettings plugin (see example *Demos\\cellsort\_2D\\cellsort\_2D\_PlayerSettings.xml*) **7.2. Runnig CompuCell3D in a GUI-Less Mode - Command Line Options.** Sometimes when you want to run CC3D on a cluster you will have to use runScript.bat which allows running CC3D simulations without invoking GUI. However, all the screenshots will be still stored. **Remark:** current version of this script does not handle properly relative paths so it has to be run from the installation directory of CC3D i.e. you have to cd into this directory prior to runnit runScript.bat. Another solution is to use full paths. The output of this script is in the form of vtk files which can be subsequently replayed in the Player (and one can take screenshots then). By default all fields present in the simulation are stored in the vtk file. If users want to remove some of the fields fro mbeing stored in the vtk format they have to pass this information in the Python script: CompuCellSetup.doNotOutputField(\_fieldName) The best place to put such stetements is directly before steppable section in the Python main script. See also commented line (and try uncommenting it) in *examples\_PythonTutorial\\ diffusion\_extra\_player\_field\\ diffusion\_2D\_extra\_player\_field.py*. Storing entire fields (as opposed to storing screenshots) preserves exact snapshots of the simulation and allows result postprocessing. In addition to the vtk files runScript stores lattice description file with .dml extension which users open in the Player (File->Open Lattice Description Summary File…) if they want to reply generated vtk files. The format of the command is: runScript.bat [options] The command line options for runScript.bat are as follows: -i - users specify simulation file they want to run. It can be either CC3DML (XML) configuration file or Python script. Remember about using full paths if you run the script from directory different than -c - allows users to specify core name for the vtk files. The default name for vtk files is “Step” -o - allows users to specify where vtk files and the .dml file will be written. Overrides default settings -f or –outputFrequency= - allows to specify how often vtk files are stored to the disk. Those files tend to be quite large for bigger simulations so storing them every single MCS (default setting) slows down simulation considerably and also uses a lot of disk space. --noOutput - instructs CC3D not to store any output. This option makes little sense in most cases. -h, --help - prints command line usage on the screen Example command may look as follows: runScript.bat –i examples\_PythonTutorial\\cellsort\_2D\_info\_printer\\ cellsort\_2D\_info\_printer.py –f 10 –o examples\_PythonTutorial\\cellsort\_2D\_info\_printer\\screenshots –c infoPrinter ***8. Managing CompuCell3D simulations (CC3D project files)*** Until version 3.6.0 CompuCell3D simulations were stored as a combination of Python, XML, and PIF files. This solution was working fine but there were significant problems with keeping track of simulations files. We still support this convention. However, starting with version 3.6.0 we introduced new way of managing CC3D simulations by enforcing that a single CC3D simulation is stored in a folder containing .cc3d project file describing simulation resources (.cc3d is in fact XML), such as XML configuration file, Python scripts, PIF files, Concentration filets etc… and a directory called Simulation where all the resources reside. The structure of the new-style CC3D simulation is presented in the diagram below: ->**CellsortDemo** CellsortDemo.cc3d **->Simulation** Cellsort.xml Cellsort.py CellsortSteppables.py Cellsort.piff FGF.txt Bold fonts denote folders. The benefit of using CC3D project files instead of loosely related files are as follows: #. 1) Previously users had to guess which file needs to be open in CC3D – XML or Python. While in a well written simulation one can link the files together in a way that when user opens either one the simulation would work but, nevertheless, such approach was clumsy and unreliable. Starting with 3.6.0 users open .cc3d file and they don’t have to stress out that CompUCell3D will complain with error message. #. 2) All the files specified in the .cc3d project files are copied to the result output directory along with simulation results (uncles you explicitely specify otherwise). Thus, when you run multiple simulations each one with different parameters, the copies of all XML and Python files are stored eliminating guessing which parameters were associated with particular simulations. #. 3) All file paths appearing in the simulation files are relative paths with respect to main simulation folder. This makes simulations portable because all simulation resources are contained withing single folder. In the example above when referring to Cellsort.piff file from Cellsort.xml you use “Simulation/ Cellsort.piff”. This effectively eliminates drawbacks of previous approach – when user one stores his simulations in Demos/cellsort and gives this simulation to his colleague who stores simulation in MySimulations/cellsort then second user will most likely see error message informing him that file “Demos/cellsort/cellsort.piff” was not found (I assume here that initial condition is specified using cellsort.piff). With approach based on relative paths such problems do not exist. Second user can put the simulation anywhere he wants and it will run just fine. #. 4) New style of storing CC3D simulations has also another advantage – it makes graphical management of simulation content and simulation generation very easy. As amatetr of fact new component of CC3D suite – Twedit++ - CC3D edition has a graphical tool that allows for easy project file management and it also has new simulation wizadrd which allows users to build template of CC3D simulation within less than a minute. Let’s now look in detail at the structure of .cc3d files: Simulation/Cellsort.xml Simulation/Cellsort.py Simulation/CellsortSteppables.py Simulation/Cellsort.piff Simulation/FGF.txt As you can see the structure of the file is quite flat. All that we are storing there is names of files that are used in the simulation. Two files have special tags which specifies name of the XML file storing “XML portion” of the simulation and which specifies main Python script. We have also PIFFile tag which is used to designate PIF files. All other files used in the simulation are referred to as Resources. For example Python steppable file is a resource of type “Python”. FGF.txt is aresource of type “Field”. Notice that all the files are specified using paths relative to main simulation directory. As we mentioned before, when you run .cc3d simulation all the files listed in the project file are copied to result folder. If for somereason oyu want to avoid coping of some of the files, simply add Copy=”No” attribute in the tag with file name specification. ***9. Keeping Track of Simulation Files (deprecated)*** CompuCell3D will store screenshots, vtk lattice snapshots and CC3DML file/Python main script in the output directory. However often simulations consist of several files: CC3DML, Python main script, Python steppable script, Python plugin script, PIF files etc. If you want those files to be archived with the rest of simulation output you need to use SimulationFileStorage steppable declared in *PySteppablesExamples*.py. The usage is very simple (see also *examples\_PythonTutorial\\cellsort\_2D\_info\_printer\\cellsort\_2D\_info\_printer.py*): from PySteppablesExamples import SimulationFileStorage sfs=SimulationFileStorage(\_simulator=sim,\_frequency=10) sfs.addFileNameToStore("\\ examples\_PythonTutorial/cellsort\_2D\_info\_printer/cellsort\_2D.xml") sfs.addFileNameToStore("\\ examples\_PythonTutorial/cellsort\_2D\_info\_printer/cellsort\_2D\_info\_printer.py") sfs.addFileNameToStore("examples\_PythonTutorial/cellsort\_2D\_info\_printer\\ /cellsort\_2D\_steppables\_info\_printer.py") steppableRegistry.registerSteppable(sfs) It wil ensure that files listed here will be writte to simulation directory. This way if you keep changing simulation files you will be able to easily recover entire simulation at