Sei sulla pagina 1di 18

Geomorphology 77 (2006) 217 234 www.elsevier.

com/locate/geomorph

Cellular automata as analysis and synthesis engines at the geomorphologyecology interface


Mark A. Fonstad
Department of Geography, Texas State University San Marcos, San Marcos, TX 78666-4616, United States Received 8 July 2004; received in revised form 9 June 2005; accepted 5 January 2006 Available online 7 February 2006

Abstract The linkages between ecology and geomorphology can be difficult to identify because of physical complexity and the limitations of the current theoretical representations in these two fields of study. Deep divisions between these disciplines are manifest in the methods used to simulate process, such as rigidly physical-deterministic methods for many aspects of geomorphology compared with purely stochastic simulations in many models of change in landcover. Practical and theoretical research into ecologygeomorphology linkages cannot wait for a single simulation schema which may never come; as a result, studies of these linkages often appear disjointed and inconsistent. The grid-based simulation framework for cellular automata (CA) allows simultaneous use of competing schemas. CA use in general geographic studies has been primarily limited to urban simulations models of change for land cover, both highly stochastic and/or expert rule-based. In the last decade, however, methods for describing physically deterministic systems in the CA framework have become much more accurate. The possibility now exists to merge separate CA simulations of different environmental systems into unified multiautomata models. Because CAs allow transition rules that are deterministic, probabilistic, or expert rule-based, they can immediately incorporate the existing knowledge rules in ecology and geomorphology. The explicitly spatial nature of CA provides a map-like framework that should allow a simple and deeply rooted connection with the mapping traditions of the geosciences and ecological sciences. 2006 Elsevier B.V. All rights reserved.
Keywords: Geomorphology; Ecology; Modeling; Simulation; Cellular automata

1. Introduction Cellular automata (CAs) are a class of numerical models based directly on a discrete spacetime grid, and they are a versatile representation of distributed dynamics that lend themselves well to the integration of geomorphology and ecology. This versatility results from a highly general form of simulation that uses local neighborhood operations to produce dynamic response
E-mail address: mfonstad@txstate.edu. 0169-555X/$ - see front matter 2006 Elsevier B.V. All rights reserved. doi:10.1016/j.geomorph.2006.01.006

from lookup tables, algebraic equations, stochastic probabilities, or expert rules centered on each discrete part of the grid. This article describes the historical development of the cellular automata approach, defines the major components of cellular automata models, focuses attention on the use of CA in geomorphology and ecology, and comments on some of the more advanced or subtle issues that underlie the use of CA in dynamic simulation. Deep down, at the heart of process description in geomorphology and ecology, lie the basic mental

218

M.A. Fonstad / Geomorphology 77 (2006) 217234

representations (schema) that allow conceptual understanding and quantification of environmental dynamics. Some of the most difficult and stubborn barriers to effectively integrating geomorphology and ecology are the vast differences in these conceptual schema. For example, the flow of water in a channel is usually described through continuum mechanics, a branch of physics that assumes the response of a fluid at macroscopic scales to be a continuum (non-particlelike), and uses this assumption to build several ahistorical laws (such as continuity and locality) of fluid motion that can then be used to solve fluid flow problems via the mathematics of partial differential equations. In contrast, models of the shifts of large-area plant species often take the form of regression equations, expert rules, discriminant equations, or other multivariate techniques. They are based on ideas such as the existence of discrete classes (species), probabilistic descriptions of transition (stochastic behavior), and the importance of previous occurrences (history) in affecting the outcome of transition. These two examples show the vast theoretical divides that must be crossed if effective interdisciplinary research is to take place, even if the systems under study are highly simplified. In ecology and geomorphology cries of anguish are heard over the lack of suitable data to analyze or synthesize. Such cries are often well-founded, but it is immediately apparent that both of these disciplines are far more theory-poor than data-poor. Indeed, in many ways, both of these disciplines are data-rich; think of the enormous amount of information contained in an air photo, a satellite image, or a digital elevation model. We have thousands of such images, but no theories in geomorphology or ecology can fully explain the patterns in any of them. In contrast, in some areas of physics (such as parts of cosmology), for example, almost no observations exist but vast understanding has been gained through a small number of theories. The lack of theory is especially apparent at the interface between geomorphology and ecology. Experimental reductionism can aid partially in this gap, as can analytical modeling, but a huge amount of theory development at the interface undoubtedly will come from numerical simulation. Sometimes resistance to simulation occurs in the development of understanding, because of the lack of exactitude that can come from numerical description, from the multiplicity of parameters that often accompany simulations, or from the reliance on computers. Nevertheless, the importance of simulation cannot be denied. Well-defined simulations allow users to ask

what if? questions, they allow direct observation and prediction of system dynamics, they can be used to observe otherwise unobservable spacetime scales, and they often generate unexpected or surprising results. Application of geomorphology and ecology to policy and decision-making often relies on simulation, so its importance stretches far beyond the analysis of hypothetical natural systems. Although the approach of cellular automata is decades old, its widespread development and acceptance across the natural sciences has been relatively recent and is rapidly accelerating. The time is approaching where CA models will become commonplace, and may become the archetypal description for certain kinds of systems. To efficiently and effectively include CA in the study of geomorphology and ecology, researchers should carefully consider and understand the strengths, weaknesses, current limitations, and future possibilities of this approach at a general level. 2. The origins and structure of the approach to cellular automata 2.1. Origins The origins of the approach of cellular automata stretch back to the 1950s when mathematician John Von Neumann devised the idea of discrete states in discrete spacetime in his search for ways to create selfreproducing structures. Fellow mathematician Stanislaw Ulam, who had worked with Von Neumann on the Manhattan Project (and who also invented the Monte Carlo simulation technique), made the idea much more practical by devising the idea of a two-dimensional lattice or grid where the distributed dynamics could take place. These grid-based dynamics were to only operate locally (locality); no global (top-down) dynamics were set by the modelers. The local dynamics were controlled by transition rules that were the same for every grid cell (universality or homogeneity). Time is viewed as discrete frames, as in a movie, where the states of all cells would update simultaneously according to the transition rules (simultaneous update or parallelism). The first popular application of cellular automata, and certainly the most well known CA application, was Cambridge mathematician John Conway's Game of Life (Gardner, 1970). Conway's original game asked players to place checkers on a checkerboard in any desired configuration. The game consisted of simple rules for these grid squares, whereby the checkers (which were alive) could live, die, or multiply depending on the states of neighbors. Some initial

M.A. Fonstad / Geomorphology 77 (2006) 217234

219

configurations were found to multiply and reform continuously, while others would die off quickly or produce only repetitive patterns. Life was converted quickly to digital computers so that highly complex collections of objects could interact with each other on a very large game board. The enormous richness of dynamics that can be produced in Life was a harbinger of things to come; indeed, the range of spatio-temporal dynamics allowed in even the simplest CAs is very similar to the range of allowed dynamics in the continuum-mechanical description of the universe (Toffoli, 1984). In the early 1980s, a veritable explosion occurred in research into the possibilities of the cellular automata approach. Mathematicians and physicists, in particular, began to take seriously the notion that a discrete representation of system dynamics may be preferable in many cases to a continuum description. Models of fluid response, based only on the single-velocity movement of particles (discrete on or off states) on a lattice with rules for what to do if a collision occurs with other particles, known as lattice gases, became much more sophisticated (d'Humieres et al., 1985; Frisch et al., 1986; Margolus et al., 1986; Salem and Wolfram, 1986) since invention in the mid-1970s (Hardy et al., 1976). Specialized computers (called cellular automata machines) built of enormous numbers of extremely simple processors, allowed much more efficient processing of models of cellular automata (Toffoli and Margolus, 1987). Wolfram (1994) began to explore systematically the overall dynamics of large classes of one-dimensional, single-state (on or off) cellular automata, finding some rules that generated completely random spacetime patterns despite being based completely on deterministic rules. In the geosciences as well, cellular automata began to be used as a simulation world-view. Tobler (1979) introduced the idea of cellular geography by showing how several distributed geographical concepts could be simulated through locally effective rules. His article became the foundation for a small industry of geographers who have designed simulations of urban growth and land cover change based on cellular automata, with probabilistic rules for state changes. In 1987, physicist Per Bak and his colleagues from Brookhaven National Laboratory developed the Sandpile CA (Bak et al., 1987); a simple 2-dimensional CA with integer states (blocks) that represented sand grains on a pile. By using local threshold rules (i.e. if more than three blocks are added to a cell, the blocks are distributed to the cardinal direction neighbor cells), the movements of material around the CA produced power

law magnitude/frequency relationships, similar to those found in many natural systems such as landslides. This initial Sandpile CA was enormously important in the developments of simulation capabilities in the past decade, as wave after wave of CA-based simulations have been proposed in many areas of geomorphology. The current research into simulations of CA in geomorphology and ecology is accelerating in many directions. The most obvious research direction has been the simple diffusion of CA models into many subfields of geomorphology and ecology. Other lines of research include questions such as (1) which formal representations of form and process can best be used in simulation, (2) how to develop transition rules, (3) what types of programming structures, software, and computer hardware best handle massive models of CA, (4) how to calibrate a model of CA to handle real time and real space scales, (5) how do simulations of CA relate to (or compete against) alternative techniques such as mean-field statistical approaches, expert systems, differential equations, and multi-agent systems, and (6) how do we validate the output of such complex simulations? 2.2. Structure The range of modeling approaches based on CA is certainly large and nebulous, but all CAs share certain commonalities. The first is the definition of the universe in CA; the model universe is discrete in space and time. In one-dimensional (1D) CA, this means that the universe is represented as a linear series of numbers. Two-dimensional (2D) CA, the most common in geomorphic and ecological simulations, are described as a grid (or lattice), much like a checkerboard, although the importance of the boundaries of these grids varies from simulation to simulation. Three-dimensional CA (3D) can be represented as a stack of two-dimensional grids (a spacetime cube). Discrete time slices in the CA universe can be added to the spatial descriptions by adding a dimension. Thus, 1D models through time can be described on a 2D grid, 2D simulations through time can be described in a 3D cube, and 3D automata models can be described by a stack of cubes, where each cube represents a time slice. Image processing provides an analogy for the computational power required for CA processing; 2D simulations are similar to processing a multispectral image per time step, and 3D simulations are similar to processing a hyperspectral image per time step. The tessellation of these spacetime grids is commonly depicted as a lattice of squares (cells), each

220

M.A. Fonstad / Geomorphology 77 (2006) 217234

with four cardinal neighbors and four diagonal neighbors. Although nearly all CA models are built upon such square-lattice tessellations, alternatives exist; triangular and hexagonal tessellations are also possible (Fig. 1). In particular, so-called latticegas and latticeBoltzmann CA techniques often employ hexagonal tessellations (Frisch et al., 1986). Nevertheless, the simplicity of the square-lattice approach and its similarity to the raster framework for spatial data make it the basis for further discussions. Although many authors do not mention explicitly what happens at the edges, or boundaries, of the CA grids, the specification of boundary conditions is important in the actual integration of a CA model with a computer (Parsons, 2004). All boundary conditions influence the response of the gridded dynamics to some degree. Many theoretical studies of CA simply use the no-boundary, or periodic, condition, where cells on one edge are neighbors with cells on the opposite edge. In a 2D lattice, this technically produces a three-dimensional torus, or donut, universe shape. Such a no-boundary condition can produce significant error in models of specific real-world examples, however, and is often replaced by some other boundary condition. For example, a periodic boundary CA grid containing a mountain with water flowing of one slope and off the grid edge would have that water strangely appear at the base of the opposite slope, certainly an unrealistic situation. An alternative includes reflecting (bounceback) boundaries, where no dynamics can go off of or come onto the grid through the boundary. Another possibility is the absorbing boundary, where material can flow off the grid, but the off-grid areas have no effect on the transition rules. One boundary condition, the no-slip condition stipulates that the cell states along a boundary do not change throughout the simulation of CA (Massalot and Chopard, 1998). Unfortunately, boundaries generally require an extra IFTHEN rule in the development of the CA transition

rules (IF the cell is a border cell, the cell state never changes, as an example). One way to avoid the effects of boundaries is to make the CA lattice much larger than the area where dynamics will be monitored, but this comes at the expense of computer time and resources. In classical cellular automata, the states of the cells are simple integers, such as 0, 1, 2, etc. Such a highly descriptive discrete state of the CA universe makes computation enormously simple and effectively restricts transition rules to expert (IFTHEN) rules and lookup tables (Wolfram, 1994). Many or even most CAs in use today, beginning with Kaneko (1985), however, allow floating point cell state values; the cell states are still technically discrete, because computers have some maximum floating point number length, but they allow algebraic and stochastic transition rules. Such floating point cell state CAs are commonly referred to as coupled-map lattices (CML), although the names coupled-lattice maps and continuous cellular automata are also often used to describe this class of CA (Wolfram, 1994). In the CA framework, dynamics are represented as a change in the state of grid cells from one time step to the following time step. The cell need not, however, necessarily change its state. What happens to each grid cell is defined by a transition rule or transition rules (Wolfram, 1994); these rules are the same for all cells in the lattice (rule universality, also known as homogeneity). If the transition rule requires that the state of a grid cell is only dependent on its state at a previous time step, such a model is called a Markov model, and is not considered a CA model. Cellular automata models have one additional feature: the transition rules operate on cells based on the local neighborhood of those cells, a property known as locality. For example, in a 2D grid, the state of a cell at time t + 1 could be a function of the states of the cells to the north, south, east, and west of the cell of interest at time t. This simple neighborhood, composed of the four cardinal neighbors, is called the

Fig. 1. Tesselations of two-dimensional space: (A) square grid; (B) triangular; (C) hexagonal.

M.A. Fonstad / Geomorphology 77 (2006) 217234

221

Von Neumann neighborhood. Other square-lattice neighborhoods include the Moore Neighborhood (all eight surrounding cells), and the extended form of these two neighborhoods (Fig. 2). In 1D CAs, the neighbors are simply the cells to the right and left (X direction) of the cell in question, in 2D CAs the neighbors in the Y direction (above and below) are added, and in the 3D case the neighbors on the Z axis above the XY plane and below the XY plane must also be added. The dependence of the transition of a cell upon the state of its neighbors directly encodes a form of spatial association into the system dynamics generated by a CA (Tobler, 1979). While this association does not necessarily lead to spatial autocorrelation, it often does, and provides a new measure (autocorrelation) of validating the results of a simulation. The form of this autocorrelation, or its nonexistence, is controlled by the form of the transition rules used in describing the dynamics of the CA system. The power of CA is its ability to handle different classes of transition rules. For example, a set of nested IFTHEN statements could be used to drive the dynamics (for example, IF all of my neighbors are of class tree, THEN the center cell either remains a tree or changes to a tree, ELSE the center cell does not change). This is an expert rule description of change, a common form of explanation in geomorphology and ecology. A set of expert rules can be encoded into a CA as a lookup table with predefined transition rules; a cell with only two tree neighbors might change to state X, one with three tree neighbors to state Y, and so on. In some situations, algebraic equations are used as the transition rule: the center cell at time t + 1 might be equal to the average of the neighbor cells at time t. Most simulations will incur a complex combination of these different types of transition rules, as well as including some level of nonspatial Markov influence, such as in CA models of differential (transformed into a Markovian difference equation) growth.

The set of transition rules used in CA can be entirely deterministic, relying only on the initial conditions of the grid and the neighboring cells states to produce the dynamics, or it can include a stochastic process. Stochastic CAs require the generation, for each time step, of a random number grid (for example, Bolliger et al., 2003). For example, a transition rule could be written as, if the random number is greater than 0.3, then change from state X to state Y, requiring a random number between 0 and 1 to work. Because of the effect of the random number grids, these stochastic CAs often require Monte Carlo-like simulations; the model must be run many times to come up with a probabilitydistribution for the overall spatio-temporal response of the system under study (Kronholm and Birkeland, 2004). Coupled map lattices and stochastic CAs are irreversible distributed systems (because of randomness or round off error), whereas true CAs (with only a few allowed discrete cell states) are, in principle, reversible. The transition rules define what happens to a cell from one time step to the next, but another important concept is that of parallelism, or simultaneous update; all cells are conceived to change at the exact same time. If some cells were to change first, the overall dynamics would be different than if other cells were to change first because of the effect of the local neighborhoods. Because CAs almost always run on standard series computers (which process one cell at a time), CA programmers generally include a buffer layer that holds the new cell states until all the cells have been updated, and then use this buffer layer in future calculations of the time step (Parsons, 2004). The student of cellular automata modeling finally may ask where do the transition rules come from? No simple answers exist to this question. The old modeling maxim holds true: a model should be made as simple as possible, but no simpler. Rule-making can proceed deductively from general rules (such as the conservation laws), or from induction (we went and measured

Fig. 2. Some common local neighborhoods for a square-tesselated lattice; (A) Von Neumann neighborhood; (B) Moore neighborhood; (C) extended Von Neumann neighborhood; (D) extended Moore neighborhood.

222

M.A. Fonstad / Geomorphology 77 (2006) 217234

empirical changes between these plant species), or some combination. Most elaborate CAs have induced and deduced rules. This ability to use either form of knowledge-building will allow future CA models to raid the past 100 years of geomorphic and ecological observations and theories to build and validate new simulations. 3. Cellular automata in ecology and geomorphology In ecology, some models of change are based upon deterministic rules, often upon continuous spatiotemporal rates of change. The LotkaVolterra model of predatorprey relationships is a commonly used example of differential equation based ecological modeling. Another classic example of this model basis is growth modeling applied to ecological organisms or populations. Based on controlled experiments and fundamental studies or organisms, models of growth are usually based on differential form equations. Simple exponential growth is rare, because of spatial limitations, food and energetics limits, and competition, so continuum models of growth incorporate a term or terms for carrying capacity. These models can vary in complexity from simple one-parameter exponential growth models (Verhuslt, 1838) to five-parameter generalized logistic models (Tsoularis and Wallace, 2002). These types of growth may be applied to the growth of an individual, or the growth of populations through space or time. Much more complex and dynamical systems can be formed by interconnecting these continuum models of individuals or groups into systems models (Odum, 1994). Simulations of these highly interconnected systems require recasting the differential equations into difference form to be solved at a finite time step. Nevertheless, growth in organisms or populations is very rarely a straightforward, continuous process; spatially varying rates of change, growth discontinuities caused by competition or death, fragmentation by external, non-biotic factors, or a wealth of other reasons may contribute to complexity. Alternatively, stochastic models of change, which include some form of probabilistic explanation of change, also provide explanations of ecological change in many situations. Simple stochastic models would include Markov models through time or space, or regression (linear, multiple linear, or logistic) to predict the presence, absence, or density of organisms in an area. Many other types of probabilistic models of ecological change exist; these are some of the common and simple tools.

The locality inherent in CAs and the flexibility of the model-builder's choice in including or not including certain rules provide an excellent rationale for using CA in ecology. Ecologists have applied CA to predator prey relationships (Tobin and Bjrnstad, 2003) and organism diffusion or spread (Zhou and Liebhold, 1995; Cannas et al., 1999; Seabloom et al., 2001, Alftine and Malanson, 2004). Also, researchers have developed cellular automata models of competition (Silvertown et al., 1992; Galam et al., 1998; Matsinos and Troumbis, 2002; van Wijk and Rodriguez-Iturbe, 2002), fragmentation and patch formation (Hogeweg, 1988; Bascompte and Sole, 1996; Keymer et al., 1998; Caswell and Etter, 1999; Favier et al., 2004), and large models built to contain many of these processes acting simultaneously (Green, 1982; Green et al., 1985; Satoh, 1989; Sato and Iwasa, 1993; Alonso and Sole, 2000; Messina and Walsh, 2001; Bolliger et al., 2003; Kupfer and Runkle, 2003). Hogeweg (1988) and Ermentrout and EdelsteinKeshet (1993) provide general overviews to the uses and considerations of cellular automata in ecological simulation. CA also provides a similar level of utility in producing simulations of the responses of forest fires. One of the very first CAs in the natural sciences was Bak et al.'s (1990) forest-fire CA, which produced similar magnitude-frequency distributions to those observed in natural forest fires. This simple CA has been much surpassed in recent years; forest fire CA simulations, developed by Clarke et al. (1994), Clar et al. (1996), and Karafyllidis and Thanailakis (1997), have improved on the basic Bak forest-fire CA by including more realistic variables and processes in the transition rules. Some of the early researchers in geomorphology to take advantage of the CA approach were researchers in Italy who built landslide and debris-flow simulations (Barca et al., 1986; Di Gregorio et al., 1994; Segre and Deangeli, 1995). Building on the foundations of Bak et al.'s (1987) original sandpile CA, these groups have made significant advances in representing mass movements in a cellular form (Avolio et al., 1999; Clerici and Perego, 2000; D'Ambrosio et al., 2003, in press; Datillo and Spezzano, 2003; Iovine et al., 2005), and have reached a level of predictive power that these simulations now are viable and important for public policy and decision-makers (Fig. 3). These simulations have since been extended to include lava flows (Crisci et al., 2003; Crosweller, 2003) and pyroclastic flows (Avolio et al., 2002). Smith (1991) provided a fascinating and unusual CA of surface erosion by using a vertical grid, with only one dimension in the horizontal plane but explicitly including vertical structure. Using simple transition

M.A. Fonstad / Geomorphology 77 (2006) 217234

223

Fig. 3. (A) Observed and (B) CA-simulated landslide paths, from D'Ambrosio et al. (2003).

rules, Smith was able to build quite realistic-looking patterns of differential erosion. Research in geomorphic CA simulation has expanded to eolian ripples (Anderson, 1990; Forrest and Haff, 1992; Anderson and Bunas, 1993; Zhang and Miao, 2003), and dune formation (Werner, 1995; Momiji et al., 2000). Werner and Fink (1993) and Ashton et al. (2001) have modeled coastal hydrodynamics and morphodynamics using the CA approach, and have connected these simulations to concepts of self-organization. The direct modeling of fluid and particle transport originally began in the 1980s by using lattice gases. These CAs have discrete particles moving from cell to cell at uniform speed. The transition rules in this approach are simple enough that large improvements, including three-dimensional space and increased physicality of the fluid movement, increased to the point where the flows are very similar to those deduced from the NavierStokes law (normally approached using partial differential equations). Nevertheless, latticegas and latticeBoltzmann approaches, with discrete particles, have not yet made a large impact on geomorphology. Bahr and Rundle (1995) have developed a model of glacial ice flow using the latticeBoltzmann approach, and Massalot and Chopard (1998) have built highly realistic simulations of water and particulate flow (1998), as well as the transport and deposition of snow. Jimenez-Hornero et al.'s (2003) work has extended the LatticeBoltzmann approach to the flow of water and sediment around complex obstacles, and Dupuis (2002) has completed a virtual river; this

LatticeBoltzmann reusable simulation produced highly realistic three-dimensional fluvial behavior. Despite the advancements made in the latticegas and latticeBoltzmann approach towards the simulation of geomorphic processes, nearly all geomorphic CA models are built using the coupled map lattice framework. Some of the most dramatic examples of CA modeling in geomorphology and ecology have been coupled map lattice models of river and watershed hydrology, fluvial sediment transport, and large-scale simulations of landscape evolution. In river and watershed CA hydrology, Thomas and Nicholas (2002), Thomas et al. (2002) and De Roo et al. (2000) have designed effective 2D simulations of steady, gradually varied flow for channels and floodplains, respectively. Parsons (2004) used time chaining approaches to simplify these runoff transition rules and to produce accurate 2D unsteady CA simulations of rainfallrunoff (Fig. 4). Brusik et al.'s (2003) hydrodynamic automaton is an unusual combination of CA and differential equation approaches to solving fluid flow that does produce realistic patterns of landform denudation. Other work in hydrology includes the movement of pollution in the soil water system (Di Gregorio et al., 1998; Bandini et al., 1999) and Kronholm and Birkeland's (2004) CA model of snowpack stability. The combination of CA approaches for fluid movement and sediment erosion, transport, and deposition has yielded important insights about the responses of rivers, watersheds, and landscape evolution. In

224

M.A. Fonstad / Geomorphology 77 (2006) 217234

Fig. 4. Three-dimensional views of Parsons (2004) CA rainfallrunoff model: (A) landcover patterns (later converted to roughness values); (B) digital elevation model; (CF) water depth during an ongoing rainstorm, with lighter colors representing deeper water (images made at time 0 min, 5 min, 20 min, and 30 min); (G) a three-dimensional exaggerated view of accumulated water depth at one time during the runoff period; (H) a 3D exaggerated view showing the distribution of infiltrated runoff water within the basin; (I) a pull-back showing that the CA simulation can be used over large areas.

particular, Murray and Paola's work on simulations of river planforms (Murray and Paola, 1994, 1996, 2003; Sapozhnikov et al., 1998) have been highly effective in showing how qualitatively different river responses are tied to specific transition rules that can be simple yet realistic. Further work in fluvial form simulations include Fonstad and Marcus's (2003) work on demonstrating self-organized criticality in riverbank dynamics and Favis-Mortlock et al.'s (2000) work simulating the development of rills. Brown et al.'s (2003) sediment transport CA focuses much more closely on local patterns of sediment sorting and movement in the fluvial

system in the search for larger patterns of material selforganization. At larger sizes, Coulthard et al.'s (1998) and Crave and Davy's (2001) CA simulations have highlighted the possibility of CA models to explore the large range of watershed geomorphic responses to long-term climate and land use changes. By coupling surface water CA models to subsurface flow and surface erosion transition rules, Luo's (Luo, 2001) LANDSAP simulations produce correctly scaled patterns of fluvial and groundwater sapping forms. Studies of long-term landscape evolution have benefited enormously from

M.A. Fonstad / Geomorphology 77 (2006) 217234

225

the CA approach, which allows fast and accurate simulations of repeated local erosion rules to digital elevation models. Chase's (1992), Rinaldo et al.'s (1992), and Rodriguez-Iturbe and Rinaldo's (1997) CA simulations of landscape evolution showed the temporal processes by which overall energy patterns were minimized and fractal networks emerged from local erosion rules. These CA models have been extended by Caldarelli (2001) and De Boer (2001), as well as Rebeiro-Hargrave (1999), who have explored the evolution of asymmetric drainage basins using cellular automata simulations. Although many geomorphic CAs allow vegetation and other layers of information to be used as boundary-

type information, very few CAs try to integrate ecological and geomorphic processes into viable cellular multiautomata. Two notable examples of interaction are the formation of shapes of meandering rivers by combining Murray and Paola's braided river CA with vegetation growth to strengthen the riverbanks (Murray and Paola, 2003). Moreover, efforts to combine earlier barchan dune (Baas, 2002) and ripple (Zhang and Miao, 2003) formation CAs with emergent vegetation growth produced highly realistic modified ridge-like forms (Fig. 5). Even these simulations which have explored the intricate relationships between these two groups of dynamics processes have barely scratched the surface in what is possible. Large-scale fully formed

Fig. 5. (A) A no vegetation, CA model of barchan dune formation; and (B) the dune formation CA modified by a rule to grow emergent vegetation through time (5, 10, 30, and 50 steps), modified from Baas (2002).

226

M.A. Fonstad / Geomorphology 77 (2006) 217234

CAs exist in both disciplines, and even simple, toy models with chained multiautomata will likely provide vast improvements in our understanding of system behavior. The model of bog patterning, put forward by Rietkerk et al. (2004), is a fairly complex combination of dense vascular plant growth coupled with spatial variable subsurface water flow. Both of these subsystems are based on a set of discretized differential equations, which allows for the emergence of banding patterns in bogs. The framework of numerical solution in two dimensions is considerably more complex that would likely be required if the same physical formulations were recast as a CA model in the mold of Murray and Paola (2003). Similarly, the model of dunefield activation and stabilization by Hugenholtz and Wolfe (2005) includes a differential equation framework for the changes in vegetation cover. This detailed model subsystem might be enhanced if the components of sand dynamics model were allowed to have spatial dynamics based on a CA model of sand dune growth and decay, a framework that would allow the existing difference-equation basis for the vegetation components. The topic of a geomorphology/ecology interface at a radically different scale is the influence of emergent vegetation in river channels. In this case, the river flows shape the spatial patterns of submerged plant community and partly alter its temporal dynamics, whereas the emergent vegetation contributes flow resistance and causes flow partitioning laterally and vertically. In this complex plantwater interaction, a latticegas or latticeBoltzmann approach might well allow a modeler to determine some of these relationships explicitly in three dimensions, from physical first principles rather than relying solely on measured values from flumes and streams. Because widespread CA modeling in the geosciences and ecological sciences is relatively new, its practitioners are often criticized for using the approach rather than some more traditional approach. Continuum mechanics, for example, is so engrained in the geosciences, that any alternative to strict differential equation formulations is often branded as arbitrary or non-advantageous. These views, while historically understandable, are in conflict with the very physical and mathematical sciences that produced continuum mechanics over the past 300 years. A huge number of recent physics articles in the past twenty years contain a wide range of modeling approaches, including differential equations, cellular automata, CMLs, multi-agent systems, and many others. For example, at the beginning of the period of major physics research into cellular automata, Toffoli (1984) carefully compared the

physicality of differential equation approaches in physics to automata approaches. He found that the automata approach, even with limited discrete states per cell, offered just as large a range (in phase space) of modeling possibilities, making it a viable alternative to differential equation methods. The nature of his comparison should be of utility to geoscience modelers as well. The great advantage of differential equations is that we have three centuries' experience with methods for their symbolic integration. As long as all computation had to be done by hand, it paid to stylize the physics in a certain direction so as to be able to handle the resulting mathematics. But few differential equations have closed-form solution anyway, and the past fifty years have seen numerical computation make bolder and bolder claims at being recognized as an essential part of mathematics. The moment one gives up symbolic manipulation as a major motive for using differential equations, one starts wondering whether one should still keep them as the starting point for numerical modeling. In fact, they lead to concrete numerical computation (e.g., as run on a general-purpose computer) that is at least three levels removed from the physical world that they try to represent. That is, first (a) we stylize physics into differential equations, then (b) we force these equations into the mold of discrete space and time and truncate the resulting power series, so as to arrive at finite-difference equations, and finally, in order to commit the later to algorithms, (c) we project real-valued variables onto finite computer words (round-off). At the end of the chain we find the computer again a physical system; isn't there a less roundabout way to make nature model itself? (Toffoli, 1984, p. 121) Certainly differential equations have enormous value in the simulation of geomorphology and ecology; most spatially explicit models are based on them. The critique here is not that differential equation methods are necessarily inappropriate or incorrect, but rather that they may not be necessary. Similarly, other schemas such as blind empiricism, statistical mechanics (as opposed to classical continuum mechanics), multi-agent systems, and other approaches do not necessarily provide the only (or even best) choice for process representation; model choice is fundamentally arbitrary, and the CA approach is simply another tool in the toolbox of choices.

M.A. Fonstad / Geomorphology 77 (2006) 217234

227

4. Advanced issues in CA implementation Despite the strengths and successes of the CA approach, important issues in implementation and theory need to be addressed if this schema will be useful in integrating geomorphic and ecologic modeling. Some of these issues can be identified, considered, and addressed early in the model-building process, others have rarely been considered, much less overcome. The structure of the CA universe as a lattice of locally connected sites sometimes can produce strange, unrealistic effects. For example, in some nondiffusive CAs where the number of possible cell states is low (for example the states could be 1 or 0), some dynamics will have preferred directions of motion. Such anisotropic response is the macroscopic realization of the inability of a simple lattice and simple neighborhood structure to allow dynamics in any direction (Frisch et al., 1986; Margolus et al., 1986). In strong anisotropic situations, the directional effects may be enough to invalidate the model completely. Much of the history of latticegas models is centered on finding a lattice where such anisotropies are minimized or removed altogether, so that the emergent fluid dynamics have the so-called Galilean invariance, where the direction of the lattice does not affect the dynamics of the CA. Many or even most complex CAs used in geomorphology and ecology avoid this problem because they contain some diffusive or probabilistic properties. Hence, even with a simple Von Neumann neighborhood, a CML fluid (Thomas and Nicholas, 2002; Thomas et al., 2002; Parsons, 2004) poured onto a cell (with no nearby obstructions) will appear to be shaped into a nearly circular fluid front within just a few time steps. Ideally, CAs would be constructed on a hexagonal lattice or one with an even higher degree of symmetry (requiring additional dimensions), but most simulations will continue to be constructed on square-grid lattices because of the inherent raster structure of computer graphics and image-processing software. Simple observation and testing after the construction of the model should identify potential anisotropy problems in a newly created CA. Another important issue in the development and application of CA simulations is the careful consideration of time. As mentioned in earlier sections, some models carefully consider what amount of real time is being spent during each time step or time slice in a CA simulation. Many current models do not. When qualitative response of single-CA systems is all that is required, non-real time steps function adequately and

are certainly the most straightforward method of modeling (Bak et al., 1987). This approach begins to have problems, however, when more than one CA model will be interlinked (multiautomata), or when researchers are interested in truly dynamic temporal processes (Favis-Mortlock et al., 2000). An excellent example of the problem of timing is the observation of diffusion in a coupled-map lattice fluid. According to the principle of locality, the amount of fluid (volume, for example) in a cell is a function of its neighbor cells in the previous time step. This would mean, therefore, that the rate of overall spread of the fluid will be one pixel per time step, regardless of the actual rules of volume transfer. Such a single-velocity response of diffusional flow is certainly not realistic at macroscopic scales. A certain amount of temporal control and understanding can be gained by scaling the observed rates of change in a CA to known rates of change in observed reality. Such a scaling procedure (for example, Murray and Paola, 2003) allows a certain level of coupling of different CA layers into a multiautomata. In systems where overall equilibrium response is the focus of study, scaling is an appropriate method for chaining processes together with the correct rates of change. If the system dynamics of interest are fast relative to the time it would take the process to affect the entire CA grid, however, equilibrium scaling may not be appropriate. As an alternative, the concept of time chaining may be applied. The effect of the response of one CA cell cannot, by definition, spread past its neighbors faster than one CA time step. Standard time-less CAs also show that the response effect of a cell cannot propagate slower than one cell per time step. The locality principle shows that the CA cannot propagate faster than one cell per time step, but no restrictions for developing rules would keep the dynamics from moving slower than one cell per time step. This is essentially the same computational issue as was early described by Courant et al. (1928) in the solution of partial differential equations. I term a method for slowing the propagation of a CA process to a desired rate time chaining, and the encoded rule for doing the slowing a timing chain (it has also been called phase locking, but I dislike the term because it suggests that processes are physically operating at the same rate, or phase). The simplest example of a timing chain is to consider a stochastic CA with one black cell surrounded by a huge lattice of white cells. An initial transition rule could be that a white cell to the right of a black cell will become black at the next time step, and any cell that is black will become white at the next step. The overall effect of this rule is to make the black square move, at

228

M.A. Fonstad / Geomorphology 77 (2006) 217234

one cell per time step, to the right. Now what if we wished to explicitly define time on this grid? Suppose we were modeling the movement of a particle (the black cell) through a matrix. We wished to specify that each time step was equal to one second. Also, the cells are all 1 cm wide. This means that, as previously described, the particle will always move 1 cm/s. Observation of a real system of particles moving through a matrix reveals, however, that this sized particle only moves at 0.4 cm/s. How do we include this observation? If we were only dealing with one size of particle, we could simply scale the grid so that the cells were 0.4 cm on a side, and the correct timing would be produced. Such a method is unwieldy, however, if we were to have more than one particle size in the model. Instead, we alter the original transition rule: the black cell moves to the right if (and only if) a randomly generated number for the black cell is greater than or equal to 0.4. Now we require a layer of random numbers (between 0 and 1) to be generated at each time step, but when they are (and the transition rule is applied) the effect is that the black cell, albeit in a rather jerky motion, is held back from moving the full one-cell-for-each-time-step by this stochastic rule that we have implemented. Averaged through time, the cell will appear to propagate at 0.4 cm/s. In this way, stochastic CAs can be time chained to any desired explicit length of time by altering the probability of transition accordingly. Time chaining of deterministic CAs, however, is slightly more challenging. If we have a grid with 1 m pixels, and a coupled-map lattice fluid on the grid, we might calculate, for example, that the fluid in a particular cell should move across that cell at the rate of 0.25 m/s. It does not make much physical sense to have a stochastic process determine the time chaining (i.e. the probability of motion across the pixel is 0.25). We know that the fluid should move over the pixel in four time steps. One approach is to have two additional states, or layers, for the cell. The first is a clock, or counter, that is reset to 0 when the fluid is removed from the cell, and increases its value by one at each time step that fluid occupies the cell. The second is an alarm clock which holds the value of the amount of time steps needed to hold the fluid before it can be moved to the next cell. In this example, the fluid speed is 0.25 m/s, and the cell size is 1 m, so we would calculate that the alarm clock needs to be set to 4, so that after four time steps (four clicks of the clock layer), we have an IF THEN rule that allows the fluid to move. This alarm clock method is essentially that used by D'Ambrosio et al. (2003) in their SCIDDICA CML models of debris flows, Favis-Mortlock et al.'s (2000) model of rill

initiation, and Parsons's (2004) CML model of the timing of rainfallrunoff. An important consideration in time chaining is the principle of locality, which will not allow processes to propagate faster than one cell per time step. If a CML fluid would be calculated to move at 10 m/s across a grid with 1 m pixels and time step of 1 s, serious inaccuracies will occur because the fluid will not be able to flow at the speed that is needed. Instead, a researcher needs to make a best-guess estimate of the maximum speed of propagation possible, and then adjust the explicit length of time devoted to each time step, or alternatively adjust the spatial size of the lattice cells. This problem is essentially the same as the wellknown CourantFriedrichsLewy Condition (or Criterion) in the solution of differential equation models of systems (Courant et al., 1928), where computational instabilities arise when processes are calculated to move faster over a space than is allowed by the amount of time per time step of computational iteration. Time chaining is an important consideration in using CAs to model real-world systems, and more research is needed to provide general chaining techniques for large classes of CAs. At the interface between geomorphology and ecology, widely varying rates of change occur in space and time between the interconnected or nested systems. In some cases, models can and should be designed to hold some variables constant and exogenous. This idea certainly is not new; Schumm and Lichty's (1965) discussion of time, space, and causality posits such regions of timespace where variables are independent or dependent. In many cases the mutual interaction of many variables, geomorphic and ecologic, however, must be considered. Perhaps a reasonable starting point for defining which variables to include as endogenous variables and how to correctly chain them to the rest of the model's timing is Phillips's (1999) information criterion in geomorphology. Much like the CFL criterion, the information criterion shows the mathematical relationship between the spatial scale, temporal scale, and rate of change of a system component or multiple system components. Known rates of change in model subsystems can be used when specifying the grid resolution of the model and time step interval, and conversely specified model resolution in time and space will place limits on the allowed rate of change of a subsystem. The power of the information criterion, in addition to its utility in time chaining the model, is exposing when two systems change at such vastly different spacetime rates that one or the other might well be modeled as an exogenous boundary condition.

M.A. Fonstad / Geomorphology 77 (2006) 217234

229

Verification of a CA model (does the model do what it is supposed to do) is an inherent part of the model construction, and it is not easy to separate verification from model validation (how well does the model simulate a real-world system). Nicholas (2005) points out that the two concepts are entangled in CA modeling because, for example, the grid structure of the CA is an implicit part of process parameterization. Indeed, it was a combination of normal verification procedures and later validation studies that found the original lattice gas models of fluid flow to have unexpected and unphysical anisotropic response because of the grid structure and nonsymmetrical neighborhoods used. Currently, no standard techniques exist for CA model verification because of the large range of conceptual schemas allowed by the CA framework and because of the entangled verification and validation concepts. An equally challenging problem in CA use is the correct use of a validation technique or techniques (Murray, 2003). In simple, 1D predictions using empirical approaches or differential equation approaches, many goodness-of-fit techniques exist such as the calculating the coefficient of determination or the index of agreement. For validating prediction of object classes without consideration of spatial accuracy, error matrices, or truth tables might be appropriate. Validating a two-dimensional or three-dimensional CA poses a much larger challenge. Nonspatial techniques such as the coefficient of determination are far too simplistic, because many patterns of prediction would give the same R2 value, and many of these potential patterns will be far more wrong than others. Also, either object-like classes or real-valued cells can exist in a CA, so choosing just truth tables or R2 measures is also oversimplistic. Geostatistical methods are one avenue likely to yield promising results in searching for useful validation tools, and they are widely available in overthe-counter GIS software packages. Also, if a model correctly produces an emergent property that is expected, but was not directly programmed into the model, this is also a reasonable, if qualitative, method of validation. An example would be the production of braids in rivers with the correct spatial scaling from simple rules of sediment and water flux (Murray and Paola, 1996, 1997; Sapozhnikov et al., 1998). Observing suitable spatial and temporal scaling emerging from a model is also an indication of validity, and researchers should, in the future, consider incorporating this type of validation. No simple approach seems to be perfect for validating all CAs, and the search for useful validation techniques will continue to be an important goal in modeling.

One of the most fascinating issues in modeling is the issue of material representation. The human brain seems innately wired to classify forms into named objects (hill/ valley/glacier/cloud/tree/forest), yet a great many mental and linguistic descriptions of degree also exist as well (higher/lower/steeper/rougher/smoother/more round). The archetypal representations in science are objects and fields, where the object is materially distinct from other objects (Mark et al., 1999), and processes happen to and between objects. In the field representation, only one object (the field) exists, but it varies in degree from one place to another and from one time to another. A contour map, for example, describes changes in the degree (elevation) of the topographic field. To make matters more confusing, models in the same discipline regularly use completely different representations to model the same thing. In geomorphology, for example, rivers can be represented as one-dimensional objects (lines, as they often are for meander-bend studies), twodimensional raster fields (as they are in CML models and remote sensing images), or a set of threedimensional objects (latticegas particles in lattice Boltzmann simulations). The issue of representation is bound up with issues of the theory-ladenness of observation and measurement at different scales (which can augment the field-like or object-like characteristics of a system). Efficient communication of CA models across interdisciplinary lines will require direct communication of representational choice, and researchers will need to accept that some degree of objectfield duality may influence model-building. The raster-like gridding and tessellation of CAs make them ideal for modeling the field representation of nature. Each cell is a position in spacetime that can be observed for dynamic changes. This type of view, the Eulerian frame of reference, is particularly suited to the field representation because objects move and grid cells may not. Objects can move about in CA models (for example, the aforementioned black cell in a sea of white cells), but the choice of CAs for a strict assemblage of objects seems unwieldy and inefficient. The multi-agent systems representation of nature (also known as particle systems or agent-based models) may provide a more reasonable approach to model a system of objects than can CAs, as it provides a view of a distinct object as it moves through spacetime (the Lagrangian frame of reference). Kessler and Werner's (2003) work on modeling structures of sorted patterned ground using large particle systems is an ideal example of a situation where CA would be an inelegant approach because the necessary representations are almost completely object-like.

230

M.A. Fonstad / Geomorphology 77 (2006) 217234

Approaches using cellular automata can be criticized, on occasion, because of the dependence of the locality principle. What about nonlocal or top-down influences? Ecological hierarchy theory postulates that influences can and must flow from top and bottom directions of influence, and both directions need to be addressed in any process description. Most CAs deal with top-down influences as external boundary conditions or as aspatial variables in the transition rules. External control programs can also change external or top-down conditions in lock-step with the CA model, but such control programs are often inelegant and unwieldy. Another approach might be to have interacting grids with different cell resolutions to simulate nonlocality yet maintain spatial elements. Regardless of the approach, CAs are an inherently bottom-up approach to modeling, and spatially variable top-down influences are difficult to represent efficiently in this schema. Finally, we come to the issue of computational design. Cellular automata have been previously designed in the raster-grid, image processing mold, or using an object-oriented approach. Although the parallelism inherent in CAs make them especially suitable for multiprocessor computers, nearly all CAs (except some latticegas designs) are run on normal series-type, single processor computers. To analyze which approaches are more suitable, more efficient, or easier to design is a task far beyond this article, and time will tell whether or not a standardized, general computing method for CAs should be implemented. 5. CA in teaching and general research In addition to the many pragmatic and useful benefits of simulating geomorphology and ecology using the cellular automata approach, CA models provide a conveniently simple approach to modelbuilding as an educational exercise. Linkages between geomorphology and ecology involve competing conceptual schema and different process rates of change that require explicit description in any model, and the CA framework provides an excellent educational forum for both issues. Also, the many examples in this article show that using CA as a modeling approach does not limit a model to one particular level of complexity. Toy models, like the Sandpile CA, are simple enough to be played by hand, and certainly in simple spreadsheet programs. Slightly more complicated CAs, like the Murray and Paola braided river CA, require an intermediate level of programming knowledge, or at least a fair competency

of using raster GIS or image processing software for modeling purposes. The most advanced CA models often involve many people, a high degree of programming competency, and a large amount of time to work through difficulties and to build in functionality. This progression, from toy models to fully designed realistic simulations, is a continuum that is well-suited to step-wise environmental science education. From an applied perspective, Malamud and Turcotte (2000) argue that CA provide an excellent way to understand and predict the response of natural hazards, making them a reasonable field of study for students. As students learn the fundamentals of model design, they may experience much of the excitement of discovery that is common in science; the importance of using toy models as a starting point in learning modeling is to show that complex response can arise from simple constructs, and also that fundamental insights can arise from only a modicum of modeling experience. Starting with simple models also encourages the frugality (Carpenter, 2003) of good modeling; a good model should be just complex enough to be realistic. Rule-based modeling is an important first step in simple ecosystem simulation (Starfeld, 1990). Also, the language of CA modeling is inclusive and useful in communicating ideas across traditionally isolating disciplines. Along with the general knowledge of CA theory, only the lack of simple, standardized tools for building CA models limits its immediate use in education. One of the most overriding themes of the past ten years of CA research in geomorphology and ecology has been the lack of standardization in modeling approaches or technology. To be sure, standardization can potentially remove some of the versatility and model understanding that serves the model-building community well. Pragmatically, however, the lack of standardized software for importing spatial data sets, setting CA boundary and neighborhood conditions, constructing and iterating the transition rules, and producing output that can be displayed, analyzed, and validated has probably hampered overall CA development in these sciences. Whoever develops a general CA modelbuilding package will have a significant influence in shaping the next period of theory-building in geomorphology and ecology. An analogy can be made with the period before STELLA was developed for simple development of differential equation models, or before GISs became widely available for individual research in the late 1980s and the early 1990s. Park and Wagner (1997) called for CAs to be coupled with GIS

M.A. Fonstad / Geomorphology 77 (2006) 217234

231

for maximum simulation viability. Some software are beginning to incorporate rudimentary CA interfaces, such as Idrisi, Erdas Imagine (and its Spatial Modeler), IDL, PCRaster, and ArcGIS, but CA modeling is not the primary goal of any of these packages, and full-fledged CA capabilities are still lacking in these programs. The use of cellular automata in model and theory development in geomorphology and ecology is just beginning to flourish as a subfield. Its developers are beginning to look at CA research as an independent field of study worthy of discussion and critical analysis (Malanson, 1999). The number of practitioners from across both of these immense fields of study has reached a critical point where general methods of model design have begun to leak from one subgroup to another. The greatest power of CA is likely its potential to link models that are currently isolated because of use on radically different schema (e.g. differential equations vs. expert rules vs. stochastic mechanisms). This survey of cellular automata in geomorphology and ecology suggests that the time has arrived to introduce CA theory as a general tool to advanced undergraduates and graduate students studying modeling, just as differential equations and empirical regression are taught in such forums today. Acknowledgements The author would like to acknowledge Mike Urban, Melinda Daniels, and Martin Doyle for their sponsorship of the Linking Geomorphology and Ecology session at the 2004 Annual Meeting of the AAG. Jay Parsons provided time-chained CA watershed simulations and valuable discussions on computer implementation of CA models. Andrew Marcus and Carl Legleiter proved to be ideal sounding boards during the recognition of many of the representational and implementational issues in this paper. Finally, Martin Doyle gave important advice on the utility of fast and frugal modeling in the training of students. Brad Murray and two anonymous reviewers helped significantly tighten the manuscript's structure and clarified many details. References
Alftine, K.J., Malanson, G.P., 2004. Directional positive feedback and pattern at an alpine tree line. Journal of Vegetation Science 15, 312. Alonso, D., Sole, R.V., 2000. DivGame: a cellular automata model of rainforest dynamics. Ecological Modelling 133 (12), 131141.

Anderson, R., 1990. Eolian ripples as examples of self-organization in geomorphological systems. Earth-Science Reviews 29 (14), 7796. Anderson, R.S., Bunas, K.L., 1993. Grain-size segregation and stratigraphy in aeolian ripples modeled with a cellular-automaton. Nature 365 (6448), 740743. Ashton, A., Murray, A.B., Arnoult, O., 2001. Formation of coastline features by large-scale instabilities induced by high-angle waves. Nature 414, 296300. Avolio, M.V., Di Gregario, S., Mantovani, F., Pasuto, A., Rongo, R., Silvano, S., Spataro, W., 1999. Hexagonal cellular automaton simulation of the 1992 Tessina landslide. In: Lippard, S.J., Naess, A., Sinding-Larsen, R. (Eds.), Proceedings of the IAMG '99, Trondheim, Norway, 291297. Avolio, M.V., Crisci, G.M., De Rosa, R., Di Gregorio, S., Larosa, S., Rongo, R., 2002. Pyroclastic flow simulation be a cellular automata model, Proceedings of the Congreso Internacional Ciencias de la Tierra 2002. Baas, A.C.W., 2002. Chaos, fractals and self-organization in coastal geomorphology: simulating dune landscapes in vegetated environments. Geomorphology 48 (1), 309328. Bahr, D.B., Rundle, J.B., 1995. Theory of lattice Boltzmann simulation of glacier flow. Journal of Glaciology 41 (139), 634640. Bak, P., Tang, C., Wiesenfeld, K., 1987. Self-organized criticality: an explanation of 1/f noise. Physical Review Letters 59 (4), 381384. Bak, P., Chen, K., Tang, C., 1990. A forest-fire model and some thoughts on turbulence. Physics Letters. A (147), 297300. Bandini, S., Mauri, G., Pavesi, G., Simone, C., 1999. A parallel model based on cellular automata for the simulation of pesticide percolation in the soil. Lecture Notes in Computer Science 1662, 383394. Barca, D., Di Gregorio, S., Nicoletta, F.P., Sorriso-Valvo, M., 1986. A cellular space model for flow-type landslides. In: Messina, G., Hamzda, M.H. (Eds.), Computers and their Application for Development, Proceedings of the International Symposium of the IASTED. Acta Press, Taormina, Italy, pp. 3032. Bascompte, J., Sole, R.V., 1996. Habitat fragmentation and extinction thresholds in spatially explicit models. Journal of Animal Ecology 65 (4), 465473. Bolliger, J., Sprott, J.C., Mladenoff, D.J., 2003. Self-organization and complexity in historical landscape patterns. Oikos 100, 541553. Brown, N.E., Ramirez, J.A., Wohl, E.E., 2003. Cellular automata models of particle interactions in sediment entrainment. Hydrology Days 2003, pp. 2432. Bursik, M., Martinez-Hackert, B., Delgado, H., Gonzalez-Huesca, A., 2003. A smoothed-particle hydrodynamic automaton of landform degradation by overland flow. Geomorphology 53 (12), 2544. Caldarelli, G., 2001. Cellular models for river networks. Physical Review E 63 (Article no. 021118). Cannas, S.A., Paes, S.A., Marco, D.E., 1999. Modeling plant spread in forest ecology using cellular automata. Computer Physics Communications 121, 131135. Carpenter, S.R., 2003. The need for fast and frugal models. In: Canham, C., Lauenroth, W., Cole, J.J. (Eds.), Understanding Ecosystems: The Role of Quantitative Models in Observations, Synthesis, and Prediction. Princeton University Press, Princeton. Caswell, H., Etter, R., 1999. Cellular automaton models for competition in patchy environments: facilitation, inhibition, and tolerance. Bulletin of Mathematical Biology 61, 625649.

232

M.A. Fonstad / Geomorphology 77 (2006) 217234 Favier, C., Chave, J., Fabing, A., Schwartz, D., Dubois, M.A., 2004. Modelling forestsavanna mosaic dynamics in man-influenced environments: effects of fire, climate and soil heterogeneity. Ecological Modelling 171 (12), 85102. Favis-Mortlock, D.T., Boardman, J., Parsons, A.J., Lascelles, B., 2000. Emergence and erosion: a model for rill initiation and development. Hydrological Processes 14 (1112), 21732205. Fonstad, M.A., Marcus, W.A., 2003. Self-organized criticality in riverbank systems. Annals of the Association of American Geographers 93 (2), 281296. Forrest, S.B., Haff, P.K., 1992. Mechanics of wind ripple stratigraphy. Science 255, 12401243. Frisch, U., Hasslacher, B., Pomeau, Y., 1986. Latticegas automata for the NavierStokes equation. Physical Review Letters 56 (14), 15051508. Galam, S., Chopard, B., Masselot, A., Droz, M., 1998. Competing species dynamics: qualitative advantage versus geography. European Physical Journal. B, Condensed Matter Physics (4), 529531. Gardner, M., 1970. The fantastic combinations of John Conway's new solitary game of Life. Scientific American 223, 120123. Green, D.G., 1982. Simulated effects of fire, dispersal and spatial pattern on competition within forest mosaics. Vegetation 82, 139154. Green, D.G., House, A.P.N., House, S.M., 1985. Simulating spatial patterns in forest ecosystems. Mathematics and Computers in Simulation 27, 191198. Hardy, J., de Pazzis, O., Pomeau, Y., 1976. Molecular dynamics of a classical lattice gas: transport properties and time correlation functions. Physical Review A 13 (5), 19491961. Hogeweg, P., 1988. Cellular automata as a paradigm for ecological modeling. Applied Mathematics and Computation 27 (1), 81100. Hugenholtz, C.H., Wolfe, S.A., 2005. Biogeomorphic model of dunefield activation and stabilization on the northern Great Plains. Geomorphology 70 (12), 5370. Iovine, G., D'Abrosio, D., Di Gregorio, S., 2005. Applying genetic algorithms for calibrating a hexagonal cellular automata model for the simulation of debris flows characterized by strong inertial effects. Geomorphology 66, 287303. Jimenez-Hornero, F.J., Giraldez, J.V., Laguna, A., 2003. A description of water and sediment flow in the presence of obstacles with a twodimensional, lattice BGK-cellular automata model. Water Resources Research 39 (12) (Article no. 1369). Kaneko, K., 1985. Spatiotemporal intermittency in coupled map lattices. Progress of Theoretical Physics 74 (5), 10331044. Karafyllidis, I., Thanailakis, A., 1997. A model for predicting forest fire spreading using cellular automata. Ecological Modelling 99 (1), 8797. Kessler, M.A., Werner, B.T., 2003. Self-organization of sorted patterned ground. Science 299 (5605), 380383. Keymer, J.E., Marquet, P.A., Johnson, A.R., 1998. Pattern formation in a patch occupancy model: a cellular automata approach. Journal of Theoretical Biology 194 (1), 7990. Kronholm, K., Birkeland, K.W., 2004. Relating spatial variability to snow stability using cellular automata models initialized with field data, International Snow Science Workshop 2004, Jackson Hole, WY, pp. Conference Abstract. Kupfer, J.A., Runkle, J.R., 2003. Edge-mediated effects on stand dynamic processes in forest interiors: a coupled field and simulation approach. Oikos 101 (1), 135146.

Chase, C.G., 1992. Fluvial landsculpting and the fractal dimension of topography. Geomorphology 5 (12), 3957. Clar, S., Drossel, B., Scwabl, F., 1996. Forest fires and other examples of self-organized criticality. Journal of Physics. Condensed Matter 8 (37), 68036824. Clarke, K.C., Brass, J.A., Riggan, P.J., 1994. A cellular-automaton model of wildfire propagation and extinction. Photogrammetric Engineering and Remote Sensing 60 (11), 13551367. Clerici, A., Perego, S., 2000. Simulation of the Parma River blockage by the Corniglio landslide (northern Italy). Geomorphology 33 (1 2), 123. Coulthard, T.J., Kirkby, M.J., Macklin, M.G., 1998. Non-linearity and spatial resolution in a cellular automaton model of a small upland basin. Hydrology and Earth System Studies 2, 257264. Courant, R., Friedrichs, K., Lewy, H., 1928. ber die partiellen Differenzengleichungen der mathematischen Physik. Mathematische Annalen 100, 3274. Crave, A., Davy, P., 2001. A stochastic precipiton model for simulating erosion/sedimentation dynamics. Computers & Geosciences 27 (7), 815827. Crisci, G.M., Di Gregorio, S., Rongo, R., Scarpelli, M., Spataro, W., Calvari, S., 2003. Revisiting the 1669 Etnean eruptive crisis using a cellular automata model and implications for volcanic hazard in the Catania area. Journal of Volcanology and Geothermal Research 123, 211230. Crosweller, H.S., 2003. Simulation of lava tube propagation using cellular automata. MRes dissertation, Lancaster University, Lancaster, UK, 50 pp. D'Ambrosio, D., Di Gregario, S., Iovine, G., Lupiano, V., Rongo, R., Spataro, W., 2003. First simulations of the Sarno debris flows through cellular automata modelling. Geomorphology 54 (12), 91117. D'Ambrosio, D., Spataro, W., Iovine, G., in press. Parallel genetic algorithms for optimising cellular automata models of natural complex phenomena: an application to debris-flows. Computers & Geosciences. Datillo, G., Spezzano, G., 2003. Simulation of a cellular landslide model with CAMELOT on high performance computers. Parallel Computing 29 (10), 14031418. De Boer, D.H., 2001. Self-organization in fluvial landscapes: sediment dynamics as an emergent property. Computers & Geosciences 27, 9951003. De Roo, A.P., Wesseling, C.G., Van Deursen, W.P., 2000. Physically based river basin modelling within a GIS: the LISFLOOD model. Hydrological Processes 14, 19811992. d'Humieres, D., Lallemand, P., Shimomura, T., 1985. An experimental study of lattice gas hydrodynamics. Preprint LA-UR-85-4051, Los Alamos National Laboratory, Los Alamos. Di Gregorio, S., Nicoletta, F.P., Rongo, R., Sorriso-Valvo, M., Spataro, W., 1994. A two-dimensional cellular automata model for landslide simulation. In: Gruber, R., Tommasini, M. (Eds.), PC '94, Proc. 6th Joint EPSAPS Int. Conf. on Physics Computing. Druck Partner Ruebelmann GmbH, Lugano, Switzerland, pp. 392407. Di Gregorio, S., Serra, R., Villani, M., 1998. Simulation of soil contamination and bioremediation by a cellular automaton model. Complex Systems 11 (1), 3154. Dupuis, A., 2002. From a lattice Boltzmann model to a parallel and reusable implementation of a virtual river. PhD dissertation thesis, University of Geneva, Geneva, 188 pp. Ermentrout, G.B., Edelstein-Keshet, L., 1993. Cellular automata approaches to biological modeling. Journal of Theoretical Biology 160, 97133.

M.A. Fonstad / Geomorphology 77 (2006) 217234 Luo, W., 2001. LANDSAP: a coupled surface and subsurface cellular automata model for landform simulation. Computers & Geosciences 27, 363367. Malamud, B.D., Turcotte, D.L., 2000. Cellular-automata models applied to natural hazards. IEEE Computing in Science & Engineering 4251. Malanson, G.P., 1999. Considering complexity. Annals of the Association of American Geographers 89 (4), 746753. Margolus, N., Toffoli, T., Vichniac, G., 1986. Cellular-automata supercomputers for fluid-dynamics modeling. Physical Review Letters 56 (16), 16941696. Mark, D.M., Smith, B., Tversky, B., 1999. Ontology and geographic objects: an empirical study of cognitive categorization. In: Freksa, C., Mark, D.M. (Eds.), Spatial Information Theory: A Theoretical Basis for GIS. Lecture Notes in Computer Sciences. SpringerVerlag, Berlin, pp. 283298. Massalot, A., Chopard, B., 1998. A lattice Boltzmann model for particle transport and deposition. Europhysics Letters 100 (6), 16. Matsinos, Y.G., Troumbis, A.Y., 2002. Modeling competition, dispersal and effects of disturbance in the dynamics of a grassland community using a cellular automaton model. Ecological Modelling 149 (12), 7183. Messina, J.P., Walsh, S.J., 2001. 2.5D Morphogenesis: modeling landuse and landcover dynamics in the Ecuadorian Amazon. Plant Ecology 156 (1), 7588. Momiji, H., Carretero-Gonzalez, R., Bishop, S.R., Warren, A., 2000. Simulation of the effect of wind speedup in the formation of transverse dune fields. Earth Surface Processes and Landforms 25 (8), 905918. Murray, A.B., 2003. Contrasting the goals, strategies, and predictions associated with simplified numerical models and detailed simulations. In: Iverson, R.M., Wilcock, P.R. (Eds.), Prediction in GeomorphologyAGU Geophysical Monograph, vol. 135, pp. 151165. Murray, A.B., Paola, C., 1994. A cellular model of braided rivers. Nature 371 (6492), 5457. Murray, A.B., Paola, C., 1996. A new quantitative test of geomorphic models, applied to a model of braided streams. Water Resources Research 32 (8), 25792587. Murray, A.B., Paola, C., 1997. Properties of a cellular braided-stream model. Earth Surface Processes and Landforms 22 (11), 10011025. Murray, A.B., Paola, C., 2003. Modelling the effect of vegetation on channel pattern in bedload rivers. Earth Surface Processes and Landforms 28, 131143. Nicholas, A.P., 2005. Cellular modeling in fluvial geomorphology. Earth Surface Processes and Landforms 30, 645649. Odum, H.T., 1994. Ecological and General Systems: An Introduction to Systems Ecology, Revised Edition. University Press of Colorado, Niwot, CO. 644 pp. Park, S., Wagner, D.F., 1997. Incorporating cellular automata simulators as analytical engines in GIS. Transactions in GIS 2 (3), 213231. Parsons, J., 2004. A computational cellular automaton for modeling surface water flow within the Rocky Mountain National Park. MAG thesis, Texas State University, San Marcos. Phillips, J.D., 1999. Earth Surface Systems: Order, Complexity, and Scale. Blackwell, Oxford, UK. 320 pp. Rebeiro-Hargrave, A., 1999. Large scale modelling of drainage evolution in tectonically active asymmetric basins using cellular automata. Zeitschrift fr Geomorphologie 118, 121134.

233

Rietkerk, M., Dekker, S.C., Wassen, M.J., Verkroost, A.W.M., Bierkens, M.F.P., 2004. A putative mechanism for bog patterning. The American Naturalist 163 (5), 699708. Rinaldo, A., Rodriguez-Iturbe, I., Ringon, R., Bras, R.L., IjjaszVasquez, E.J., Mariani, A., 1992. Minimum energy and fractal structures of drainage networks. Water Resources Research 28, 21832195. Rodriguez-Iturbe, I., Rinaldo, A., 1997. Fractal River Basins: Chance and Self-Organization. Cambridge University Press, Cambridge, NY. Salem, J., Wolfram, S., 1986. Thermodynamics and hydrodynamics with cellular automata. In: Wolfram, S. (Ed.), Theory and Applications of Cellular Automata. World Scientific, pp. 362365. Sapozhnikov, V.B., Murray, A.B., Paola, C., Foufoula-Georgiou, E., 1998. Validation of braided-stream models: spatial state-scale plots, self-affine scaling, and island shapes. Water Resources Research 34, 23532364. Sato, K., Iwasa, Y., 1993. Modeling wave regeneration in subalpine Abies forests: population dynamics with spatial structure. Ecology 74, 15381550. Satoh, K., 1989. Computer experiment on the complex behavior of a two-dimensional cellular automaton as a phenomenological model for an ecosystem. Journal of the Physical Society of Japan 58 (10), 38423856. Schumm, S.A., Lichty, R.W., 1965. Time, space, and causality in geomorphology. American Journal of Science 263, 110119. Seabloom, E.W., Moloney, K.A., van der Valk, A.G., 2001. Constraints on the establishments of plants along a fluctuating water-depth gradient. Ecology 82 (8), 22162232. Segre, E., Deangeli, C., 1995. Cellular automaton for realistic modelling of landslides. Nonlinear Processes in Geophysics 2, 115. Silvertown, J., Holtier, S., Johnson, J., Dale, P., 1992. Cellular automaton models of interspecific competition for space the effect of pattern on process. Journal of Ecology 80, 527534. Smith, R., 1991. The application of cellular automata to the erosion of landforms. Earth Surface Processes and Landforms 16 (3), 273281. Starfeld, A.M., 1990. Qualitative, rule-based modeling. BioScience 40 (8), 601604. Thomas, R., Nicholas, A.P., 2002. Simulation of braided river flow using a new cellular routing scheme. Geomorphology 43, 179195. Thomas, R., Nicholas, A.P., Quine, T.A., 2002. Development and application of a cellular model to simulate braided river processform interactions and morphological change. In: Bousmar, D., Zech, Y. (Eds.), River Flow 2002. Balkema, Rotterdam, pp. 783792. Tobin, P.C., Bjrnstad, O.N., 2003. Spatial dynamics and crosscorrelation in a transient predatorprey system. Journal of Animal Ecology 72, 460467. Tobler, W.R., 1979. Cellular geography. In: Gale, S., Olsson, G. (Eds.), Philosophy in Geography. D. Reidel, Dordrecht, The Netherlands, pp. 279386. Toffoli, T., 1984. Cellular automata as an alternative to (rather than an approximation of) differential equations in modeling physics. Physica. D 10, 117127. Toffoli, T., Margolus, N., 1987. Cellular Automata Machines. MIT Press, Cambridge, MA. 259 pp. Tsoularis, A., Wallace, J., 2002. Analysis of logistic growth models. Mathematical Biosciences 179, 2155.

234

M.A. Fonstad / Geomorphology 77 (2006) 217234 Wolfram, S., 1994. Cellular Automata and Complexity: Collected Papers. Addison-Wesley, Reading, MA. Zhang, Q., Miao, T., 2003. Aeolian sand ripples around plants. Physical Review E 67 (051304), 14. Zhou, G., Liebhold, A.M., 1995. Forecasting the spatial dynamics of gypsy moth outbreaks using cellular transition models. Landscape Ecology 10 (3), 177189.

van Wijk, M.T., Rodriguez-Iturbe, I., 2002. Treegrass competition in space and time: insights from a simple cellular automata based on ecohydrological dynamics. Water Resources Research 38 (2), 18.118.15. Verhuslt, P.F., 1838. Notice sur la loi que la population suit dans son accrossement. Correspondence Mathmatique et Physique 10, 113121. Werner, B.T., 1995. Eolian dunes: computer simulation and attractor interpretation. Geology 23, 11071110. Werner, B.T., Fink, T.M., 1993. Beach cusps as self-organized patterns. Science 260, 968971.

Potrebbero piacerti anche