You are on page 1of 12

CHAPTER

Mathematical Modeling Techniques in Food and Bioprocesses: An Overview


Ashim K. Datta and Shyam S. Sablani

CONTENTS Mathematical Modeling ............................................................................................................2 Classication of Mathematical Modeling Techniques.............................................................3 Scope of the Handbook.............................................................................................................4 Short Overview of Models Presented in this Handbook..........................................................4 1.4.1 Physics-Based Models (Chapter 2 through Chapter 8) ................................................4 1.4.1.1 Molecular Dynamic Models ...........................................................................5 1.4.1.2 Lattice Boltzmann Models (Chapter 2) ..........................................................5 1.4.1.3 Continuum Models (Chapter 3 through Chapter 6) .......................................6 1.4.1.4 Kinetic Models (Chapter 7) ............................................................................6 1.4.1.5 Stochastic Models (Chapter 8)........................................................................6 1.4.2 Observation-Based Models (Chapter 9 through Chapter 15) .......................................7 1.4.2.1 Response Surface Methodology (Chapter 9)..................................................7 1.4.2.2 Multivariate Analysis (Chapter 10) ................................................................7 1.4.2.3 Data Mining (Chapter 11)...............................................................................9 1.4.2.4 Neural Network (Chapter 12) .........................................................................9 1.4.2.5 Genetic Algorithms (Chapter 13) ...................................................................9 1.4.2.6 Fractal Analysis (Chapter 14).........................................................................9 1.4.2.7 Fuzzy Logic (Chapter 15)...............................................................................9 1.4.3 Some Generic Modeling Techniques (Chapter 16 through Chapter 18)......................9 1.4.3.1 Monte-Carlo Technique (Chapter 16) ..........................................................10 1.4.3.2 Dimensional Analysis (Chapter 17)..............................................................10 1.4.3.3 Linear Programming (Chapter 18) ...............................................................10 1.4.4 Combining Models ......................................................................................................10 1.5 Characteristics of Food and Bioprocesses..............................................................................10 Acknowledgments............................................................................................................................11 References........................................................................................................................................11 1.1 1.2 1.3 1.4

1
q 2006 by Taylor & Francis Group, LLC

HANDBOOK OF FOOD AND BIOPROCESS MODELING TECHNIQUES

1.1 MATHEMATICAL MODELING A model is an analog of a physical reality, typically simpler and idealized. Models can be physical or mathematical and are created with the goal to gain insight into the reality in a more convenient way. A physical model can be a miniature, such as a benchtop version of an industrial scale piece of equipment. A mathematical model is a mathematical analog of the physical reality, describing the properties and features of a real system in terms of mathematical variables and operations. The phenomenal growth in the computing power and its associated user-friendliness

Need for understanding the detailed mechanisms Strong need Availability of time and resources, depending on the state of a-priori knowledge of the physics Available

Not really necessary

Constrained

Use fundamental laws to develop physics-based model

Obtain experimental data to develop observation-based model

Validate model against experimental data

Possibly validate against additional experimental data

Extract knowledge from the model using sensitivity analysis

Use model in optimization and control

Figure 1.1 A simple overview of model development and use.


q 2006 by Taylor & Francis Group, LLC

MATHEMATICAL MODELING TECHNIQUES IN FOOD AND BIOPROCESSES: AN OVERVIEW

have allowed models to be more realistic and have fueled rapid growth in the use of models in product, process, and equipment design and research. Many advantages of a model include (1) reduction of the number of experiments, thus reducing time and expenses; (2) providing great insight into the process (in case of a physics-based model) that may not even be possible with experimentation; (3) process optimization; (4) predictive capability, i.e., ways of performing what if scenarios; and (5) providing improved process automation and control capabilities. Mathematical models can be classied somewhat loosely depending on the starting point in making a model. In observation-based models, the starting point is the experimental data from which a model is built. It is primarily empirical in nature. In contrast, the starting point for physicsbased models is the universal physical laws that should describe the presumed physical phenomena. Physics-based models are also validated against experimental data, but in physics-based models the experimental data do not have to exist before the model. The decision on whether to build an observation-based or a physics-based model depends on a number of factors, including the need and available resources, as shown in Figure 1.1. After a model is built, its parameters can be varied to see their effectsthis process is termed parametric sensitivity analysis. A model can also be used to control a process. These conceptual steps are also shown in Figure 1.1.

1.2

CLASSIFICATION OF MATHEMATICAL MODELING TECHNIQUES

Classication of mathematical models can be in many different dimensions (Gershenfeld 1999), as shown in Figure 1.2. As implied in this gure, there is a continuum between the two extremes for any particular dimension noted in this gure. For example, it can be argued that even a model that is obviously physics-based, such as a uid ow in a porous media, has permeability as a parameter that is experimentally measured and is made up of many different parameters characterizing the porous matrix and the uid. It is possible to use a lattice Boltzmann simulation for the same physical process that will not need most of these matrix and uid parameters and, therefore, can be perceived as more fundamental. The chapters in this text cover much of the range shown in Figure 1.2 for any particular dimension. Physics-based (rst-principle-based) vs. data-driven models is the primary dimension along which the chapters are grouped. Scale of models is another dimension covered here. The lattice Boltzmann simulation in Chapter 2, for example, is at a smaller scale than the macroscale
Deterministic Numerical Macroscale

First-principle based

Data-driven

Microscale

Stochastic Analytical
Figure 1.2 Various dimensions of a model. This is not an exhaustive list.
q 2006 by Taylor & Francis Group, LLC

HANDBOOK OF FOOD AND BIOPROCESS MODELING TECHNIQUES

continuum models in Chapter 3 through Chapter 6. Another dimension is deterministic vs. stochastic. For example, the deterministic models in Chapter 3 through Chapter 6 can be made stochastic by following the discussion in Chapter 7. Analytical vs. numerical method of solution is another dimension of models. Numerical models have major advantages over analytical solution techniques in terms of being able to model more realistic situations. Thus, Chapter 2 through Chapter 5 cover mostly numerical solutions, although some references to analytical solutions are provided as well.

1.3 SCOPE OF THE HANDBOOK Each chapter in this book describes a particular modeling technique in the context of food and bioprocessing applications. Entire books have been written on each of the chapters in this handbook. However, these books are frequently not with food and bioprocess as the main focus. Also, no one book covers the breadth of modeling techniques included here. The motivation behind this handbook was to bring many different modeling techniques, as varied as physics-based and observation-based models, under one umbrella with food and bioprocess applications as the focus. Because the end goal of even very different modeling approaches, such as physics-based and observation-based models, can be the same (e.g., to understand and optimize the system), any two modeling techniques can be conceptually thought of as competing alternatives. This is more so in food and bioprocess applications in which the processes are complex enough that the superiority of any one type of modeling technique in an industrial scenario that demands quick answer is far from obvious. Another reason for discussing various models under one roof is that different types of models can be pooled to obtain models that combine the respective advantages. Succinct discussion of each model in the same context of food and bioprocess can help trigger such possibilities. The modeling techniques selected in the handbook are either already being used or have a great potential in food and bioprocess applications. Emphasis has been placed on how to formulate food and bioprocess problems using a particular modeling technique, away from the theory behind the technique. Thus, the chapters are generally structured to have a short introduction to the modeling technique, followed by the details on how that technique can be used in specic food and bioprocess applications. Although optimization is often one of the major goals in modeling, optimization itself is a broad topic that could not be included (with the exception of linear programming) in this text because of its extensive coverage of modeling, and the reader is referred to the excellent article by Banga et al. (2003).

1.4

SHORT OVERVIEW OF MODELS PRESENTED IN THIS HANDBOOK

A short description of each type of model presented in this handbook is presented in this section. There is no such thing as the best model because the choice of a model depends on a number of factors, the most obvious ones being the goal (whether to know the detailed physics), the modelers background (statistics vs. engineering or physics), and the time available (physics-based models typically take longer). Some of this is also noted in the schematic in Figure 1.1. 1.4.1 Physics-Based Models (Chapter 2 through Chapter 8)

Physics-based models follow from fundamental physical laws such as conservation of mass and energy and Newtons laws of motion; however, empirical (but fairly universal) rate laws are needed to apply the conservation laws at the macroscopic scale. For example, to obtain temperatures using a physics-based model, combine conservation of energy with Fouriers law (which is empirical)
q 2006 by Taylor & Francis Group, LLC

MATHEMATICAL MODELING TECHNIQUES IN FOOD AND BIOPROCESSES: AN OVERVIEW

of heat conduction. The biggest advantages of physics-based models are that they provide insight into the physical process in a manner that is more precise and more trustable (because we start from universal conservation laws), and the parameters in such models are measurable, often using available techniques. Physics-based models can be divided into three scales: molecular, macro, and meso (between molecular and macro). An example of a model at the molecular scale is the molecular dynamic model discussed later. Models such as the lattice Boltzmann model discussed in this book are in the mesoscale. Macroscopic models are the most common among physics-based models in food. Examples of macroscopic models are the commonly used continuum models of uid ow, heat transfer, and mass transfer. As we expand food and biological applications at micro- or nanoscale, such as in detection of microorganisms in a microuidic biosensor, scales will be approached where the continuum models in Chapter 2 through Chapter 5 will break down (Gad-el-Hak 2005). Similarly, in very short time scales, continuum assumption breaks down, and mesoscale or molecular scale models become necessary (Mitra et al. 1995). General discussion of models when continuum assumption breaks down can be seen in Tien et al. (1998). Physics-based models today are less common in food and bioprocessing product, process, and equipment design than in some manufacturing, such as automobile and aerospace. This can be primarily attributed to variability in biomaterials and the complexities of transformations that food and biomaterials undergo during processing; however, this scenario is changing as the appropriate computational tools are being developed. In fact, the physics-based model (such as computational uid dynamics, or CFD) is one of the areas in food process engineering experiencing rapid growth. 1.4.1.1 Molecular Dynamic Models Molecular dynamic (MD) models are physics-based models at the smallest scale. In its most rudimentary version, repelling force between pairs of atoms at close range and attractive force between them over a range of separations are represented in a potential function (such as Lennard Jones), for which there are many choices (Rapaport 2004). The spatial derivative of this potential function provides the corresponding force. Forces between one atom and a number of its neighbors are then added to obtain the combined force, and Newtons second law of motion is then used to obtain the acceleration from the force. This acceleration is then numerically integrated to obtain the trajectory describing the way the molecule would move. Physical properties of the system can be calculated as the appropriate time average over the trajectory, if it is of sufcient length. Although applications of molecular dynamics relevant to food processing (such as protein functionality and solution properties of carbohydrates) have been reported (Schmidt et al. 1994; Ueda et al. 1998), there appears to be very little ongoing work in applying MD to systems of direct relevance to food processing. Thus, MD has been excluded from this handbook. 1.4.1.2 Lattice Boltzmann Models (Chapter 2) The lattice Boltzmann (LB) method is physics-based, but at an intermediate scale (referred to as mesoscale) between the molecular dynamic model mentioned above that is at the microscale and continuum models mentioned below that are at the macroscale, where physical quantities are assumed to be continuous. LB is based on kinetic theory describing the dynamics of a large system of particles. The continuum assumption breaks down at some point going from the macroscale toward the microscale. Examples of such systems can be colloidal suspensions, polymer solutions, and ow-through porous media. This is where the lattice Boltzmann model is useful and is currently being pursued in relation to food processes.
q 2006 by Taylor & Francis Group, LLC

HANDBOOK OF FOOD AND BIOPROCESS MODELING TECHNIQUES

Other mesoscale simulations are also being used in food. For example, in Pugnaloni et al. (2005), large compression and expansion of viscoelastic protein lms are studied in relation to stability of foams and emulsions during formation and storage. 1.4.1.3 Continuum Models (Chapter 3 through Chapter 6) Continuum models presented in Chapter 3 through Chapter 6 primarily deal with transport phenomena, i.e., uid ow, heat transfer, and mass transfer. These physics-based models are based on fundamental physical laws. Typically, these models consist of a governing equation that describes the physics of the process along with equations that describe the condition at the boundary of the system. The conditions at the boundary determine how the system interacts with the surroundings. Mathematically, they are needed to obtain particular solutions of the governing equation. The solution of the combined governing equation-boundary condition system can be made as exact an analog of the physical system as desired by including as much detail of the physical processes as necessary. Physics-based models have several advantages over observation-based models: (1) they can be exact analogs of the physical process; (2) they allow in-depth understanding of the physical process as opposed to treating it as a black box; (3) they allow us to see the effect of changing parameters more easily; and (4) models of two different processes can share the same basic parameter (such as mass diffusivity and permeability measured for one process can be useful for other processes). The disadvantages of a physics-based model are as follows: (1) high level of specialized technical background is required; (2) generally more work is required to apply to real-life problems; and (3) often longer development time and more resources are needed. In the past 10 years or so, physics-based continuum models have really picked up because of the available powerful and user-friendly software. These software programs do have limitations, however, that apply to food related problems because of complexities in the process and signicant changes in the material due to processing. For example, rapid evaporation, as is true in baking, frying, and some drying operations, is hard to implement in most of these software. Also, these continuum models rely heavily on properties data that are only sparsely available for food systems. There are other physics-based continuum models for which more details could not be included because of the scope of this handbook. For example, electromagnetic heating of food such as microwave and radio frequency heating is modeled using the governing Maxwells equations, some details of which are provided in Chapter 3. Likewise, solid mechanics problems in food, such as during chewing, pufng, texture development, etc., are governed by the equations of solid mechanics, which also are not included in the book. 1.4.1.4 Kinetic Models (Chapter 7) Kinetic models mathematically describe rates of chemical or microbiological reactions. They generally can be considered to be physics-based. However, in complex chemical and microbiological processes, as is true for food and bioprocesses, the mechanisms are generally hard to obtain and are not always available. The kinetic models for such systems are more data-driven than fundamental (as could be true for simple systems). 1.4.1.5 Stochastic Models (Chapter 8) The physics-based continuum models have material properties that are typically measured. These models are often treated as deterministic ones, i.e., the parameter values are considered xed. However, due to biological and other sources of variability, these measured parameters can have random variations. For example, viscosity of a sample can have random variation because of
q 2006 by Taylor & Francis Group, LLC

MATHEMATICAL MODELING TECHNIQUES IN FOOD AND BIOPROCESSES: AN OVERVIEW

its biological variability. In a uid ow model that uses viscosity, the nal answer of interest, such as pressure drop, would also have the random uctuations corresponding to the random variations in viscosity. Inclusion of such random variations makes the physics-based models more realistic. Techniques to include such uncertainty are presented in Chapter 6 and Chapter 8. 1.4.2 Observation-Based Models (Chapter 9 through Chapter 15)

The physics-based modeling process described in part I assumes that a model is known, which is frequently difcult to achieve in complex processes. Although a physics-based model may also be adjusted based on measured data, observation-based models (see Figure 1.3) are inferred primarily from measured data. Observational models are black box models to different degrees in relation to the physics of the process. The classical statistical models can have a model in mind (often based on some understanding of the process) before obtaining the measured data. This makes them less of a black box than models such as neural network or genetic algorithm that are frequently completely data driven; no prior assumption is made about the model and no attempt is made to physically interpret the model parameters once the model is built. Loosely speaking, though, all observational models are referred to as data-driven models. For this handbook (Figure 1.3), we separate the classical statistical models from the rest of the observation-based models and refer to the rest as data-driven models. There are many practical situations in which time and resources do not permit a complete physics-based understanding of a process. Physics-based models often require more specialized training and/or longer development time. In some applications, detailed understanding provided by the physics-based model may not be necessary. For example, in process control, detailed physicsbased models often are not needed, and observation-based models can sufce. Observation-based models can be extremely powerful in providing a practical, useful relationship between input and output parameters for complex processes. The types of data available and the purpose of modeling usually inuence the kind of observation models to be used. General information on how to choose a model for a particular situation is hard to locate. An excellent Internet source guiding data-driven model choice and development can be seen in NIST (2005). Because observation-based models are built from data without necessarily considering the physics involved, use of such models beyond the range of data used (extrapolation) is more difcult than in the case of physics-based models. 1.4.2.1 Response Surface Methodology (Chapter 9) This is a statistical technique that uses regression analysis to develop a relationship between the input and output parameters by treating it as an optimization problem. The principle of experimental design is used to plan the experiments to obtain information in the most efcient manner. Using experimental design, the most signicant factors are found before doing the response surface and nding the optimum. This method is quite popular in food applications. It is important to note that nding the optimum using response surface is not limited to experimental data. Physics-based models can also be used to generate data that can be optimized using the response surface methodology similar to the method for experimental data (Qian and Zhang 2005). 1.4.2.2 Multivariate Analysis (Chapter 10) Multivariate analysis (MVA) is a collection of statistical procedures that involve observation and analysis of multiple measurements made on one or several samples of items. MVA techniques are classied in two categories: dependence and interdependence methods. In a dependence technique, the dependent variable is predicted or explained by independent variables. Interdependence methods are not used for prediction purposes and are aimed at interpreting the analysis output to opt
q 2006 by Taylor & Francis Group, LLC

Modeling of food and bioprocesses Part I Physics-based


2
Microscale Mol. Dynamics
Mesoscale (L. Boltzmann)

Part II Observation-based

Part III
Techniques that can be useful in either model

HANDBOOK OF FOOD AND BIOPROCESS MODELING TECHNIQUES

Macroscale continuum

Stochastic

Classical statistical

Data driven

Monte-Carlo

Dimensional analysis

Linear programming

16
Kinetics

17

18

Fluid flow

Data mining

11
Response 9 surface meth.

Neural network

12

Heat transfer Multivariate analysis

Genetic algorithm

13

10
Fractal analysis

Mass transfer

14

Heat & Mass transfer

Fuzzy logic

15

Figure 1.3 Various models presented in this handbook and their relationships.

q 2006 by Taylor & Francis Group, LLC

MATHEMATICAL MODELING TECHNIQUES IN FOOD AND BIOPROCESSES: AN OVERVIEW

for the best and most representative model. MVA is likely to be used in situations when one is not sure of the signicant factors and how they interact in a complex process. It is also a popular modeling process in food. 1.4.2.3 Data Mining (Chapter 11) Data mining refers to automatic searching of large volumes of data to establish relationships and identify patterns. To do this, data mining uses statistical techniques and other computing methodology such as machine learning and pattern recognition. Data mining techniques can also include neural network analysis and genetic algorithms. Thus, it can be seen as a meta tool that can combine a number of modeling tools. 1.4.2.4 Neural Network (Chapter 12) An articial neural network model (as opposed to a biological neural network) is an interconnected group of functions (equivalent to neurons or nerve cells in a biological system) that can represent complex inputoutput relationships. The power of neural networks lies in their ability to represent both linear and nonlinear relationships and in their ability to learn these relationships directly from the modeled data. Generally, large amounts of data are needed in the learning process. 1.4.2.5 Genetic Algorithms (Chapter 13) Genetic algorithms are search algorithms in a combinational optimization problem that mimick the mechanics of the biological evolution process based on genetic operators. Unlike other optimization techniques such as linear programming, genetic algorithms require little knowledge of the process itself. 1.4.2.6 Fractal Analysis (Chapter 14) Fractal analysis uses the concepts from fractal geometry. It has been primarily used to characterize surface microstructure (such as roughness) in foods and to relate properties such as texture, oil absorption in frying, or the Darcy permeability of a gel to the microstructure. Although fractal analysis may use some concepts from physics, the models developed are not rst principle-based. Processes governed by nonlinear dynamics can exhibit a chaotic behavior that can also be modeled by this procedure. Applications to food have been only sporadic. 1.4.2.7 Fuzzy Logic (Chapter 15) Fuzzy logic is derived from the fuzzy set theory that permits the gradual assessment of the membership of elements in relation to a set in contrast to the classical situation where an element strictly belongs or does not belong to a set. It seems to be most successful for the following: (1) complex models where understanding is strictly limited or quite judgmental; and (2) processes in which human reasoning and perception are involved. In food processing, the applications have been in computer vision to evaluate food quality, in process control, and in equipment selection. 1.4.3 Some Generic Modeling Techniques (Chapter 16 through Chapter 18)

Included in this part of the book are three generic modeling techniques that are somewhat universal and can be used in either physics-based or observation-based model building or for optimization once a model is built.
q 2006 by Taylor & Francis Group, LLC

10

HANDBOOK OF FOOD AND BIOPROCESS MODELING TECHNIQUES

1.4.3.1 Monte-Carlo Technique (Chapter 16) Monte Carlo refers to a generic approach whereby a probabilistic analog is set up for a mathematical problem, and the analog is solved by stochastic sampling. Chapter 7 shows the application of this technique to physics-based models. 1.4.3.2 Dimensional Analysis (Chapter 17) This is typically an intermediate step before developing mostly physics-based (but can be datadriven) models that is used to reduce the number of variables in a complex problem. This can reduce the computational or experimental complexity of a problem. 1.4.3.3 Linear Programming (Chapter 18) This is a well-known technique that is used for the optimization of linear models. It can be used in the context of a physics-based or a data-driven model. 1.4.4 Combining Models

Various modeling approaches can be combined to develop models that are even closer to reality and that have greater predictive power. For example, a physics-based model can be combined with an observation-based model by treating the output from the physics-based model as analogous to experimental data. See, for example, Eisenberg (2001) or work in a different application (Sudharsan and Ng 2000). Such a combined model is useful when only a portion of the system can be represented using a physics-based model or when the parameters in the physics-based model are uncertain. Two or more observation-based modeling techniques can also be combined (e.g., Panigrahi 1998), which is sometimes referred to as a hybrid model. A challenge, however, is to combine diverse methods in a seamless manner to provide a model that is easy to use.

1.5 CHARACTERISTICS OF FOOD AND BIOPROCESSES Some characteristics of food and bioprocesses are as follows: (1) they often involve drastic physical, chemical, and biological transformation of the material, during processing. Many of these transformations have not been characterized, primarily because of the following: (1) such a large variety of possible materials; (2) their biological origin, variabilities are signicant, even in the same material; (3) because the material contains large amounts of water, unless temperatures are low, there is always evaporation in the food matrix. This evaporation is hard to handle in physicsbased models and increases complexity of the process; and (4) many food processes involve coupling of different physics (e.g., microwave heating involves heat transfer and electromagnetics), thus compounding complexities. As novel processing technologies are introduced and combination technologies such as hurdle technology become more popular, complexities will only increase in the future. The industry in this area is characterized by a lower prot level and less room for drastic changes, than, for instance, automotive and aerospace industries. This translates to lower investment in research and development, which in turn leads to the generally lower level of technical sophistication as compared to other industries. Modeling, particularly physics-based modeling, often requires time and resources that are not available in the food industry. Consequently, with the exception of a handful of large multinational companies, modeling in general and physics-based modeling in particular are viewed as less critical and somewhat esoteric. It is expected that as the
q 2006 by Taylor & Francis Group, LLC

MATHEMATICAL MODELING TECHNIQUES IN FOOD AND BIOPROCESSES: AN OVERVIEW

11

computer technology continues to advance, modeling (particularly physics-based modeling) will become easier and perhaps more of a viable alternative in the industry.

ACKNOWLEDGMENTS Author Datta greatly acknowledges discussions with Professor James Booth of the Department of Biological Statistics and Computational Biology, Professor John Brady of the Department of Food Science, Professor Jean Hunter of the Department of Biological and Environmental Engineering, and Mr. Parthanil Roy of the School of Operations Research and Industrial Engineering, all of Cornell University. REFERENCES
Banga, J. R., Balsa-Canto, E., Moles, C. G., and Alonso, A. A., Improving food processing using modern optimization methods, Trends in Food Science and Technology, 14, 131144, 2003. Eisenberg, F. G., Virtual experiments using computational uid dynamics. Proceedings of 7th Conference on Food Engineering, American Institute of Chemical Engineers, New York, 2001. Gad-el-Hak, M., Liquids: The holy grail of microuidic modeling, Physics of Fluids, 17, 113, 2005. Gershenfeld, N., The Nature of Mathematical Modeling, Cambridge: Cambridge University Press, 1999. Mitra, K., Kumar, S., Vedavarz, A., and Moallemi, M. K., Experimental evidence of hyperbolic heat conduction in processed meat, Journal of Heat Transfer, Transactions of the ASME, 117(3), 568573, 1995. NIST. 2005. NIST/SEMATECH e-Handbook of Statistical Methods, http://www.itl.nist.gov/div898/handbook/ pmd/pmd.htm. Panigrahi, S., Neuro-fuzzy systems: Applications and potential in biology and agriculture, AI Applications, 12(13), 8395, 1998. Pugnaloni, L. A., Ettelaie, R., and Dickinson, E., Brownian dynamics simulation of adsorbed layers of interacting particles subjected to large extensional deformation, Journal of Colloid and Interface Science, 287, 401414, 2005. Qian, F. P. and Zhang, M. Y., Study of the natural vortex length of a cyclone with response surface methodology, Computers and Chemical Engineering, 29(10), 21552162, 2005. Rapaport, D. C., The Art of Molecular Dynamics Simulation, Cambridge: Cambridge University Press, 2004. Schmidt, R. K., Tasaki, K., and Brady, J. W., Computer modeling studies of the interaction of water with carbohydrates, Journal of Food Engineering, 22(14), 4357, 1994. Sudharsan, N. M. and Ng, E. Y. K., Parametric optimization for tumor identication bioheat equation using ANOVA and the Taguchi method. Proceedings of the IMechE, Part H, Journal of Engineering in Medicine, 214(H5), 505512, 2000. Tien, C.-L., Majumdar, A., and Gerner, F. M., Microscale Energy Transport, Washington, DC: Taylor & Francis, 1998. Ueda, K., Imamura, A., and Brady, J. W., Molecular dynamics simulation of a double-helical b-Carrageenan hexamer fragment in water, The Journal of Physical Chemistry A, 102(17), 27492758, 1998.

q 2006 by Taylor & Francis Group, LLC

You might also like