Fact-checked by Grok 2 weeks ago

Physical system

A physical system is a portion of the physical selected for study, comprising a collection of interacting material entities such as particles, fields, or objects, distinct from its surroundings known as the . This demarcation allows physicists to apply fundamental laws—like conservation principles and —to predict and explain the system's behavior, often simplifying complex interactions by idealizing boundaries or neglecting minor influences. Physical systems are classified based on their interactions with the , particularly regarding exchanges of and . An open system can transfer both and across its boundary, as seen in processes like in fluids or chemical reactions in living organisms. A closed system exchanges (such as or work) but not , exemplified by a sealed piston-cylinder containing gas undergoing compression. An isolated system exchanges neither, representing an idealization where total and remain constant, like the entire in cosmological models. These categories facilitate analysis in , , and other branches of physics, enabling the application of laws such as the first law of , which states that the change in equals added minus work done by the system. The of a physical system is fully specified when all its relevant physical properties—such as , , , , or quantum —have definite values, allowing deterministic predictions under classical or quantum frameworks. Properties evolve over time through internal dynamics or external influences, governed by equations like Newton's laws for classical systems or the for quantum ones. In practice, real-world systems are modeled approximately, accounting for uncertainties via for large ensembles or for small deviations. This framework underpins diverse applications, from engineering devices to understanding natural phenomena like planetary motion or subatomic interactions.

Definition and Fundamentals

Core Definition

A physical system is any portion of the physical selected for analysis, consisting of , , or both, with specified that separate it from its surroundings or . This demarcation allows physicists to focus on interactions within the defined region while treating external influences as inputs or outputs across the boundary. The encompasses diverse scales, from subatomic particles to astronomical structures, enabling the application of physical laws to predict behavior. While the general concept of a physical system traces back to in the with , the notion of a developed in the within , notably through the work of , who in 1850 established the foundational principles of modern by unifying and work under . Clausius's work, building on earlier ideas from Sadi Carnot and Émile Clapeyron, introduced the notion of a as a bounded entity undergoing processes like and work, with his 1854 analysis of the providing a key mathematical framework. This development evolved from Newtonian mechanics, where 's 17th-century laws described the motion of point masses and forces, laying the groundwork for analyzing isolated mechanical systems before extending to energetic and thermal interactions in the . In physics, defining a physical system serves to simplify the complexity of the universe by isolating variables and interactions, facilitating predictions and deeper understanding of natural phenomena through controlled analysis. For instance, a gas confined in a piston-cylinder represents a simple physical system where , , and can be studied under thermodynamic laws, contrasting with the entire Earth's atmosphere as a highly involving myriad coupled processes like and .

System Boundaries and Components

In physics, the boundaries of a physical system are defined as real or imaginary surfaces that separate the system from its surroundings, delineating the portion of the under study. These boundaries can be fixed or movable and are often conceptualized as infinitely thin interfaces across which properties such as , , or may change abruptly. The permeability of these boundaries with respect to and depends on the nature of exchanges allowed, enabling the isolation of specific interactions for . The internal components of a physical system consist of in forms such as particles, molecules, or continuous fields (e.g., electromagnetic fields), along with various energy manifestations including of motion, due to position or configuration, and internal . These components interact through fundamental forces, such as gravitational, electromagnetic, or forces, mediated by fields that govern the system's . For instance, in a mechanical system, components might include masses connected by springs, where interactions arise from elastic forces. Criteria for selecting system boundaries and components are guided by the goal of making the analysis mathematically tractable, often prioritizing simplicity and relevance to the physical phenomena of interest. Boundaries are chosen at convenient locations to enclose relevant interactions while excluding extraneous influences, such as drawing them around a single object or a group of interacting elements to apply specific frameworks. In conservative systems, where forces derive from a potential, boundaries are selected to encompass all such interactions, facilitating the use of Lagrangian mechanics, which reformulates dynamics in terms of generalized coordinates and minimizes computational complexity for multi-body problems. Defining boundaries can present challenges, particularly when they are arbitrary or ill-defined, leading to necessary approximations in modeling. In , for example, interfaces between fluids or between a fluid and a solid may be fuzzy due to mixing or thin transition layers, requiring techniques like approximations to simplify the governing equations while capturing essential flow behaviors near surfaces. Such approximations introduce errors but enable solvable models for complex, real-world scenarios where exact boundaries are impractical to specify.

Classification of Physical Systems

By Interaction with Surroundings

Physical systems are classified based on their interactions with the surrounding , particularly regarding the exchange of and . This categorization influences how physical laws, such as those in , apply to the system. The three primary types are isolated, closed, and open systems. An exchanges neither nor with its surroundings, making it an idealization rarely achieved in practice. The entire is often considered an example of an , as there is no known external with which it can interact. In such systems, all processes occur internally without external influence. A permits the exchange of , such as or work, but not with its surroundings. A sealed thermos flask serves as an approximate example, where can slowly transfer across the while the contents remain contained. This type of system maintains fixed composition but can undergo changes in due to external energy flows. An open system allows both and to exchange freely with the surroundings. A pot of with the lid off exemplifies this, as (matter) escapes while enters from the stove. Open systems are common in natural and engineering contexts, where continuous inputs and outputs drive dynamic behavior. These classifications have significant implications for the applicability of thermodynamic laws. The of , which states that is conserved and can neither be created nor destroyed, holds universally for all system types, as it reflects the invariance of total in any process. In contrast, the second , concerning , asserts that the of an never decreases and typically increases over time for irreversible processes, providing a directionality to spontaneous changes within the system. For closed and open systems, changes must account for external exchanges, often requiring consideration of the combined system and surroundings to apply the second fully.

By Scale and Complexity

Physical systems are classified by spatial scale, which determines the dominant physical principles governing their behavior. At the microscopic scale, systems involve atomic or subatomic particles, such as electrons orbiting atomic nuclei, where provides the fundamental description due to wave-particle duality and probabilistic outcomes. These systems exhibit phenomena like superposition and tunneling, which are negligible at larger scales, and their dynamics are captured by the rather than classical trajectories. The mesoscopic scale occupies an intermediate regime, typically at the nanoscale (1–100 ), where systems like quantum dots or nanowires display behaviors that bridge quantum and classical regimes. In these nanoscale devices, quantum effects such as and coexist with classical dissipation, enabling applications in and sensors. This scale is characterized by finite-size effects and that blur strict quantum-classical boundaries, often modeled using mesoscopic transport theories. At the macroscopic scale, systems encompass everyday objects and larger structures, such as planetary orbits or fluid flows, where suffices due to the averaging out of quantum fluctuations over vast numbers of particles. For instance, the motion of planets follows Newtonian , treating bodies as point masses without quantum corrections. These systems are analyzed using approximations, insensitive to atomic details, as links microscopic interactions to bulk properties like and . Beyond scale, physical systems are categorized by internal complexity, reflecting the number and interaction strength of components. Simple systems feature few and predictable dynamics, exemplified by a single , whose motion is governed by linear or weakly nonlinear equations yielding periodic oscillations. In contrast, complex systems involve numerous interacting elements, leading to emergent behaviors like sensitivity to initial conditions, as seen in weather patterns modeled by . These systems, such as , exhibit deterministic yet unpredictable evolution due to nonlinearities, where small perturbations amplify into large divergences, a hallmark of chaotic dynamics.

Key Properties

Conservation Laws

Conservation laws represent foundational principles in physics, asserting that specific quantities—such as , , and —remain unchanged within physical systems under defined conditions, reflecting underlying symmetries in nature. These laws enable the prediction and analysis of system evolution without tracking every interaction, applying most rigorously to isolated systems that exchange neither nor with their surroundings. Derived from empirical observations and theoretical frameworks, they form the for understanding diverse phenomena from to . The , encapsulated in of , posits that the total of a is invariant; it can neither be created nor destroyed, only converted between forms. Mathematically, for a , this is expressed as the change in \Delta U equaling the Q added to the minus the work W done by the : \Delta U = Q - W where U denotes . This principle was independently formulated by in 1847 and in 1850, building on earlier work by demonstrating the mechanical equivalent of . It governs processes in closed s, ensuring energy balance in everything from chemical reactions to planetary motion. Conservation of momentum maintains that the total momentum of an isolated system remains constant if no external forces act upon it. Linear momentum for a particle is defined as \mathbf{p} = m \mathbf{v}, where m is and \mathbf{v} is , while angular momentum is \mathbf{L} = I \omega, with I as the and \omega as . These arise from Newton's laws of motion, particularly the third law stating that action and reaction forces are equal and opposite, as detailed in his 1687 Philosophiæ Naturalis Principia Mathematica. In collisions or interactions within isolated systems, momentum redistribution occurs without net change, exemplified by the recoil of a gun when firing a bullet. The asserts that the total mass in a —impermeable to exchange—remains constant throughout any . First established by through precise experiments in the late , this law revolutionized chemistry by quantifying reactions and refuting earlier notions of creation or annihilation. In the framework of , extended it in 1905 to the mass-energy equivalence principle, E = mc^2, where energy E interchanges with mass m at the c, allowing transformations like while conserving total mass-energy. These conservation laws apply strictly to isolated or closed physical systems, where external influences are absent or negligible; in open systems, apparent non-conservation arises from unaccounted exchanges with the . Violations typically signal overlooked interactions, errors, or the of new physical regimes, such as quantum or relativistic effects. As noted in classifications, isolated systems provide the context for these principles' exact adherence, guiding analyses across scales from subatomic particles to cosmological structures.

Equilibrium and Stability

In physical systems, equilibrium refers to a state where the system experiences no net change over time, maintaining balance among its internal processes and interactions. Thermal equilibrium occurs when two or more systems have reached the same , resulting in no net flow between them. This condition is formalized by the , which states that if two systems are each in thermal equilibrium with a third system, they are in thermal equilibrium with each other, enabling the consistent measurement of across systems. Mechanical equilibrium describes a state in which a system is at rest or moving with constant velocity, with no net acceleration. For this to hold, the vector sum of all forces acting on the system must be zero (\sum \mathbf{F} = 0), and the sum of all torques about any axis must also be zero (\sum \boldsymbol{\tau} = 0). These conditions ensure that the system's linear and angular momentum remain constant, preventing translational or rotational motion from changing. Stability assesses how a responds to small from its state. In stable , the returns to its original position after a disturbance, as seen in a ball at the bottom of a valley where is minimized and restoring act to restore balance. Unstable occurs when a causes the to diverge further, such as a ball balanced on a hilltop where any deviation increases and amplifies . Neutral features no net restoring or diverging , allowing the to remain in a new position after , exemplified by a ball on a flat surface where is constant./09%3A_Statics_and_Torque/9.03%3A_Stability) In the context of dynamical systems, equilibrium points and their are analyzed in , which represents all possible states of the system as points in a multidimensional space of variables like and . Attractors in are subsets toward which trajectories evolve over time, such as fixed points for stable or limit cycles for oscillatory behaviors. specifically characterizes an as stable if nearby trajectories remain close for all future times following small perturbations, and asymptotically stable if they converge to the ; this is determined by the existence of a that decreases along trajectories, quantifying the system's resilience to infinitesimal disturbances.

Modeling and Analysis

Mathematical Frameworks

Physical systems are described and analyzed using various mathematical frameworks that capture their dynamics at different scales and levels of complexity. Newtonian mechanics serves as the cornerstone for classical macroscopic systems, where the behavior of particles and rigid bodies is governed by deterministic laws. Central to this approach is Newton's second law, expressed as \mathbf{F} = m \mathbf{a}, where \mathbf{F} is the acting on a of m and \mathbf{a}. This yields a system of second-order ordinary differential equations for the trajectories, solvable under specified initial conditions and forces derived from potentials or interactions. For more complex systems involving constraints or , offers a reformulation that simplifies the analysis by focusing on energy rather than forces directly. Introduced by , this framework defines the L = T - V, where T is the and V is the . The follow from the Euler-Lagrange equation: \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}_i} \right) - \frac{\partial L}{\partial q_i} = 0, for each generalized coordinate q_i and its time derivative \dot{q}_i. This , derived from the principle of least action, is particularly advantageous for systems with symmetries or , enabling the use of coordinates like angles or arc lengths. Hamiltonian mechanics extends the Lagrangian formulation into phase space, providing a symmetric treatment of position and momentum variables that proves essential for statistical mechanics and quantum transitions. developed this approach, defining the Hamiltonian H = T + V as the total energy in terms of generalized coordinates q_i and conjugate momenta p_i. The dynamics are then given by Hamilton's canonical equations: \dot{q}_i = \frac{\partial H}{\partial p_i} and \dot{p}_i = -\frac{\partial H}{\partial q_i}. This first-order system preserves the symplectic structure of phase space, facilitating the study of conserved quantities via Poisson brackets and long-term stability. At the , quantum frameworks replace with probabilistic wave mechanics to model systems where particles exhibit wave-like properties. The , formulated by , governs the time evolution of the wave function \psi: i \hbar \frac{\partial \psi}{\partial t} = \hat{H} \psi, where \hat{H} is the operator incorporating kinetic and potential energies, \hbar is the reduced Planck's constant, and i is the . Solutions to this yield probabilities via |\psi|^2, enabling predictions for atomic and subatomic phenomena that cannot address.

Computational Approaches

Computational approaches play a crucial role in analyzing physical systems where analytical solutions are infeasible, enabling the simulation of through numerical techniques that approximate continuous equations on discrete computational grids or ensembles. These methods discretize the governing equations derived from Newtonian or , allowing researchers to predict system behavior under various conditions, such as in fluid flows, material deformations, or atomic interactions. By leveraging , these techniques have transformed the study of physical systems from idealized models to realistic, large-scale simulations. Numerical integration methods, particularly the Runge-Kutta family, are widely used to solve ordinary differential equations (ODEs) arising from Newtonian in simulations of mechanical systems. Developed in the early , the classical fourth-order Runge-Kutta method (RK4) provides a balance of accuracy and computational efficiency by evaluating the derivative at multiple intermediate points within each time step, achieving an error of order O(h^5) where h is the step size. In physics simulations, such as or particle trajectories, RK4 is applied to integrate second-order ODEs like \mathbf{\ddot{r}} = \mathbf{F}/m, where \mathbf{r} is position and \mathbf{F} is force, enabling stable predictions over long timescales without excessive computational cost. For instance, in N-body gravitational simulations, RK4 has been employed to model planetary orbits with high fidelity, outperforming lower-order methods like Euler integration in preserving . Monte Carlo methods employ statistical sampling to estimate properties of probabilistic physical systems, particularly in where exact solutions are intractable. Originating from work on diffusion but adapted for molecular systems, these methods generate ensembles of random configurations according to a , such as the [Boltzmann distribution](/page/Boltzmann distribution), to compute averages like or . The seminal Metropolis algorithm, introduced in , uses a to sample states by proposing moves and accepting or rejecting them based on an difference \Delta E, with acceptance probability \min(1, e^{-\Delta E / kT}), where k is Boltzmann's constant and T is ; this ensures and to distributions. In applications to particle interactions in gases, Monte Carlo simulations have quantified phase transitions in hard-sphere systems, revealing phenomena like the fluid-solid transition that align with experimental data. Finite element analysis (FEA) addresses continuum problems in physical systems by discretizing the domain into a of finite elements, solving partial differential equations (PDEs) numerically for fields like or distribution. Pioneered in the 1950s for , the method approximates solutions within each element using basis functions, such as linear polynomials for , and assembles a global to enforce via \mathbf{K} \mathbf{u} = \mathbf{f}, where \mathbf{K} is the , \mathbf{u} the nodal displacements, and \mathbf{f} the forces. This approach excels in simulating in solids under complex loads, as seen in early applications to analysis, where refinement improves accuracy for irregular geometries without analytical tractability. Modern FEA implementations handle nonlinear materials and large deformations, providing quantitative insights into failure modes validated against experimental benchmarks. Molecular dynamics (MD) simulates the time evolution of nanoscale physical systems by integrating Newton's laws for interacting atoms, using empirical potentials to model interatomic forces. Introduced in the late 1950s for hard-sphere gases, MD tracks trajectories via velocity , a method that updates positions and velocities in a single step with second-order accuracy, preserving energy better than explicit Euler schemes. Potentials like the Lennard-Jones form V(r) = 4\epsilon [(\sigma/r)^{12} - (\sigma/r)^6] approximate van der Waals interactions, enabling studies of motions in liquids or solids. For example, early MD simulations of revealed diffusion coefficients matching experimental values within about 15%, establishing the method's validity for probing thermodynamic at the scale.

References

  1. [1]
    [PDF] 2 The Structure of Physics
    A physical system is said to be “in a state” when all its physical properties have certain values. In practice, any physical system that can be investigated is.
  2. [2]
    3.1 Thermodynamic Systems - University Physics Volume 2
    Oct 6, 2016 · A thermodynamic system includes anything whose thermodynamic properties are of interest. It is embedded in its surroundings or environment; ...
  3. [3]
    None
    ### Summary of Definition of Physical System and Historical Notes
  4. [4]
    Rudolf Clausius - The Information Philosopher
    In 1865, Clausius coined a word and a new symbol for the part of the heat that cannot be converted into mechanical work. ...the author proposed to call this ...
  5. [5]
    A History of Thermodynamics: The Missing Manual - PMC
    As noted by Gibbs, in 1850, Clausius established the first modern form of thermodynamics, followed by Thomson's 1851 rephrasing of what he called the Second Law ...
  6. [6]
    [PDF] Newtonian Mechanics and its Philosophical Significance Ryan ...
    physical systems ... That is, they provide frameworks of constraints in which physical quantities can be constructed and whose evolution can be determined.
  7. [7]
  8. [8]
    Physical and Conceptual Systems - incose
    A physical system is an arrangement of parts or elements that together exhibit behaviour that the individual constituents do not. (This definition includes ...
  9. [9]
    Physical System - an overview | ScienceDirect Topics
    A physical system is a set of elements that operate together to accomplish an objective. A system is not a randomly assembled set of elements.
  10. [10]
    Systems, boundaries, and surroundings - Thermodynamics I - Fiveable
    By defining what's inside and outside our system, we can track energy flow and figure out how things change.
  11. [11]
    [PDF] Boundary layers
    If the mainstream flow accelerates, i.e. grows with x, the boundary layer tends to remain thin. This happens at the front of a moving body, where the fluid must.
  12. [12]
    [PDF] Fluid Dynamics II (D) - DAMTP
    Sep 18, 2020 · This gives an estimate for the pressure variations along the flow. ∆ x p = O(ρU2). Second boundary layer approximation Consider the y-momentum ...
  13. [13]
    Basic Definitions - Chemistry LibreTexts
    Jun 15, 2023 · An isolated system does not allow either matter or energy to enter or leave. A thermos or cooler is approximately an isolated system. There are ...
  14. [14]
    Isolated Systems in Physics | Overview, Types & Examples - Lesson
    Isolated System Examples · The inside of a vacuum thermos is an isolated system. · Our universe is an isolated system because, as far as we know, there is nothing ...
  15. [15]
    The laws of thermodynamics (article) | Khan Academy
    An isolated system is one that cannot exchange either matter or energy with its surroundings. A perfect isolated system is hard to come by, but an insulated ...
  16. [16]
    2nd Law of Thermodynamics - Chemistry LibreTexts
    Jan 1, 2025 · The Second Law of Thermodynamics states that the state of entropy of the entire universe, as an isolated system, will always increase over time.Major players in developing... · Derivation and Explanation · Gibbs Free Energy
  17. [17]
    5.5: Quantum Theory and Atomic Orbitals - Chemistry LibreTexts
    Jul 18, 2022 · Macroscopic objects act as particles. Microscopic objects (such as electrons) have properties of both a particle and a wave. Their exact ...
  18. [18]
    DOE Explains...Quantum Mechanics - Department of Energy
    Quantum mechanics also helped explain the structure of the atom. It helped make sense of the photoelectric effect, which involves how materials emit electrons ...
  19. [19]
    Mesoscopic physics of nanomechanical systems | Rev. Mod. Phys.
    Dec 8, 2022 · This review describes, at a conceptual level, basic theoretical ideas and explicative experiments pertaining to mesoscopic physics of nanomechanical systems.
  20. [20]
    [PDF] Mesoscopic Methods in Engineering and Science
    Matter, conceptually classified into fluids and solids, can be completely described by the microscopic physics of its con- stituent atoms or molecules.
  21. [21]
    [PDF] From Microscopic to Macroscopic Behavior - Princeton University
    Our goal is to understand the properties of macroscopic systems, that is, systems of many electrons, atoms, molecules, photons, or other constituents.
  22. [22]
    8 MATHEMATICAL AND NUMERICAL METHODS
    The macroscopic level deals with bulk averages across microstructure and is the domain of continuum mechanics and thermodynamics. The mesoscopic level is ...Missing: classification | Show results with:classification<|control11|><|separator|>
  23. [23]
    Chaos Theory - an overview | ScienceDirect Topics
    A simple pendulum swing would have a two-dimensional phase space of velocity and angle. Once the movement is represented or mapped on the coordinates, a pattern ...
  24. [24]
    Chaos Theory - an overview | ScienceDirect Topics
    Chaos theory is defined as the study of how simple physical systems governed by deterministic laws can exhibit unpredictable behavior, leading to regimes ...
  25. [25]
    Chaos - Stanford Encyclopedia of Philosophy
    Jul 16, 2008 · The big news about chaos is supposed to be that the smallest of changes in a system can result in very large differences in that system's behavior.
  26. [26]
    Conservation Laws - HyperPhysics
    In mechanics, examples of conserved quantities are energy, momentum, and angular momentum. The conservation laws are exact for an isolated system.
  27. [27]
    Thermodynamic Asymmetry in Time
    Nov 15, 2001 · The First Law expresses the conservation of energy and is founded upon the impossibility of creating a machine that can create energy. The law ...
  28. [28]
    [PDF] LECTURE NOTES ON THERMODYNAMICS
    May 17, 2025 · Where possible, citations of original work are provided; most citations are linked directly to the original source material via click-able text.
  29. [29]
    Newton's Philosophiae Naturalis Principia Mathematica
    Dec 20, 2007 · Newton originally planned a two-book work, with the first book consisting of propositions mathematically derived from the laws of motion, ...
  30. [30]
    Newton's Laws of Motion - Glenn Research Center - NASA
    Jun 27, 2024 · In 1686, he presented his three laws of motion in the “Principia Mathematica Philosophiae Naturalis.” By developing his three laws of motion, ...
  31. [31]
    Law of Conservation of Matter (Antoine Lavoisier)
    Lavoisier found that mass is conserved in a chemical reaction. The total mass of the products of a chemical reaction is always the same as the total mass of the ...
  32. [32]
    E= mc^2 - University of Pittsburgh
    It was used by Einstein in his 1905 E=mc2 paper "Is the energy content of a body dependent on its inertial?". Divide the kinetic energy equally over two ...
  33. [33]
    Conservation of energy
    Energy for an isolated system is always conserved. It may change forms, but the total amount of energy in an isolated system is constant.
  34. [34]
    Zeroth Law - Thermal Equilibrium | Glenn Research Center - NASA
    May 2, 2024 · The zeroth law of thermodynamics is an observation. When two objects are separately in thermodynamic equilibrium with a third object, they are in equilibrium ...
  35. [35]
    Thermal equilibrium - HyperPhysics
    The "zeroth law" states that if two systems are at the same time in thermal equilibrium with a third system, they are in thermal equilibrium with each other.
  36. [36]
    8.1 The First Condition for Equilibrium – Douglas College Physics ...
    The first condition necessary to achieve equilibrium is that the net external force on the system must be zero, so that net F = 0. Glossary. static equilibrium ...
  37. [37]
    static equilibrium - Oregon State University
    A rigid object is in static equilibrium if the net torque is zero at every point on the object, and the net force on the object is zero.
  38. [38]
    [PDF] 12.006J F2022 Lecture 19: Introduction to Strange Attractors
    Oct 31, 2022 · In phase space, trajectories on an aperiodic attractor can diverge, e.g.,. We shall see that the divergence of trajectories is exponential in ...
  39. [39]
    [PDF] 4 Lyapunov Stability Theory
    In this section we review the tools of Lyapunov stability theory. These tools will be used in the next section to analyze the stability properties.
  40. [40]
    [PDF] Eighty Years of the Finite Element Method - arXiv
    As Ray Clough commented in his 1980 paper, One aspect of the FEM, mathematical modeling of continua by discrete elements, can be related to work done ...
  41. [41]
    [PDF] A history of Runge-Kutta methods f ~(z) dz = (x. - x.-l) - People
    The paper by W. Kutta~ which appeared in 1901, took the analysis of Runge--Kutta methods as far as order 5. He made a complete classification of order 4 ...Missing: 1899 | Show results with:1899
  42. [42]
    Numerical Integration of Newton's Equation of Motion - ComPADRE
    In general, a Runge-Kutta method adjusts the partition of dt as well as the constants ai and ci so as to minimize the error. In the fourth-order Runge-Kutta ...
  43. [43]
    Equation of State Calculations by Fast Computing Machines
    A general method, suitable for fast computing machines, for investigating such properties as equations of state for substances consisting of interacting ...