Fact-checked by Grok 2 weeks ago

Thermodynamic system

A thermodynamic system is a defined portion of the , consisting of or a in space, that is selected for thermodynamic study and separated from its surroundings by a conceptual or physical . This delineates the system from everything else, which constitutes the surroundings, allowing analysis of energy, work, , and exchanges across it. The is typically a large collection of atoms or molecules, enabling the definition of macroscopic average properties such as , , and . Thermodynamic systems are classified into three main types based on their interactions with the surroundings: open, , and . An , also known as a , permits the exchange of both and (such as and work) with its surroundings, as seen in devices like engines or turbines where fluids flow in and out. A , or control mass, allows energy transfer but no net exchange, conserving within the boundary while or work can cross it, exemplified by a sealed piston-cylinder . An exchanges neither nor with the surroundings, representing an idealized scenario where the total energy and remain constant, such as the entire in theoretical considerations. The behavior of a thermodynamic system is characterized by its thermodynamic properties and , which are macroscopic observables that describe its condition under . occurs when the system exhibits no net changes over time, with uniform mechanical, thermal, and chemical potentials, allowing the to be fully specified by a minimal set of independent properties via an , such as the pV = nRT. Processes involving the system—changes from one to another—can be reversible (quasi-static and without dissipative effects like ) or irreversible, influencing quantities like and in thermodynamic cycles. These concepts form the foundation for applying the to predict system behavior in and physical applications.

Fundamentals

Definition and Scope

A thermodynamic system is defined as a specific, identifiable portion of or a designated in space that is selected for the purpose of thermodynamic analysis, set apart from the rest of the known as the surroundings. This delineation allows researchers to focus on the , work, and transformations within the system while treating the surroundings as the external context influencing or being influenced by it. In thermodynamic studies, the entire is partitioned into the system and its surroundings, providing a for applying principles to isolated or interacting components. The scope of thermodynamic systems encompasses both macroscopic collections, such as a volume of gas in a piston-cylinder device, and microscopic ensembles, like assemblies of molecules where average properties can be defined despite individual fluctuations. Systems are typically chosen for their controllability, enabling the precise application of thermodynamic laws to predict behavior under controlled conditions of heat, work, or matter exchange. A classic example is an ideal gas confined in a rigid container, where the system's properties remain uniform and analyzable without significant intermolecular interactions. Central to this concept are state variables, such as (P), (V), and (T), which fully characterize the state of the system and determine its thermodynamic properties. The first law of provides a foundational tool for these systems, ensuring that changes—through addition or work done—are conserved within the defined boundaries. Such systems form the basis for further classifications, including isolated, closed, and open types, depending on their interactions with the surroundings.

Historical Development

The concept of a thermodynamic system emerged in the early amid efforts to understand heat engines and energy conversion. In 1824, Sadi Carnot published Reflections on the Motive Power of Fire and on Machines Fitted to Develop that Power, which analyzed the efficiency of idealized engines operating between hot and cold reservoirs, laying the groundwork for treating engines as bounded systems interacting with thermal surroundings. This mechanical perspective marked the initial formalization of systems as entities capable of reversible work cycles, influencing subsequent thermodynamic theory. Building on Carnot's ideas, advanced the framework in the by introducing the notion of a thermodynamic system in the context of and . In his 1850 paper "On the Moving Force of ," Clausius reformulated Carnot's using the emerging , treating the system as a closed where and work exchanges could be quantified without . By the mid-1850s, Clausius's work on —formalized in 1865—explicitly described systems as aggregates of subject to irreversible processes, where measures the unavailable energy within the system. Concurrently, William Thomson (later ) proposed an absolute temperature scale in 1848, based on Carnot's efficiency principle, which provided a universal metric for system states independent of material properties. Key milestones further refined the system's scope for complex materials. In 1876, J. Willard Gibbs published "On the Equilibrium of Heterogeneous Substances," deriving the F = C - P + 2 (where F is , C components, and P phases), which formalized the constraints on multi-component systems at equilibrium. This work extended the concept beyond simple engines to heterogeneous mixtures, emphasizing compositional variables in system behavior. The evolution from macroscopic mechanical descriptions to microscopic foundations occurred through Ludwig Boltzmann's contributions in the 1870s, particularly his 1872 H-theorem and subsequent papers, which linked thermodynamic properties to statistical ensembles of particles, deriving as S = k \ln W (with k Boltzmann's constant and W microstates). In the , the concept expanded to non-equilibrium conditions through Ilya Prigogine's irreversible thermodynamics. From the onward, Prigogine, building on Théodore de Donder's affinity concept, developed theories for open systems far from equilibrium, introducing dissipative structures in works like Introduction to Thermodynamics of Irreversible Processes (1955) and From Being to Becoming (1980). His 1945–1949 formulations showed how non-equilibrium systems could self-organize via , as recognized in his 1977 for advancing . These refinements broadened the thermodynamic system to encompass dynamic, far-from-equilibrium phenomena in and .

Boundaries and Interactions

Walls and Boundaries

In thermodynamics, walls refer to the real or imaginary surfaces that enclose a thermodynamic system, delineating its spatial boundaries and separating it from the external . These walls can be physical, such as the material casing of a , or conceptual, like an abstract surface drawn around a to define the system's extent. The choice of wall type fundamentally influences the possible interactions between the system and its surroundings by regulating the flow of , work, and across the . Walls are classified based on their permeability to heat, matter, and mechanical deformation. Diathermic walls allow the conduction of heat while typically remaining impermeable to matter, as exemplified by thin metal sheets that facilitate thermal equilibrium without particle exchange. In contrast, adiabatic walls provide thermal insulation, preventing heat transfer; this is often achieved through materials like vacuum layers or highly reflective barriers that minimize conduction, convection, and radiation. Impermeable walls block the passage of matter entirely, ensuring no mass crosses the boundary, which is common in setups designed to maintain constant composition. Additionally, walls can be semi-permeable, selectively permitting the transfer of specific components or species while restricting others, though such configurations are more specialized and often involve membranes. Regarding mechanical properties, walls may be rigid, fixing the system's volume and prohibiting expansion or contraction, or movable, such as a frictionless piston, which enables volume changes and associated mechanical work. The role of these walls is to control the permissible exchanges, thereby shaping the system's behavior during processes. For instance, an insulated cylinder with rigid adiabatic walls exemplifies a setup for adiabatic processes, where no heat enters or leaves, preserving the system's solely through work if the allows displacement. In such cases, a movable introduces the possibility of expansion work, quantified conceptually as times volume change (PdV), highlighting how flexibility couples interactions to thermodynamic changes without altering the no-heat-transfer condition. The external to these walls constitutes the surroundings, providing the context for any permitted interactions.

Surroundings

In thermodynamics, the surroundings refer to everything external to the defined thermodynamic system, encompassing the rest of the that interacts with it. This division allows for the analysis of energy and matter exchanges across the system , where the surroundings serve as the counterpart to the system's changes./University_Physics_II_-Thermodynamics_Electricity_and_Magnetism(OpenStax)/03%3A_The_First_Law_of_Thermodynamics/3.02%3A_Thermodynamic_Systems) The surroundings may be finite, such as another subsystem in a larger setup, or idealized as infinite, like vast reservoirs that maintain uniform conditions despite interactions. Interactions between the system and surroundings involve transfers of , work, and , which drive changes in the system's state. flows as across the due to differences, while work arises from mechanical forces, such as or , acting through the interface. occurs in permeable boundaries, allowing flow that alters both composition and properties. These exchanges can be reversible, occurring infinitesimally slowly with no net increase in the , or irreversible, involving finite gradients that generate . Surroundings are often modeled as thermal reservoirs with large heat capacity to sustain constant temperature during heat transfer, or as mechanical reservoirs providing constant pressure, such as the atmosphere acting on a piston. For instance, in an open flask experiment, the laboratory atmosphere serves as the surroundings, supplying or absorbing heat and matter while approximated as unchanging due to its scale. This idealization simplifies analysis by assuming the surroundings' properties remain fixed, focusing attention on the system's response. The boundary, such as a wall, defines the interface for these interactions but does not alter the surroundings' external role.

Classification by Matter and Energy Exchange

Isolated Systems

An isolated thermodynamic system is defined as one that exchanges neither nor , including or work, with its surroundings, ensuring that both total energy and are conserved within the system itself. This isolation is achieved through boundaries that are impermeable to mass, rigid to prevent work transfer, and adiabatic to block flow. In such systems, all internal processes redistribute existing energy and without external influence, maintaining a constant total as per of , where the change in ΔU equals zero due to the absence of or work inputs. A key property of isolated systems is their evolution toward driven by the second law of thermodynamics, which states that the of an isolated system never decreases and tends to increase until it reaches a maximum. This increase reflects the natural progression toward greater disorder or randomness within the system, such as the of gases or the equalization of temperatures among components, without any external intervention. Consequently, isolated systems provide an ideal framework for studying spontaneous processes, as their behavior is solely determined by internal and the irreversible tendency toward . Representative examples of isolated systems include the entire , considered the ultimate isolated system where no or can escape its boundaries, and an idealized thermos bottle containing a sealed, insulated that approximates over short timescales. Another example is a rigid, insulated filled with gas, where no occurs due to the sealed boundary, and no or work exchange happens because of the insulation and immovability. The implications of isolated systems are profound for thermodynamic analysis, as they allow researchers to isolate the effects of the first and second laws in pure form, revealing how remains fixed while maximization governs the direction of spontaneous changes. This makes isolated systems invaluable for theoretical studies of and irreversibility, though perfect isolation is rarely achieved in practice and is often approximated for experimental purposes.

Closed Systems

A closed thermodynamic system is defined as a region of space bounded by an impermeable barrier that prevents the exchange of matter with its surroundings, while allowing the transfer of energy in the form of and work. This boundary ensures that the within the system remains constant throughout any , making closed systems ideal for analyzing changes in internal properties without complications from mass flow. Unlike isolated systems, which prohibit all exchanges, closed systems permit interactions that can alter their content without altering their composition. The fundamental relation governing energy changes in a closed system is the first law of thermodynamics, expressed as \Delta U = Q - W, where \Delta U is the change in internal energy of the system, Q is the net heat transferred to the system, and W is the net work done by the system. This equation arises from the principle of conservation of energy, which states that energy cannot be created or destroyed, only transformed or transferred; thus, any increase in the system's internal energy must equal the energy added via heat minus the energy expended via work. To derive it, consider a differential process: the infinitesimal change in internal energy dU equals the heat added đQ minus the work done đW, leading to the integrated form \Delta U = Q - W for finite changes between states. The sign convention is crucial: Q > 0 indicates heat absorbed by the system (increasing U), while W > 0 indicates work performed by the system (decreasing U); conversely, heat rejected or work received by the system carry negative signs. Internal energy U is a state function depending solely on the system's state (e.g., for an ideal gas, U = U(T)), ensuring \Delta U is path-independent. Common processes in closed systems include isothermal, adiabatic, and isobaric types, each characterized by specific constraints on thermodynamic variables and implications for state functions like internal energy U and enthalpy H = U + PV. In an isothermal process, temperature remains constant (T = constant), so for an ideal gas \Delta U = 0 and thus Q = W; heat input exactly balances expansion work. An adiabatic process involves no heat transfer (Q = 0), so \Delta U = -W, meaning the system's internal energy change directly reflects work done, often leading to temperature variations in gases. For an isobaric process at constant pressure (P = constant), work is W = P \Delta V, and the heat transfer equals the enthalpy change \Delta H = Q, highlighting H's utility for constant-pressure analyses. These processes demonstrate how closed systems evolve while conserving mass, with changes in U and H determined by initial and final states. Representative examples of closed systems include a gas confined in a piston-cylinder with impermeable seals, where the piston allows work via volume change but no escapes, enabling study of or processes. Another is a conducted in a sealed rigid container, such as a bomb calorimeter, where reactants evolve or perform minimal work against fixed boundaries, allowing precise measurement of changes without loss. These setups underscore the practical role of closed systems in and scientific applications.

Open Systems

An open thermodynamic system, also referred to as a , is defined as a where both and can cross the system boundaries, enabling mass flow into and out of the . This permeability distinguishes open systems from those with impermeable boundaries, allowing for processes involving inflow and outflow of substances such as fluids or gases. In applications, open systems are modeled using to analyze flows where the itself may not contain all the material undergoing transformation. A fundamental concept for open systems is the continuity equation, which ensures mass conservation by balancing the accumulation of mass within the control volume against the net mass flow across its boundaries. The general form of the continuity equation is \frac{dm_{cv}}{dt} = \sum \dot{m}_i - \sum \dot{m}_e, where m_{cv} is the mass inside the control volume, \dot{m}_i represents inlet mass flow rates, and \dot{m}_e denotes outlet mass flow rates. For steady-state conditions, where the mass within the system remains constant over time, this simplifies to \sum \dot{m}_i = \sum \dot{m}_e, implying no net accumulation or depletion of mass. In open systems, energy analysis relies on , defined as H = U + PV, where U is , P is , and V is . accounts for the total transported by flowing matter, incorporating not only but also the flow work PV required to push the mass across the system boundaries. This property is particularly useful in flow processes, as it simplifies the bookkeeping of energy changes associated with . The first law of thermodynamics for a steady-state open system, neglecting kinetic and potential energy changes, is expressed as Q - W = \Delta H, where Q is the heat transfer to the system, W is the work done by the system, and \Delta H is the net change in enthalpy between outlet and inlet streams (\Delta H = \sum \dot{m}_e h_e - \sum \dot{m}_i h_i, with h as specific ). This equation is derived from the general energy balance for an open system: the rate of energy accumulation equals the net energy inflow by , work, and mass flow. For steady state, accumulation is zero, yielding \dot{Q} - \dot{W} + \sum \dot{m}_i (h_i + \frac{V_i^2}{2} + g z_i) = \sum \dot{m}_e (h_e + \frac{V_e^2}{2} + g z_e). Omitting kinetic (V^2/2) and potential (g z) terms for many applications simplifies to \dot{Q} - \dot{W} = \sum \dot{m}_e h_e - \sum \dot{m}_i h_i, or in integrated form Q - W = \Delta H. Examples of open systems include a pot of boiling water, where liquid water evaporates and mass leaves as vapor while heat enters from the stove; a , which ingests air and fuel, combusts them, and expels hot exhaust gases to produce thrust; and chemical reactors in , such as continuous stirred-tank reactors, where reactants flow in and products flow out to maintain steady production. In cases of very low mass flow rates, open systems can be approximated as closed systems for simplified analysis.

Equilibrium and Dynamics

Equilibrium States

In thermodynamics, an equilibrium state of a system is characterized by the absence of net changes in its macroscopic properties, such as temperature, pressure, and composition, over time. This condition requires simultaneous thermal, mechanical, and chemical equilibrium. Thermal equilibrium occurs when there is no net heat flow within the system or between the system and its surroundings, leading to a uniform temperature distribution. Mechanical equilibrium implies no unbalanced forces or net macroscopic motion, resulting in uniform pressure throughout the system. Chemical equilibrium is reached when there are no net chemical reactions or phase transitions, with the rates of forward and reverse processes being equal. The criteria for these equilibria are grounded in fundamental thermodynamic laws. Thermal equilibrium follows from the , which states that if two systems are each in with a third system, then they are in with each other; this transitivity defines as an empirical property. Mechanical equilibrium is ensured by the uniformity of and the absence of stresses or gradients that would drive . For , particularly in multiphase systems or reacting mixtures, the key criterion is the equality of chemical potentials (\mu) for each component across phases or between reactants and products, ensuring no net transfer of matter. A profound thermodynamic links to maximization, as dictated by the second law. For an , the state corresponds to the condition of maximum (S), where any increases until this maximum is attained, and no further changes occur. This maximization reflects the most probable distribution of microscopic states consistent with the system's constraints. Diathermic walls, which permit , can facilitate the approach to by allowing temperature equalization. Representative examples illustrate these concepts. Consider an confined in an insulated container: after initial mixing or expansion, it reaches with uniform temperature and , exemplifying thermal and with maximum for the given volume. Another case is phase equilibrium in a closed vessel containing at its under constant , where liquid and vapor phases coexist stably because their chemical potentials for are equal, preventing net or .

Spontaneous Processes

In thermodynamics, spontaneous processes in systems are driven by internal imbalances, such as or gradients, leading to state changes without imposed external work beyond associated with volume changes. These processes evolve irreversibly toward , generating internally due to dissipative effects, without mechanisms like shafts or stirrers introducing additional mechanical work from the surroundings. Key examples include irreversible expansions and heat diffusion. In free expansion, a gas in an insulated container expands into an adjacent upon partition removal, performing no work and exchanging no heat; for an , internal energy remains constant, but increases as the system reaches uniformity. Similarly, heat diffusion in a , such as in a solid rod, flows from hotter to cooler regions, producing locally according to \sigma = -\frac{J_q}{T^2} \frac{dT}{dx} > 0, where J_q is the and \frac{dT}{dx} is the gradient, reflecting irreversible thermal dissipation. Illustrative cases include gas leaking into a , where molecules diffuse spontaneously without work input, leading to uniform and increased . Spontaneous mixing of fluids, such as two gases in an insulated mixing upon partition removal, proceeds via driven by concentration gradients, resulting in entropy increase without or work exchange. These dynamics resolve internal imbalances to reach .

Driven Processes

In non-equilibrium thermodynamics, systems can be driven by internal devices or external fields performing work beyond boundary displacement, such as shaft work from rotating elements like paddles or turbines, or field-induced work in gravitational, electric, or magnetic contexts. This contrasts with spontaneous processes limited to pressure-volume changes. In the first law, the work term includes such contributions: \Delta U = Q - (W_\text{boundary} + W_\text{other}), where W_\text{other} accounts for these mechanisms. For magnetic fields, the work is often \delta W = \mu_0 \mathbf{H} \cdot d\mathbf{M}, with \mathbf{H} the magnetic field strength and \mathbf{M} the magnetization, enabling cycles like magnetic refrigeration. Driven systems maintain operation far from , such as steady-state flows with continuous work input, dissipating and generating excess . The rate \sigma arises from irreversible processes like viscous , quantified as \sigma = \sum J_k X_k > 0, where J_k are fluxes and X_k affinities, with work inputs amplifying . This supports functionality but is limited by second law constraints. Examples include stirred tank reactors, where impellers provide shaft work for homogenization and steady-state reactions. Systems in , like paramagnetic materials, undergo work during field changes, aiding applications like cryocoolers. Biological cells harness from ATP hydrolysis to drive molecular motors and ion pumps, sustaining non-equilibrium states for life processes. Many open systems use such driving for mixing or .

Special Cases and Extensions

Selective Matter Transfer

A thermodynamic system exhibiting selective matter transfer is characterized by boundaries that permit the passage of certain matter components while restricting others, typically facilitated by semi-permeable membranes. These membranes allow selective based on molecular size, charge, or , enabling processes such as where solvent molecules traverse the barrier in response to solute concentration differences across the . In applications involving multi-component equilibria, the Gibbs-Duhem equation provides a for maintaining thermodynamic . The equation arises from the of the G = \sum_i \mu_i n_i, where \mu_i is the and n_i the amount of component i. Differentiating yields dG = \sum_i n_i d\mu_i + \sum_i \mu_i dn_i, while the standard form is dG = -S dT + V dP + \sum_i \mu_i dn_i. Equating these and rearranging at constant composition gives the Gibbs-Duhem : S dT - V dP + \sum_i n_i d\mu_i = 0, which constrains changes in chemical potentials across phases separated by selective boundaries. Representative examples include bags, which function as semi-permeable enclosures allowing small solute molecules or ions to diffuse out while retaining larger macromolecules like proteins, thus purifying solutions through selective . In fuel cells, proton-exchange selectively transport hydrogen ions (protons) between electrodes while blocking electrons and other species, enabling efficient electrochemical energy conversion. A key quantitative aspect is , which counteracts solvent flow across the membrane and is given by \pi = i M R T, where \pi is the osmotic pressure, i the van't Hoff factor accounting for , M the molarity of the solute, R the , and T the . These systems drive critical separation processes in and by exploiting differential to isolate components, differing from non-ideal closed or open systems through their controlled, component-specific exchange that enhances efficiency in purification and concentration tasks.

Composite Systems

A composite thermodynamic system comprises multiple subsystems in mutual contact, such as through diathermal or adiabatic walls that permit selective exchanges of , work, or volume while potentially restricting , resulting in emergent macroscopic properties that depend on the of the components rather than individual subsystems alone. For instance, the total of the composite is the sum of the entropies of its subsystems, and is achieved when intensive variables like and equalize across permeable boundaries, maximizing the overall under given constraints. Analysis of such systems can proceed by treating the entire composite as a unified entity, applying laws to the aggregate properties, or by decomposing it into subsystems with separate and balances, accounting for interactions via shared interfaces that impose constraints like fixed volumes or impermeable partitions. This dual approach facilitates understanding how local equilibria in subsystems contribute to global behavior, such as uniform distribution after through a wall. Often, the bounding surface of the composite as a whole defines it as an with no net exchange with surroundings, simplifying the application of to internal processes. Prominent examples include multi-phase systems, where multiple phases (e.g., solid, liquid, gas) coexist within a composite framework, governed by the Gibbs phase rule that determines the degrees of freedom as F = C - P + 2, with C representing the number of independent components and P the number of phases, ensuring thermodynamic consistency across interfaces. Another key illustration is the heat engine, modeled as a composite involving a working substance interacting with a hot reservoir (supplying heat at high temperature) and a cold reservoir (absorbing exhaust heat at low temperature), where efficiency emerges from the temperature differential and cyclic processes across these subsystems. In modern extensions, non-equilibrium composite systems exhibit dissipative structures, where far-from-equilibrium interactions among subsystems—driven by continuous energy dissipation—generate spatiotemporal order and , as pioneered by in his work on irreversible processes leading to dynamic stability. These structures, such as networks or cells, highlight how composites can sustain through nonlinear couplings, contrasting with composites and addressing gaps in classical .

References

  1. [1]
    1.2 Definitions and Fundamental Ideas of Thermodynamics - MIT
    A thermodynamic system is a quantity of matter of fixed identity, around which we can draw a boundary (see Figure 1.3 for an example).
  2. [2]
    3.1 Thermodynamic Systems – University Physics Volume 2
    A thermodynamic system includes anything whose thermodynamic properties are of interest. It is embedded in its surroundings or environment.
  3. [3]
    [PDF] Thermodynamic Systems and Processes
    A thermodynamic system is a collection of particular atoms, molecules or other fundamental particles that is large enough to allow definition of the average ...
  4. [4]
    What is Thermodynamics? - Glenn Research Center - NASA
    Mar 2, 2023 · Thermodynamics is a branch of physics which deals with the energy and work of a system. It was born in the 19th century as scientists were first discovering ...
  5. [5]
    Thermodynamic systems and properties
    In thermodynamics the totality of the world is usually termed the universe, and is divided into two parts. 1. The system is the part of the universe in which ...
  6. [6]
    Systems, Surroundings, and the Universe - Chemistry 301
    Systems, Surroundings, and the Universe. In thermodynamics we follow energy flowing from one place to another during chemical or physical changes.<|control11|><|separator|>
  7. [7]
    7.2 Microscopic and Macroscopic Descriptions of a System - MIT
    The microscopic description of a system is the complete description of each particle in this system.
  8. [8]
    [PDF] From Microscopic to Macroscopic Behavior - Princeton University
    Thermodynamics is concerned only with macroscopic quantities and ignores the microscopic variables that characterize individual molecules.
  9. [9]
    The first law of thermodynamics - Physics
    Dec 8, 1999 · A good example of a thermodynamic system is gas confined by a piston in a cylinder. If the gas is heated, it will expand, doing work on the ...
  10. [10]
    3.6 Adiabatic Processes for an Ideal Gas - UCF Pressbooks
    In an adiabatic process, an ideal gas's temperature increases during compression and decreases during expansion, with no heat flow.
  11. [11]
    Definition of Equation of State (EOS) | PNG 520
    Assuming an equilibrium state, the three properties needed to completely define the state of a system are pressure (P), volume (V), and temperature (T).
  12. [12]
    [PDF] Thermodynamics and Statistical Mechanics - Rutgers Physics
    What are some of the state variables that characterize the system? One could think of amount of gas (in moles), volume, pressure, temperature, total energy, ...
  13. [13]
    First Law of Thermodynamics - HyperPhysics
    The first law of thermodynamics is the application of the conservation of energy principle to heat and thermodynamic processes.
  14. [14]
    2.3 First Law of Thermodynamics | METEO 300 - Dutton Institute
    The First Law of Thermodynamics tells us how to account for energy in any molecular system, including the atmosphere.
  15. [15]
    June 12, 1824: Sadi Carnot Publishes Treatise on Heat Engines
    May 26, 2009 · In 1824 he published Reflections on the Motive Power of Fire, which described a theoretical “heat engine” that produced the maximum amount of work for a given ...
  16. [16]
    The Second Law: From Carnot to Thomson-Clausius, to the Theory ...
    At its origins, thermodynamics was the study of heat and engines. Carnot transformed it into a scientific discipline by explaining engine power in terms of ...
  17. [17]
    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 ...
  18. [18]
    Exploring the Science and History of Thermodynamics - NYAS
    May 1, 2006 · Clausius saw that something was being conserved in Carnot's perfectly reversible engine – but something other than heat. He called it entropy, ...
  19. [19]
    Kelvin: History | NIST - National Institute of Standards and Technology
    May 14, 2018 · In 1848, he published a paper, On an Absolute Thermometric Scale, that stated that this absolute zero was, in fact, -273 degrees Celsius.Missing: significance | Show results with:significance
  20. [20]
    Historic Note No. 1: Gibbs' Phase Rule - Thermo-Calc Software
    This article is the first in a three-part series tracing the evolution of CALPHAD from the Gibbs Phase Rule to the first lattice stabilities.Missing: 1876 | Show results with:1876
  21. [21]
    Boltzmann's Work in Statistical Physics
    Nov 17, 2004 · The 1872 paper contained the Boltzmann equation and the H-theorem. Boltzmann claimed that the H-theorem provided the desired theorem from ...Missing: 1870s | Show results with:1870s
  22. [22]
    Ilya Prigogine (1917–2003): Structure Formation Far from Equilibrium
    Jul 24, 2003 · Prigogine built on De Donder's results and in 1945–1949 worked out a general macroscopic theory of systems that operate under constraints that ...
  23. [23]
    [PDF] prigogine-lecture.pdf - Nobel Prize
    To obtain a thermodynamic theory for this type of structure we have to show that that non-equilibrium may be a source of order. Irreversible processes may lead.Missing: 1940s- | Show results with:1940s-
  24. [24]
    2.1: The System, Surroundings, and Boundary
    ### Definitions and Summaries
  25. [25]
    [PDF] III. Thermodynamic Phenomenology
    Closed systems have walls that are impermeable for particles but permeable for energy. The corre- sponding containers have diathermic walls. For such systems, ...
  26. [26]
    Chemical thermodynamics | McGraw Hill's AccessScience
    A boundary which is restrictive to one component of a system but not to the other components is called a semipermeable wall or membrane. A system whose ...
  27. [27]
    [PDF] DOE Fundamentals Handbook Thermodynamics, Heat Transfer, and ...
    Everything external to the system is called the thermodynamic surroundings, and the system is ... However, to apply these relationships, one additional definition ...
  28. [28]
    [PDF] LECTURE NOTES ON THERMODYNAMICS
    May 17, 2025 · These are lecture notes for AME 20231, Thermodynamics, a sophomore-level undergraduate course taught in the Department of Aerospace and ...<|separator|>
  29. [29]
    [PDF] The Second Law of Thermodynamics Thermal Energy Reservoirs
    Thermal energy reservoirs are hypothetical bodies with a relatively large thermal energy capacity (mass x specific heat) that can supply or absorb finite ...Missing: mechanical | Show results with:mechanical
  30. [30]
    [PDF] Introduction to Thermodynamics Definitions - Purdue Engineering
    Why take Thermodynamics I? • Develop basic problem solving, organizational ... surroundings. • A control volume (CV) (aka open system) is a particular ...<|control11|><|separator|>
  31. [31]
    1.2 System and surroundings
    A boundary may be fixed or movable, real or imaginary, rigid or flexible. A ... system Figure 1.2.3 Piston cylinder device as an example of closed systems.Missing: adiabatic impermeable
  32. [32]
    [PDF] System, Surrounding and Boundary • State, Property and Process
    Thermodynamics deals with the science of “motion” (dynamics) and/or the transformation of “heat” (thermo) and energy into various other energy–.
  33. [33]
    Entropy and the 2nd & 3rd Laws of Thermodynamics
    Chemical Thermodynamics ; Second Law: In an isolated system, natural processes are spontaneous when they lead to an increase in disorder, or entropy.
  34. [34]
    Entropy - Richard Fitzpatrick
    Thus, the entropy of an isolated system tends to increase with time and can never decrease. This proposition is known as the second law of thermodynamics. One ...
  35. [35]
    [PDF] Lecture 5. Entropy and the Second Law (Ch. 2 ) - Rutgers Physics
    2nd law: The entropy of an isolated system never decreases. Nernst's Theorem: The entropy of a system at T = 0 is a well-defined constant. For any processes ...
  36. [36]
    Thermodynamic systems - Bluffton University
    The physical universe is an isolated system; a closed thermos bottle is essentially an isolated system (though its insulation is not perfect).
  37. [37]
    [PDF] ESCI 341 – Atmospheric Thermodynamics Lesson 2 – Definitions
    energy with its surroundings. Examples would be: o The inside of a ThermosTM ... This is what the subject of thermodynamics attempts to do. Page 4. 4.
  38. [38]
    [PDF] Thermodynamics
    The universe is the ultimate isolated system; therefore, the total energy of the universe increases during every natural process.
  39. [39]
    Energy Balance for Closed Systems – Thermodynamics
    The First Law of Thermodynamics applied to stationary closed systems as a conservation of energy principle. For a closed system (no mass transfer) process ...
  40. [40]
    [PDF] 1st Law for Closed Systems Guide - ME 201
    First Law for Closed Systems Guide. Recall that for a closed system there are no mass inflows or mass outflows, so the mass of the closed system remains ...
  41. [41]
    Chapter 3a - Closed Systems - Energy (updated 1/17/11)
    Jan 17, 2011 · Chapter 3: The First Law of Thermodynamics for Closed Systems. a) The Energy Equation for Closed Systems. We consider the First Law of ...
  42. [42]
    first law of thermodynamics: conservation of energy - MIT
    On a p-v diagram for a closed-system sketch the thermodynamic paths that the system would follow if expanding from volume = v1 to volume = v2 by isothermal ...
  43. [43]
    [PDF] Some notes on the First Law of Thermodynamics for a closed system
    The first law of thermodynamics is a statement of the conservation of energy. : represents the net heat interactions. : represents the net work interaction.
  44. [44]
    Thermodynamics of Systems of Constant Composition (Closed ...
    From our basic courses in thermodynamics, we recall that the first law of thermodynamics for a closed system is written as follows: ∂W=PdV. (14.1). where,. U ...
  45. [45]
    3.4 Thermodynamic Processes – University Physics Volume 2
    There are several types of thermodynamic processes, including (a) isothermal, where the system's temperature is constant; (b) adiabatic, where no heat is ...
  46. [46]
    Thermodynamic processes
    An adiabatic process is a process during which no heat enters or leaves the system. · An isobaric process is a process that occurs at constant pressure. · An ...
  47. [47]
    The First Law of Thermodynamics and Some Simple Processes
    Key thermodynamic processes include isobaric (constant pressure), isochoric (constant volume), isothermal (constant temperature), and adiabatic (no heat ...
  48. [48]
    15.2 The First Law of Thermodynamics and Some Simple Processes
    Some thermodynamic processes, including isothermal and adiabatic processes, are reversible in theory; that is, both the thermodynamic system and the environment ...
  49. [49]
    [PDF] 249 Chapter 7 Heat, Work, and the First Law of Thermodynamics ...
    Aug 7, 2021 · Examples of constant volume systems include autoclaves, kitchen canning jars, closed aerosol spray cans, and capped soda bottles. The most ...
  50. [50]
    [PDF] Sample problems
    Classify the thermodynamic systems described below as open, closed, or isolated: ... The system is a solution contained in a sealed, thermally insulated flask.
  51. [51]
    [PDF] Thermodynamics
    The microstate of a macroscopic system at thermodynamic equilibrium is permanently changing. However, these changes are irrelevant, whereas the important ...<|separator|>
  52. [52]
    Thermodynamic Foundations – Introduction to Aerospace Flight ...
    Thermodynamics is the science of energy and the principles governing its transformation from one form to another, such as work and heat, and the production of ...
  53. [53]
    Continuity Equation – Introduction to Aerospace Flight Vehicles
    It is a general governing equation applicable to three-dimensional, unsteady flows, covering all types, including compressible or incompressible, viscous or ...
  54. [54]
    [PDF] Notes on Thermodynamics, Fluid Mechanics, and Gas Dynamics 5.2 ...
    Dec 15, 2021 · The Continuity Equation, which is Conservation of Mass for a differential fluid element or control volume, can be derived several different ...Missing: open system
  55. [55]
    Energy Balances — Introduction to Chemical and Biological ...
    Energy balances determine energy requirements and temperatures in systems, tracking energy transfer, and considering internal, kinetic, and potential energy.
  56. [56]
    Steady Flow Energy Equation - MIT
    In an open flow system, enthalpy is the amount of energy that is transferred across a system boundary by a moving flow.
  57. [57]
    [PDF] Unit Five – First Law for Steady, Open Systems - CSUN
    Sep 23, 2010 · Topic is first law for open systems, i.e., systems in which mass flows across the boundary. • Will look at general results and focus on.
  58. [58]
    Steady Flow Energy Equation
    The first law relates the change in energy between states 1 and 2 to the difference between the heat added and the work done by the system.
  59. [59]
    Thermodynamic Equilibrium - Objectives_template
    In thermodynamics, equilibrium refers to a state of equilibrium with respect to all possible changes, thermal, mechanical and chemical.
  60. [60]
    Thermodynamic equilibrium – Knowledge and References
    A system is said to be in thermodynamic equilibrium if it is in thermal, mechanical, and chemical equilibrium with itself and with the surroundings.
  61. [61]
    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 ...
  62. [62]
    4: Chemical Potential and Equilibrium - Chemistry LibreTexts
    Aug 23, 2023 · Equilibrium occurs when the total chemical potential of products equals that of reactants. Pressure changes affect equilibrium composition, but ...
  63. [63]
    Chemical equilibrium | McGraw Hill's AccessScience
    When equilibrium is reached, the total chemical potentials of products and reactants become equal. For reactions at constant temperature and pressure, the ...
  64. [64]
    17: Equilibrium and the Second Law of Thermodynamics
    Aug 14, 2020 · Third, the entropy of a substance increases with increasing temperature. The temperature is, of course, a measure of the average kinetic energy ...
  65. [65]
    Equilibrium and the Second Law of Thermodynamics - Physics
    The entropy of an isolated system never decreases: in equilibrium, the entropy stays the same;; otherwise the entropy increases until equilibrium is reached.
  66. [66]
    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.
  67. [67]
    1.10: Phase Equilibria and Phase Diagrams - Chemistry LibreTexts
    May 3, 2023 · A typical binary phase diagram (Figure 1) indicates those phases present in equilibrium at any particular temperature and composition at the ...
  68. [68]
    4.2 Difference between Free and Isothermal Expansions - MIT
    The overall result when we have restored the system to the initial state, however, is quite different for the reversible expansion than for the free expansion.
  69. [69]
    None
    Summary of each segment:
  70. [70]
    First Law - Conservation of Energy | Glenn Research Center - NASA
    Jul 18, 2024 · Some of the work, called the shaft work (wsh), opposes the motion of the cylinder, while the rest of the work goes into changing the state of ...
  71. [71]
    [PDF] The Form of Magnetic Work in a Fundamental Thermodynamic ...
    For a paramagnetic crystal in a uniform magnetic field B, with total mag- netic dipole moment M, there are two forms for the work done when B and M change ...
  72. [72]
    Thermodynamics based stability analysis and its use for nonlinear ...
    The first one is an example of single reaction system subject to steady-state ... Quadratic autocatalysis in an extended continuous-flow stirred tank reactor ...
  73. [73]
    Entropy Production: Its Role in Non-Equilibrium Thermodynamics
    In this work we will review the ideas behind the entropy production concept and we will give some insights about its relevance.Missing: shaft | Show results with:shaft
  74. [74]
    Non-equilibrium Thermodynamical Principles for Chemical ...
    Keywords. reaction networks · non-equilibrium thermodynamics · interacting ... stirred tank reactor. In such reactors, all species are deemed to be present ...
  75. [75]
    Non-equilibrium thermodynamics and the free energy principle in ...
    Aug 23, 2021 · The motivation for Definition 1 [ active systems] is simple: systems ... biological significance because they are not observed in actual cells.
  76. [76]
    Osmosis and thermodynamics explained by solute blocking - PMC
    It is the selective transport of water across a semipermeable membrane from high to low chemical potential caused by a difference in solute concentrations and/ ...
  77. [77]
    [PDF] Osmotic pressure and mechanics of cell membranes
    Small water molecules can pass through a semipermeable membrane, which blocks large solute macromolecules. In thermodynamic equilibrium the Gibbs free energy G ...
  78. [78]
    Dialysis Methods for Protein Research | Thermo Fisher Scientific - US
    Dialysis is a separation technique that facilitates the removal of small, unwanted compounds from macromolecules in solution by selective and passive diffusion.
  79. [79]
    [PDF] Fuel Cell Handbook (Seventh Edition)
    ... thermodynamics of fuel cells to provide an understanding of fuel cell operation. Sections 3 through 7 describe the five major fuel cell types and their ...
  80. [80]
    Membrane Separation Performance in Energy, Water, and Industrial ...
    Gains by selective mechanisms other than size would have the greatest impact in rubbery polymers. 4. Contribution of Driving Forces to Separation Processes.<|control11|><|separator|>
  81. [81]
    Making wastewater obsolete: Selective separations to enable ...
    Jan 6, 2021 · In particular, wastewater will be mined for the valuable product precursors it contains, which will require highly selective separation ...
  82. [82]
    [PDF] Review of Classical Thermodynamics and VLE
    A composite system consists of one or more simple systems. For example, two closed containers containing a mixture of gases considered together is a composiste.
  83. [83]
  84. [84]
  85. [85]
    Time, Structure, and Fluctuations - Science
    It is shown that nonequilibrium may become a source of order and that irreversible processes may lead to a new type of dynamic states of matter called ...Missing: composites | Show results with:composites