Fact-checked by Grok 2 weeks ago

Exergy

Exergy is a fundamental concept in that quantifies the maximum amount of useful work a can produce as it is brought into complete with its reference , effectively measuring the or "usefulness" of rather than its total quantity. Unlike , which is conserved according to of , exergy is not conserved and is inevitably destroyed due to irreversibilities in real processes, as governed by the second law. This property makes exergy a powerful tool for assessing the and potential improvements in energy conversion systems, where it highlights losses that simple energy balances overlook. The origins of exergy trace back to early 19th-century developments in , such as Sadi Carnot's work on heat engines in 1824, which emphasized the role of temperature differences in work production, and ' 1873 formulation of "available energy" as the maximum work extractable from a system interacting with its surroundings. The modern term "exergy" was coined in 1956 by Slovenian mechanical engineer Zoran Rant, derived from the Greek words ex (out) and ergon (work), to describe the "technical working capacity" of energy systems. By the mid-20th century, following the , exergy analysis gained prominence as a method for optimizing resource use in power plants, chemical processes, and industrial operations, with early applications documented in works like Obert and Birnie's 1949 study on steam power cycles. Mathematically, the specific physical exergy \psi of a is often expressed as \psi = (h - h_0) - T_0(s - s_0), where h and s are the specific and at the system's state, subscript 0 denotes the reference environment state, and T_0 is the environmental temperature; the total exergy includes this physical component plus separate contributions from kinetic, potential, and chemical effects. In practice, exergy balances are applied to evaluate efficiencies in diverse fields, including where it assesses fuel utilization in systems, and sustainable where it optimizes low-grade heating and cooling to minimize environmental impact. Exergy-based metrics, such as exergetic efficiency defined as the ratio of exergy output to input, provide a unified for comparing energy technologies and promoting by focusing on the degradation of energy quality.

Definitions and Fundamentals

Terminology and Basic Concepts

Exergy represents the maximum amount of useful work that can be extracted from a as it interacts with its reference environment to reach complete , often described as the available or work potential inherent in the system. Unlike total , which is conserved according to of , exergy is not conserved and is inevitably destroyed during irreversible processes due to entropy generation as per the second law. The reference environment, also known as the , defines the baseline for exergy calculations as the state of where the system shares the same , , and chemical potentials as the ambient surroundings, rendering no further useful work possible. In this , the system is in thermal, mechanical, and with the environment, eliminating all gradients that could drive work extraction. Exergy can be categorized into physical exergy, which arises from differences in and relative to the reference ; chemical exergy, which stems from differences in or concentration compared to the environmental ; kinetic exergy, due to the system's relative to the ; and potential exergy, arising from the system's position in a gravitational or other . Physical exergy quantifies the work potential from thermomechanical disequilibria, while chemical exergy accounts for the potential from compositional variations, such as in reactive mixtures or fuels. The term "exergy" derives from the Greek roots "ex" meaning "out of" or "from" and "ergon" meaning "work," and was coined in the 1950s to encapsulate this concept of extractable work.

Physical Implications

Exergy serves as a fundamental measure of energy quality, quantifying the maximum useful work that can be extracted from a system relative to a reference environment, thereby distinguishing between high-quality energy forms capable of performing work and low-quality forms that cannot. For instance, electricity and mechanical work represent high-exergy energy carriers due to their near-complete convertibility into useful work, whereas ambient heat at environmental temperature possesses negligible exergy because it is already in equilibrium with the surroundings and offers no potential for work extraction. In alignment with the second law of thermodynamics, exergy analysis reveals that all real-world processes involve irreversibilities, such as , across finite differences, and mixing, which inevitably lead to exergy destruction and a degradation of quality. This destruction quantifies the thermodynamic inefficiencies inherent in practical systems, providing a clearer picture of losses than first-law balances alone, as it accounts for the generation that renders less available for work. Exergy's emphasis on quality underscores the scarcity and value of high-grade resources in the context of resource depletion, as natural high-exergy sources like concentrated fuels or minerals are finite and their degradation through use contributes to broader environmental and economic constraints. By framing resource consumption in terms of exergy loss, this approach highlights why low-entropy, high- energy is inherently more valuable and prone to rapid depletion compared to abundant but low-quality forms like solar radiation dispersed at ambient conditions. Exergy efficiency, defined as the ratio of useful exergy output to exergy input, offers a superior metric for evaluating and comparing process performance over traditional , which conserves quantity but ignores quality degradation. This second-law-based measure better identifies opportunities for improvement by pinpointing where exergy destruction occurs, enabling more rational assessments of in energy conversion and utilization systems.

Illustrative Examples

One illustrative example of physical exergy involves a quantity of hot water maintained at a temperature above the ambient environment, such as a tank of water at 70°C in a room at 25°C. In this scenario, the exergy represents the maximum useful work that could theoretically be extracted by operating a reversible heat engine between the hot water and the reference environment, gradually cooling the water to ambient temperature while rejecting heat to the surroundings. If the hot water simply cools passively without work extraction, all of this exergy is destroyed through irreversible heat transfer, highlighting how temperature differences relative to the environment determine available work potential. Another example demonstrates chemical exergy in the discharging of a , where stored is converted into electrical work. Consider a car used to start an : the chemical exergy inherent in the difference between the electrodes enables the production of electrical current to power the starter motor, performing useful work against the . This process illustrates how chemical exergy, tied to the 's relative to environmental substances, can be harnessed directly for mechanical or electrical tasks, with minimal destruction if the discharge is reversible. Fuel combustion provides a case involving both chemical and thermal exergy, where the high chemical exergy of the is partially converted to exergy but largely destroyed due to irreversibilities. For instance, in the burning of in an , the fuel's chemical exergy—arising from its molecular structure differing from environmental compounds—is released rapidly, generating hot combustion products with some exergy for work . However, approximately 20% of the fuel's chemical exergy is typically destroyed during the itself through from mixing, temperature gradients, and chemical reactions, reducing the overall compared to a reversible . Solar radiation exemplifies exergy in radiative processes, characterized by its high-quality originating from the Sun's surface of about 5800 . Incident solar radiation on carries substantial exergy—roughly 93% of its content—due to the concentrated, high-temperature blackbody that allows for efficient to work, such as in photovoltaic or systems. In contrast, the low-temperature re-radiated by Earth's surface at around 300 possesses much lower exergy, often less than 10% of its , as it is closer to with the environment and yields minimal work potential.

Historical Development

Early Thermodynamic Foundations

The 19th-century transition in from the , which posited as an indestructible , to the principle of fundamentally reshaped understandings of transformations and paved the way for concepts of available work. This shift was driven by experimental evidence, such as James Prescott Joule's demonstrations of generation through mechanical work in the 1840s, and theoretical advancements that integrated and work as interchangeable forms of . By the 1850s, and William Thomson () had formalized the first law of , emphasizing conservation over the caloric model's limitations, which had constrained earlier analyses of engines and processes. A pivotal early contribution came from Sadi Carnot's 1824 publication, Reflections on the Motive Power of Fire, which analyzed the efficiency of engines operating between hot and cold reservoirs, even while adhering to . Carnot introduced the ideal reversible , demonstrating that the maximum work extractable from is limited by the temperature difference, with efficiency given by the ratio of these temperatures on an . This work established fundamental limits on reversible work from engines, influencing later thermodynamic principles without yet invoking explicitly. Building on these foundations, developed the concept of in the 1870s as a measure of the maximum reversible work available in a system at constant temperature and pressure. In his 1873 and 1876 papers, Gibbs termed this "available energy," defining it as the portion of a system's that can perform useful work, excluding that dissipated as heat due to increase. This potential captured the spontaneity of processes under isothermal-isobaric conditions, providing a criterion for and phase transitions. Complementing Gibbs' work, introduced the in during a lecture on the of chemical processes, framing it as the maximum work obtainable from a at constant temperature and volume. Defined as the minus the temperature-entropy product, this function quantified the useful work potential in processes like chemical reactions, where volume is fixed, and highlighted affinities driving reversibility. Helmholtz's formulation extended availability ideas to non-expansion work, solidifying as a cornerstone for analyzing thermodynamic potentials.

Formulation of Exergy Concept

The concept of exergy emerged as a unified term in the mid-20th century, primarily driven by the need to quantify useful work potential in thermodynamic systems amid post-World War II reconstruction efforts in energy-constrained . Following the devastation of the war, European engineers and scientists sought tools to optimize limited resources for industrial recovery, shifting focus from mere quantity to its quality and convertibility into work. This context propelled the formalization of exergy as a practical metric for analysis in power plants, chemical processes, and systems. In 1956, Slovenian mechanical engineer Zoran Rant introduced the term "exergy" (from ex meaning "out" and ergon meaning "work") in German technical literature to denote "technische Arbeitsfähigkeit," or technical work capacity. Rant's proposal replaced disparate earlier terms like "" and "work potential," which had roots in 19th-century by figures such as Gibbs and Helmholtz, providing a standardized for the maximum useful work extractable from a relative to its . His seminal papers in Forschung im Ingenieurwesen emphasized exergy's role in evaluating irreversibilities and resource utilization, marking the birth of exergy as a distinct analytical concept. By the , exergy gained adoption in textbooks and curricula, particularly in and the , as researchers like Peter Grassmann, Wladimir Brodyansky, and Myron Tribus expanded its framework for broader applications. Grassmann's works in integrated exergy into , while Tribus and others in the U.S. promoted it through educational texts, solidifying its transition from a niche idea to a core principle in applied . This period saw exergy evolve into a systematic tool for comparing energy forms, unifying previous availability-based approaches under a single, environmentally referenced . The 1970s oil crises further accelerated interest in exergy, as soaring fuel prices and supply disruptions in and globally heightened demands for rational use. Engineers like Jan Szargut and Hans Fratzscher applied exergy analysis to industrial optimization, revealing inefficiencies in conversion and inspiring policies for . This era marked exergy's maturation into a framework for , bridging historical work potential concepts with modern efficiency imperatives.

Mathematical Formulation

General Definition and Second Law Basis

Exergy represents the maximum amount of useful work that can be extracted from a as it reaches complete with a specified reference through a reversible process, thereby quantifying the system's deviation from thermodynamic and its potential for conversion into work. This concept is intrinsically tied to the second of thermodynamics, which dictates that irreversible processes generate and degrade the quality of , limiting the extractable work to less than the total content. The general mathematical formulation of exergy for a system, relative to the reference environment denoted by subscript 0 (with temperature T_0 and pressure P_0), is given by Ex = (U - U_0) - T_0 (S - S_0) + P_0 (V - V_0) + \sum_i \mu_{i,0} (N_i - N_{i,0}), where U, S, and V are the internal energy, entropy, and volume of the system; \mu_{i,0} and N_{i,0} are the chemical potential and number of moles of species i in the reference environment; and the summation accounts for chemical contributions. This expression captures the work potential arising from differences in thermal, mechanical, and chemical properties between the system and the reference state, where exergy vanishes at equilibrium. The derivation stems directly from the first and second laws applied to a reversible bringing the system to the dead state. For a , the first law yields W = Q - (U_0 - U), where W is the work output and Q is absorbed by the . The second law, for reversibility, imposes Q = T_0 (S_0 - S), and accounting for useful work excluding atmospheric boundary work P_0 (V_0 - V), results in the maximum reversible work W_\text{rev} = (U - U_0) - T_0 (S - S_0) + P_0 (V - V_0), equivalent to exergy; any irreversibility reduces this by T_0 times the generation. For open systems, the formulation shifts to flow exergy using , Ex_f = (H - H_0) - T_0 (S - S_0) + \text{kinetic and potential terms}, reflecting steady-flow processes where mass crosses boundaries. Exergy is synonymous with the older term "availability" in thermodynamics, both denoting the same measure of useful work potential relative to the environment, though exergy more explicitly encompasses chemical potentials and is preferred in modern analyses. Conceptually, exergy embodies a inherent "potential" embedded in every thermodynamic state, extending beyond mere engine cycles to characterize the quality of any energy form or system deviation from environmental equilibrium.

Exergy for Heat, Work, and Mechanical Systems

In thermal systems, the exergy associated with Q from a at T relative to the environmental T_0 is given by the expression X_Q = Q \left(1 - \frac{T_0}{T}\right). This formula arises from the maximum work obtainable via a reversible Carnot engine operating between the heat source and the environment, where the term \left(1 - \frac{T_0}{T}\right) represents the Carnot factor, quantifying the fraction of that can be converted to useful work. For reservoirs, this exergy measures the potential to perform work while rejecting the unavoidable to the surroundings at T_0. Pure work, such as electrical or shaft work, possesses full exergy content equivalent to its magnitude, meaning its exergy transfer is X_W = W, with a 100% exergy-to-energy ratio. This stems from the direct convertibility of work into other forms of useful without thermodynamic losses in conditions, limited only by practical irreversibilities like . Electrical and work sources thus represent the highest-quality forms in exergy analysis. For mechanical systems, the exergy contributions from pressure differences in compressible fluids are captured by X = (P - P_0) V, where P is the system , P_0 the environmental , and V the ; this term reflects the work potential from or against the atmosphere. Additionally, kinetic energy exergy is X_{KE} = \frac{V^2}{2} per unit , and potential energy exergy is X_{PE} = g z per unit , where V is , g is , and z is relative to the reference level. These forms retain 100% exergy content, as they can be fully transformed into work through reversible processes like turbines or brakes. In open systems involving flowing fluids, the specific flow exergy \psi (also called ) is expressed as \psi = (h - h_0) - T_0 (s - s_0) + \frac{V^2}{2} + g z, where h and s are the specific and at the system state, subscript 0 denotes environmental conditions, and the kinetic and potential terms are included as above. This formulation extends the closed-system exergy by incorporating flow work via the enthalpy difference, enabling assessment of the maximum work extractable from a stream brought to environmental . The term accounts for the irreversibility penalty due to interactions with the surroundings.

Chemical and Reactive Exergy

Chemical exergy quantifies the maximum useful work derivable from the difference in between a and its reference , achieved through reversible reactions that bring the into with environmental substances at constant and . This component of exergy is particularly relevant for reactive systems, such as fuels or materials, where potential arises from affinities for oxidation, , or other transformations into stable environmental forms like CO₂, H₂O, and N₂. Unlike physical exergy, which depends on and deviations, chemical exergy is independent of these for the but requires corrections for non-ideal conditions or concentrations. For a pure substance or , the standard chemical exergy is fundamentally tied to the standard of reaction:
ex_{ch}^0 = -\Delta G^0 + \text{corrections},
where \Delta G^0 represents the standard change for the hypothetical converting the substance (along with environmental ) to the stable components of the reference environment, such as the oxidation of carbon to CO₂ or to H₂O. The corrections account for non-standard , , or activity coefficients, ensuring the value reflects realistic conditions beyond the ideal 298.15 and 1 . This formulation stems from the second , as the reaction affinity drives the work potential, with \Delta G^0 directly linking thermodynamic to chemical disequilibrium. Seminal work by Szargut established this approach by modeling reactions based on a reference environment, emphasizing probable products like dissolved ions and gases.
Standard chemical exergy values for elements and compounds have been tabulated using this method, providing benchmarks for analysis. For instance, in a reference environment approximating Earth's atmosphere and oceans (with O₂ at 20.95% and CO₂ at about 0.033%), the standard chemical exergy of pure O₂ gas is 3.97 /, arising primarily from its concentration gradient relative to the environment, while CO₂ is 19.48 / due to its lower ambient of approximately 0.0003. These values, derived from constants and environmental compositions, illustrate how even ubiquitous possess exergy when concentrated. Comprehensive tables cover 49 elements and numerous compounds, with updates refining reference for accuracy; for example, iron () has a standard chemical exergy of 374.3 /, reflecting its potential for oxidation to environmental iron oxides. Such tables facilitate rapid assessment in without recalculating full paths each time. For ideal mixtures, such as gaseous fuels or solutions, the total chemical exergy combines reactive potentials with concentration effects:
Ex_{ch} = \sum_i n_i (\mu_i^0 - \mu_{i0}^0) + RT_0 \sum_i n_i \ln \left( \frac{y_i}{y_{i0}} \right),
where n_i is the number of moles of component i, \mu_i^0 and \mu_{i0}^0 are the standard chemical potentials of the component in the and , respectively, R is the , T_0 the reference temperature, and y_i, y_{i0} the mole fractions. The first term captures the inherent reactive exergy from compositional differences (analogous to -\Delta [G](/page/G)^0 for the ), while the logarithmic term accounts for diffusive work potential due to non-uniform concentrations, assuming ideal behavior. This equation applies to systems like air-fuel mixtures, where enhances overall availability.
The total exergy of a is the sum of its physical and chemical components, Ex = Ex_{ph} + Ex_{ch}, allowing separate evaluation of thermal-mechanical and compositional potentials. This decomposition is essential for processes involving changes or , as chemical exergy often dominates in fuels (e.g., methane's standard chemical exergy of 831.2 / far exceeds its physical counterpart at ambient conditions). By prioritizing the environment's composition—typically seawater-augmented air—calculations ensure consistency across global analyses.

Exergy in Radiative and Non-Equilibrium Processes

The exergy flux of for arbitrary spectra is determined by integrating the weighted by the local Carnot factor, accounting for the non-blackbody nature of the radiation through its . Specifically, the radiative exergy E is given by E = \int_0^\infty I_{\lambda} \left(1 - \frac{T_0}{T_b(\lambda)}\right) d\lambda, where I_{\lambda} is the at \lambda, T_0 is the environmental , and T_b(\lambda) is the corresponding to the local . This formulation extends classical exergy concepts to distributions by treating each monochromatic component as a reversible limited by its , enabling precise quantification of work potential in non-ideal radiative fields. For , which exhibits a broad, non-equilibrium approximating a blackbody at around 5800 but diluted by the Sun's finite angular size, the exergy is approximately % of the total energy at the top of Earth's atmosphere. With an incident of 1366 W/m², the corresponding exergy is about 1270 W/m², reflecting the high-quality, low-entropy nature of relative to terrestrial conditions. This ratio arises from the evaluation over the , where shorter wavelengths contribute disproportionately to exergy due to their higher effective temperatures. In non-equilibrium processes, exergy analysis extends beyond to systems where distributions deviate significantly from Maxwell-Boltzmann statistics, such as gases, plasmas, and biological media. For s in non-equilibrium states, exergy incorporates the of the radiation field, quantifying the work extractable from population inversions or lasing conditions as b = u - T_0 s - \mu_0 n, where u, s, and n are the , , and particle number densities, respectively, and \mu_0 is the reference . In plasmas, non-equilibrium exergy captures the disparity between and temperatures, enabling assessment of work potential in or processes through generalized functions that include kinetic and contributions. For biological systems, such as cellular , non-equilibrium exergy measures the dissipative structures maintaining far-from-equilibrium states, with applications in quantifying metabolic efficiency via rates exceeding baselines by factors of 10-100 in living tissues. Radiative heat transfer exergy is analyzed using components of the , which represents the directional in electromagnetic fields and extends to exergy as \mathbf{S}_e = \mathbf{S} \left(1 - \frac{T_0}{T_b}\right), where \mathbf{S} is the Poynting vector magnitude and T_b is the local . This decomposition allows tracking of exergy streams in participating media, such as combustion chambers or solar receivers, where irreversibilities arise from spectral and , reducing the transferable work by up to 20-30% compared to ideal blackbody . Seminal formulations by Petela emphasize that the Poynting-derived exergy flux aligns with second-law efficiency in undiluted radiation, providing a vectorial basis for optimizing radiative systems.

Properties and Analysis

Exergy Balance and Irreversibility

The exergy provides a framework for analyzing energy conversion processes by accounting for both the quantity and of , integrating the . For a , the balance is expressed as the change in exergy of the system equals the net exergy transfer by , work, and minus the exergy destroyed due to irreversibilities: \Delta Ex_{\text{system}} = Ex_{\text{in}} - Ex_{\text{out}} - Ex_{\text{destroyed}} where Ex_{\text{in}} and Ex_{\text{out}} represent exergy flows entering and leaving the system, and Ex_{\text{destroyed}} quantifies the loss of useful work potential. This equation highlights that while is conserved, exergy is not, due to inherent thermodynamic imperfections. Exergy destruction, often denoted as I or Ex_{\text{destroyed}}, directly measures irreversibility and is given by I = T_0 \sigma, where T_0 is the reference environment temperature and \sigma is the entropy generation rate. This relationship stems from the Gouy-Stodola theorem, which establishes that the lost available work in any process is proportional to the entropy produced, providing a quantitative link between thermodynamic inefficiency and environmental conditions. In practical terms, I \geq 0 always holds per the second law, with equality only for reversible processes. In engineering applications, the exergy is used to determine minimum work requirements for processes such as separation and . For separation systems, like of a binary , the minimum work input equals the change in exergy between the feed and product streams, W_{\min} = \Delta Ex, revealing opportunities to minimize losses from mixing and . Similarly, in cycles, the identifies exergy destruction in components like the and , where the minimum work for cooling is the exergy difference between the heat source and sink relative to the , guiding designs to reduce irreversibilities from drops and finite differences. Unlike the energy balance, which enforces conservation (\Delta E_{\text{system}} = E_{\text{in}} - E_{\text{out}}) and treats all forms of energy as equivalent, the exergy balance explicitly captures degradation due to irreversibilities such as friction, unrestrained expansion, mixing of dissimilar streams, and heat transfer across finite temperature gradients. This distinction allows exergy analysis to pinpoint where high-quality energy is wasted on low-quality tasks, enabling targeted improvements in process efficiency beyond what first-law analysis provides.

Quality Metrics for Energy Forms

Exergy serves as a fundamental quality metric for evaluating different forms of , distinguishing it from the first-law concept of by incorporating the irreversibilities dictated by the second law of thermodynamics. The quality of an energy form is quantified by its potential to perform useful work when brought into with the reference environment, typically defined by ambient T_0 and P_0. This is captured through the exergy , often denoted as f = \frac{\text{exergy}}{\text{[energy](/page/Energy) content}}, which ranges from 0 (no useful work potential, as in ambient ) to 1 (full convertibility to work). For energy carriers beyond pure work, the exceeds 1 in some cases due to the structured nature of the , such as in chemical bonds. This metric enables comparisons across disparate energy types, highlighting why high-quality forms like are preferable for tasks requiring precise control, while low-quality forms like low- heat are limited in versatility. For and , the exergy factor is exactly 1, as these forms are fully convertible to work without thermodynamic losses under ideal conditions. Chemical energy in fuels also exhibits high quality, with exergy factors typically between 1.0 and 1.1, reflecting that the maximum work from or closely matches or slightly exceeds the heating value due to the changes involved. Representative values for common carriers are summarized below, based on standard chemical exergy calculations relative to lower heating values:
Energy CarrierExergy Factor
1.00
(e.g., kinetic, potential)1.00
Oil, petroleum products1.06
1.06
Fuel wood (20% )1.11
These values underscore the near-equivalence of chemical exergy to enthalpic content for fuels, making them high-quality carriers for power generation, though actual utilization efficiencies are lower due to process irreversibilities. Thermal energy's quality varies markedly with temperature, as its exergy is given by \text{Ex} = Q \left(1 - \frac{T_0}{T}\right), where Q is the heat quantity and T is the absolute temperature of the source. Thus, the exergy factor f = 1 - \frac{T_0}{T} approaches 1 for high-temperature heat (e.g., f \approx 0.70 at T = 1000 K with T_0 = 298 K) but nears 0 for heat near ambient conditions (e.g., f \approx 0.17 at T = 350 K). This temperature dependence illustrates why high-grade heat from combustion is valuable for work production via engines, while low-grade waste heat requires upgrading (e.g., via heat pumps) to enhance its utility. In practice, exergy analysis reveals significant degradation when high-quality energy is cascaded to lower temperatures, such as in heating systems where exergy efficiency drops to 20-40%. Radiative energy, such as , presents a more complex metric due to its non-equilibrium spectrum. The exergy factor for approximating the sun's surface (around 6000 ) is approximately 0.93-0.96, derived from integrating the distribution with the Carnot analogue for photons. For terrestrial incident , effective factors range from 0.70 to 0.85, accounting for atmospheric dilution and directionality, making it a moderate- that necessitates concentration or conversion technologies to approach work potential comparable to fuels. These metrics emphasize exergy's role in assessing renewable integration, where solar's inherent limits direct without auxiliary processes.

Total Exergy and System Potentials

Total exergy represents the comprehensive measure of a 's useful work potential relative to a reference environment, encompassing all contributing forms of exergy within the system. It is calculated as the sum of physical exergy (arising from and differences), chemical exergy (from compositional disparities), kinetic exergy (from ), potential exergy (from in a force ), and nuclear exergy (from isotopic differences), excluding minor effects like magnetic or unless relevant. This aggregation provides a holistic of the system's departure from environmental , enabling evaluation of its overall thermodynamic value. For instance, in a flowing fluid at elevated and altitude, the total exergy integrates driving forces with gravitational positioning to quantify extractable work. Exergy potentials persist in various states, including those at ambient conditions, where certain components remain non-zero due to relative differences from the reference state. exergy, for example, is present for any mass elevated above the defined reference level, expressed as m g z, where z is the difference, even if the is thermally equilibrated with the surroundings. Similarly, chemical potentials in ambient mixtures, such as atmospheric gases or compositions, retain exergy if their concentrations deviate from environmental , highlighting that dead states are context-specific and not universally zero. This relative nature underscores exergy's dependence on the chosen reference environment, often standardized to sea-level conditions at 25°C and 1 atm. In multi-component systems, total exergy extends to account for interactions among constituents, particularly in mixtures and phase-changing materials. For gaseous or mixtures, the chemical exergy is determined by the sum of partial exergies, \sum_i n_i (\mu_i - \mu_i^0), where \mu_i is the of component i and \mu_i^0 its standard value, incorporating non-ideal effects via activity coefficients rather than simple summation of individual exergies. In phase change scenarios, such as in or solidification, the total exergy includes contributions from the of transformation, capturing the work potential from transitioning between phases at non-equilibrium conditions. These extensions ensure accurate representation in complex setups like alloys or multiphase flows, where synergies or antagonisms among components influence the aggregate potential. A key limitation of total exergy lies in its non-additivity during interactions between subsystems, as irreversibilities cause mutual exergy destruction that reduces the combined potential below the sum of isolated values. For example, when two mix irreversibly, the exergy loss due to exceeds the individual contributions, reflecting generation per the second law. This property necessitates careful analysis in , preventing overestimation of potentials in coupled processes.

Applications

Engineering Design and Efficiency

Exergy serves as a critical tool in engineering for identifying and minimizing irreversibilities in power generation systems, enabling more efficient resource use. In steam power plants, this involves applying the exergy balance to components like and turbines, where destruction is quantified as the difference between exergy input and output. The process in typically exhibits the highest exergy destruction due to chemical irreversibilities and temperature mismatches during , often accounting for 50-80% of the total plant destruction. For instance, in a conventional steam power plant, and within the contribute approximately 53.83% of overall exergy destruction, while turbines account for about 13-20%, primarily from and losses. Design improvements leverage exergy principles to align the quality of supply with demand, reducing avoidable losses. In systems, high-exergy from is first converted to work in turbines, with residual low-exergy cascaded for process or , achieving exergy efficiencies of 25-35% compared to 15-25% for separate power and production. This matching minimizes destruction by avoiding the use of high-grade for low-grade tasks, as demonstrated in combined heat and power plants where exergy-based optimization increases overall by up to 15%. Exergy flow diagrams, analogous to Sankey diagrams but emphasizing destruction hotspots, facilitate this by visually mapping exergy streams across components, guiding iterative design refinements. Specialized software enhances these analyses by simulating exergy flows and testing design variants. Tools like EBSILON Professional enable detailed modeling of power plant cycles, incorporating exergy balances to optimize staging and configurations for minimal destruction. Similarly, ProSimPlus supports steady-state simulations of complex processes, calculating component-specific exergy efficiencies and suggesting improvements like advanced heat exchangers. A practical case is the automotive , where remains below 30% primarily due to irreversible rejection. In a typical , fuel chemical exergy input yields only about 20-25% as useful mechanical work exergy, with over 60% destroyed or rejected as low-grade exhaust and , highlighting opportunities for recovery via thermoelectric or Rankine cycles. Exergy analysis reveals that irreversibilities alone destroy around 25-30% of fuel exergy, informing designs like to better match load conditions.

Resource Utilization and Sustainability

Exergy plays a central role in resource accounting by quantifying the total useful energy extracted from the natural environment, enabling a thermodynamic assessment of depletion rates for both renewable and non-renewable resources. The Cumulative Exergy Extraction from the Natural Environment (CEENE) method provides a comprehensive framework for life cycle impact assessment, aggregating exergy inputs across upstream processes such as exploration, extraction, and initial processing. For fossil fuels, CEENE accounts for the high chemical exergy inherent in hydrocarbons; for instance, crude oil has a standard chemical exergy of approximately 45 MJ/kg, reflecting its concentrated potential for work relative to the reference environment. Minerals and metal ores are evaluated similarly, with cumulative exergy extraction coefficients capturing the embodied exergy in beneficiation and concentration, such as 20-30 MJ/kg for iron ore depending on grade. This approach highlights the irreversibility of resource use, where global annual exergy extraction from fossils exceeds 10^20 J, underscoring the scale of depletion for non-renewable sources. Sustainability metrics based on exergy emphasize costs, comparing the thermodynamic effort required to sustain flows from renewables depletable sources. The levelized exergy cost of (LExCOE) illustrates this disparity, showing that non-renewable sources like and have historically lower costs but higher fuel-related exergy demands, with LExCOE values around 3-4 MJ/MJ in the early dropping to 1-2.6 MJ/MJ by 2050 due to gains. In contrast, renewables such as and exhibit rising exergy costs initially (e.g., 143 TJ/MW for manufacturing in 2010) but project to undercut non-renewables by 2050, with LExCOE falling to 0.226 MJ/MJ for and 0.067 MJ/MJ for , driven by material and technological maturation. Exergy costs further quantify by estimating the equivalent primary exergy needed to replenish depleted resources, revealing that substituting fossil fuels with or equivalents demands 2-5 times more cumulative exergy upfront but avoids long-term depletion. These metrics promote , as renewables achieve lower overall exergy footprints when lifecycle emissions and remediation are included. In ecological applications, exergy analysis traces flows through to evaluate organizational and , positioning as a for exergy maximization. Eco-exergy, an extension of thermodynamic exergy, incorporates the informational content of organisms, calculated as the sum of contributions weighted by their genetic and structural distance from states. maximize exergy by optimizing capture and cycling, such as through diversified food webs that prolong throughflow and minimize dissipative losses, leading to higher eco-exergy in mature, biodiverse systems like coral reefs or old-growth forests compared to stressed or early-successional ones. enhances this maximization by enabling adaptive structures that increase network and buffer against perturbations, with studies showing eco-exergy indices rising 20-50% in recovering wetlands due to gains. This principle, rooted in the work of S.E. Jørgensen, underscores exergy as a function for development, where flows from solar radiation (the primary exergy input) are directed toward building ordered rather than . Recent advances since 2000 have integrated exergy-based indicators into global sustainability frameworks, particularly aligning with the (SDGs). Extended Exergy Accounting (EEA) extends traditional exergy analysis by incorporating labor, capital, and costs into a unified thermodynamic , applied in case studies like Romania's system to monitor progress on SDGs 7 (clean ), 12 (), and 13 (). Key indicators include the Exergy Footprint, which measures load in primary exergy units (e.g., kWh equivalents), and thermo-ecological cost, which penalizes non-renewable use to favor circular economies. These tools have informed SDG reporting by quantifying , with EEA revealing that transitioning to renewables could reduce a nation's total exergy demand by 30-40% while enhancing equity in energy access. Such indicators bridge and policy, prioritizing high-impact strategies like to minimize cumulative exergy extraction.

Environmental Policy and Economic Valuation

Exergy analysis has been proposed as a tool for informing environmental policies by quantifying the inefficiencies in energy conversion processes that lead to emissions. One seminal approach is the carbon exergy tax (CET), a thermo-economic mechanism that imposes a tax on CO₂ emissions proportional to the exergy destroyed and rejected in energy systems, thereby incentivizing higher efficiency and lower pollution. Developed to promote the sustainable use of exergy resources, CET calculates costs based on thermodynamic productivity deficits and residual exergy losses, outperforming traditional carbon taxes by directly linking penalties to inefficiency rather than just emission volume. For instance, in comparative studies of coal-fired plants and advanced fuel cell systems, CET encourages shifts toward technologies with exergy efficiencies exceeding 50%, reducing both fuel consumption and environmental impact. In broader frameworks, exergy serves as a metric for evaluating the quality of flows, enabling policymakers to address more precisely than energy-based measures alone. By identifying locations and magnitudes of exergy destruction—often the root of wasteful emissions—exergy guides regulations aimed at , such as those promoting renewable integration and standards in the . Although EU directives primarily focus on targets, exergy-based assessments have been advocated to refine these policies, highlighting opportunities for reducing irreversibilities in sectors like power generation and . This approach supports long-term environmental goals by quantifying the thermodynamic costs of and fostering economically viable transitions to low-exergy-loss systems. Exergy provides a unified for economic valuation in , capturing the and of resources by measuring their work potential relative to the . Unlike mere content, which ignores quality differences, exergy reflects the thermodynamic value of fuels and processes, allowing consistent across diverse forms—such as electrical versus —based on their capacity to drive economic activity. Studies in exergy demonstrate that this better correlates with value-added in production, as higher-quality exergy inputs yield greater economic output per unit, underscoring in non-renewable resources. For example, integrating exergy costs into models has shown potential for optimizing resource allocation while internalizing environmental externalities. The exergy content of manufactured goods offers a thermodynamic basis for assigning economic and resource , tracing the work potential in from materials. In , for instance, the exergy consumption via the blast -basic oxygen furnace route is approximately 22 per tonne of , encompassing reduction, transport, and refining stages where chemical and thermal exergy is converted into the material's structural . This quantification reveals the scarcity , as 's high exergy derives from irreversible processes that destroy significant input potential, informing that accounts for both market and thermodynamic costs. Such analyses promote practices by valuing recycled , which requires up to 70% less exergy than . Environmentally, exergy dissipation underscores the thermodynamic penalties of and , where high-quality energy is degraded into low-value waste forms. Emissions like CO₂ represent exergy destruction through entropy generation, as ordered dissipates into diffuse atmospheric , exacerbating by trapping low-exergy . In power plants, for example, over 60% of input exergy is lost as in condensers and exhausts, contributing to and indirectly to effects via inefficient . Exergy analysis thus aids climate mitigation by pinpointing these losses, advocating for technologies that minimize dissipation and recover residual potential, ultimately reducing the environmental footprint of human activities.

Advanced and Interdisciplinary Uses

Exergy in Life Cycle Assessment

Exergy-based (LCA) integrates thermodynamic principles to quantify the total useful energy extracted from natural resources across a product's entire , from acquisition to end-of-life disposal or . The primary indicator in this framework is the cumulative exergy demand (CExD), which aggregates the exergy content of all inputs required to deliver a product or service, thereby highlighting in terms of work potential rather than mere quantity. This approach, operationalized for LCA databases like ecoinvent, categorizes resources into fossil fuels, , , , other renewables, water, minerals, and metals, using specific exergy factors to compute the overall demand. Bösch et al. formalized CExD's application in LCA in 2007, extending earlier concepts of cumulative exergy consumption developed by Szargut and Morris in 1987. CExD is calculated as the sum of exergy contributions from raw materials and production processes, with credits applied for recoverable exergy in recyclables or by-products at the end of the . The can be expressed as: \text{CExD} = \sum \left( \text{exergy of raw materials} + \text{exergy of processes} \right) - \text{exergy of recyclables} where exergy values are derived from standard values for each resource type, ensuring a consistent thermodynamic basis. This accounts for direct and indirect exergy flows, making it suitable for process chains in LCA software. In contrast to traditional LCA methods, which often employ indicators like cumulative (CED) that treat all forms equally based on calorific value, exergy-based LCA emphasizes quality and efficiencies. For instance, CED might undervalue the superior work potential of (exergy equal to 100% of its content) compared to low-grade (exergy around 5-20%, depending on temperature), leading to incomplete assessments of impacts where high-grade is degraded into low-grade . thus provides a more precise tool for identifying irreversibilities and inefficiencies, as demonstrated in comparative studies fusing exergy with conventional LCA frameworks. Applications of exergy-based LCA span sectors like buildings and vehicles, with growing emphasis in the 2020s on strategies. In building assessments, CExD has revealed that material choices and operational energy sources dominate exergy demands, with case studies of residential and commercial structures showing up to 20-30% potential reductions through optimized designs and renewable integrations. For vehicles, exergy LCA evaluates battery production and drivetrain efficiencies, as seen in analyses of components where credits significantly lower net CExD. Recent 2020s studies have further applied CExD to contexts, such as aluminum recycling in automotive parts and components in cars, demonstrating exergy savings of 50-70% through loops compared to linear production.

Complex Systems and Cosmological Perspectives

In complex systems far from , exergy serves as a driving force for , enabling the emergence of ordered structures through the maximization of exergy flows and storage. Ilya Prigogine's theory of dissipative structures posits that open systems, by dissipating energy and matter, can maintain and evolve intricate patterns that counteract local increases, with exergy providing the available work potential for such processes. For instance, in a fluid layer heated from below, the transition to Bénard cells—a classic dissipative structure—results in an increase in system exergy due to enhanced efficiency, while decreases locally as ordered patterns form, though overall exergy destruction occurs via irreversible heat conduction. This maximization of exergy flows aligns with Prigogine's framework, where fluctuations amplify to produce stable, self-sustaining structures in non-equilibrium conditions. The implications of exergy dynamics extend to evolutionary processes in biological and technological domains, manifesting as hierarchies of increasing complexity. In biological , self-organizing systems like organisms tend toward higher rate densities, as evidenced by phylogenetic trends where rate density per unit rises from prokaryotes (around 10^4 erg s^{-1} ^{-1}) to more complex eukaryotes and multicellular , reflecting adaptations that optimize utilization for and . Similarly, technological development can be viewed as constructing exergy hierarchies, where human societies leverage flows to build layered systems—from simple tools to advanced —achieving rate densities up to 5 × 10^5 erg/s/, far exceeding biological levels and driving through optimized dissipation. These hierarchies illustrate how exergy gradients foster progressive organization, with natural and artificial selection favoring configurations that enhance throughput per unit . From a cosmological viewpoint, the universe's total exergy diminishes over time in accordance with the second law of thermodynamics, as the expansion dilutes available energy gradients, leading toward a state of maximum entropy and thermal equilibrium. The Big Bang represents the primordial high-exergy condition, characterized by extreme density and low entropy (approximately 10^88 k_B total), providing the initial potential for all subsequent structure formation through gravitational and nuclear processes that locally concentrate exergy. This cosmic evolution unfolds via increasing local exergy densities—from galaxies (∼0.5 erg/s/g) to stars, planets, life, and technology—against the backdrop of global exergy depletion, as nuclear fuels are consumed and radiation homogenizes. Philosophically, exergy quantifies the potential for in physical systems, serving as a metric of distance from and thus underpinning the , where irreversible processes dictate the unidirectional progression from ordered states to . In this , exergy's role highlights a tension between local complexity growth—driven by flows that enable and —and the inexorable global increase in , framing the universe's trajectory as one of transient hierarchies emerging within an overarching thermodynamic decline. This perspective integrates physical laws with the of structure, suggesting that the arises from the finite availability of exergy, constraining the scope and duration of ordered phenomena across scales.

References

  1. [1]
    Exergy - an overview | ScienceDirect Topics
    Exergy is defined as the amount of work (= entropy-free energy) a system can perform when it is brought into thermodynamic equilibrium with its environment.
  2. [2]
    [PDF] An introduction to exergy and its evaluation using Aspen Plus - K-REx
    concept which is derived from the second law of thermodynamics. Exergy has been defined as. “the ability to do work” or “the amount of work that can be ...
  3. [3]
    [PDF] A brief Commented History of Exergy From the Beginnings to 2004
    As stated above, the modern definition of exergy is a rephrasing of Gibbs' original statement: The exergy of a thermodynamic system S in a certain state SA is ...
  4. [4]
    Energy and exergy analyses of a VVER type nuclear power plant
    The term “Exergy” was initially used by Rant [1] in 1956. At that time it ... Zoran Rant. Exergie, Ein neues Wort für “technische Arbeitsfähigkeit ...Missing: coinage | Show results with:coinage
  5. [5]
    [PDF] Exergy Analysis of Rocket Systems
    Szargut defines exergy as “the amount of work obtainable when some matter is brought to a state of thermodynamic equilibrium with the common components of the ...
  6. [6]
    [PDF] Introduction to the Concept of Exergy - - IEA EBC
    Apr 25, 2002 · Chapter 1 describes the characteristics of a thermodynamic concept, exergy, in association with building heating and cooling systems.
  7. [7]
    [PDF] 67.pdf - Stanford School of Earth, Energy & Environmental Sciences |
    Jul 13, 2021 · Exergy, or availability, is a thermodynamic concept representing the useful work that can be extracted from.
  8. [8]
    16.1 Availability (Exergy) Analysis
    A system is said to be in the dead state when it is in thermodynamic equilibrium with its surroundings. At the dead state: 1. A system is at the same ...
  9. [9]
    Exergy Efficiency of Closed and Unsteady-Flow Systems - PMC
    Sep 10, 2025 · 4. Definition of Exergy Efficiency. Exergy efficiency or second-law efficiency is a measure of the actual performance of an energy system ...<|control11|><|separator|>
  10. [10]
    Physical Exergy - an overview | ScienceDirect Topics
    The thermodynamics or physical exergy is defined as the maximum work obtained in reversible processes when a quantity of material is taken from the initial ...
  11. [11]
    An Introduction to Chemical Thermodynamics for Engineers
    Two components of exergy are considered: physical exergy, which is related to changes in temperature, pressure and concentration; and chemical exergy, which is ...<|separator|>
  12. [12]
    Energy quality and energy grade: concepts, applications and ...
    Mar 22, 2022 · The concept of exergy was coined by Rant [3] in 1956. Available energy (or availability) represents the same physical meaning as exergy. Exergy ...
  13. [13]
    [PDF] Think ExErgy, noT EnErgy - Science Europe
    ... exergy approach can identify and quantify the causes of inefficiencies. Exergy is therefore the right metric to value energy use and resource scarcity. in ...<|control11|><|separator|>
  14. [14]
    Exergy Destruction - an overview | ScienceDirect Topics
    In accordance with the second law, the exergy destruction is positive in an irreversible process and vanishes in a reversible process. The change in exergy of a ...
  15. [15]
    [PDF] Energy, exergy, and Second Law performance criteria
    thermodynamic definition is to define the numerator terms as having a monetary benefit value, and those in the denominator as having a monetary cost, to the ...
  16. [16]
    Exergy as a Measure of Resource Use in Life Cycle Assessment ...
    Jun 29, 2016 · The energetic argument claims that useful energy (i.e., exergy) is the ultimate limiting and scarce resource because every material resource ...
  17. [17]
    Understanding energy and exergy efficiencies for improved energy ...
    Although exergy analysis can be generally applied to energy and other systems, it appears to be a more powerful tool than energy analysis for power cycles ...
  18. [18]
    None
    ### Summary of Exergy Concepts with Illustrative Examples
  19. [19]
    Exergy destruction during the combustion process as functions of ...
    Aug 7, 2025 · Approximately 20% of the chemical exergy of fuel is destroyed during the combustion process [7, 11,12], and studies have been conducted to ...
  20. [20]
    Exergy Destruction - an overview | ScienceDirect Topics
    Chemical exergy caused by the combustion procedure was the principal source of exergy destruction (irreversibility) for System 4. In addition, heat losses in ...
  21. [21]
    [PDF] EXERGY OF SOLAR RADIATION
    Exergy of solar radiation is a measure of its working capacity, estimated by energy and exergy, which is the ability to work in the human environment.
  22. [22]
    How much exergy one can obtain from incident solar radiation?
    Feb 27, 2009 · A thermodynamic model is proposed to study the exergetic content of incident solar radiation reaching on the Earth's surface which can be ...
  23. [23]
    A History of Thermodynamics: The Missing Manual - PMC
    We present a history of thermodynamics. Part 1 discusses definitions, a pre-history of heat and temperature, and steam engine efficiency, which motivated ...
  24. [24]
    [PDF] REFLECTIONS ON THE MOTIVE POWER OF FIRE AND ON ... - ASME
    In 1824, Carnot published Reflections on the. Motive Power of Fire, which detailed his research and presented a well-reasoned theoretical treatment for the ...
  25. [25]
    Gibbs (Free) Energy - Chemistry LibreTexts
    Jan 29, 2023 · Gibbs energy was developed in the 1870's by Josiah Willard Gibbs. He ... The definition of Gibbs energy can then be used directly.Introduction · Gibbs Energy in Reactions · Gibbs Energy in Equilibria
  26. [26]
    Hermann von Helmholtz - Stanford Encyclopedia of Philosophy
    Feb 18, 2008 · Helmholtz proposed the notion of a “free energy” to account for cases involving heat and entropy. The Helmholtz free energy is defined as F ...
  27. [27]
    A brief Commented History of Exergy From the Beginnings to 2004
    ArticlePDF Available. A brief Commented History of Exergy From the Beginnings to 2004. March 2007; International Journal of Thermodynamics 10(1). DOI:10.5541 ...
  28. [28]
    [PDF] EXERGY AS A TOOL FOR DESIGNING AND OPERATING ...
    In the 1950s and 1960s, contributions to the exergy concept were also made by Z. Rant, P. Grassmann, W. Μ. Brodyansky, Ε. A. Bruges, Μ. Tribus,. Ε. F. Obert ...
  29. [29]
    None
    ### Summary of Exergy from Lecture 5 Slides (MIT 2.813, T. G. Gutowski, Feb 24, 2006)
  30. [30]
  31. [31]
    [PDF] EXERGY
    Equation (3.2) may be read as follows: the useful work exchanged by a system is the total work exchanged minus the work exchanged with the atmosphere, assumed ...Missing: formula | Show results with:formula
  32. [32]
    [PDF] EXERGY
    Electrical and mechanical energies are very “high quality” forms of energy: their exergy index is 100% since exergy is equal to energy. ... Moreover, this graph ...
  33. [33]
    [PDF] Exergy Analysis of Engineering Applications - Dr. Md. Zahurul Haq
    Exergy Analysis. Equations: CM & CV Systems. CM System: • Q − W = ∆U. • Wu = W − P0∆V = W − W0. • φ = (u − u0) + P0(v − v0) − T0(s − s0). • ΦQ = Pn j=1 Qj 1 − ...
  34. [34]
    Standard chemical exergy of some elements and compounds on the ...
    Values of the standard chemical exergy of 49 elements, and some inorganic and organic compounds of those elements, are proposed.
  35. [35]
    [PDF] 1 APPENDIX 1. STANDARD CHEMICAL EXERGY (Tn= 298.15 K ...
    devaluation. Standard chemical exergy. M, kg/kmol. D o , kJ/mol e o x,ch, kJ/mol. NiO s. 74.71. 0. 23.0. Ni(OH)2 s. 92.72. -48.13. 25.5. NiS s. 909.77. 883.15.Missing: formula sum μ_i
  36. [36]
    Standard chemical exergy of elements updated - ScienceDirect
    According to Szargut [3], [7], to calculate chemical exergy of a substance, reference species should be chosen as the most probable products of the interactions ...
  37. [37]
    Calculating the chemical exergy of materials - Wiley Online Library
    Mar 31, 2021 · For a known and homogenous mixture of species, the standard molar chemical exergy can be computed using the formula by Szargut (Szargut, 2005) ...
  38. [38]
  39. [39]
    [PDF] EXERGY OF SOLAR RADIATION
    The derived exergy formulae take into account that the solid angle in which the radiation propagates and its energy spectrum are determined by measurements. The ...
  40. [40]
    [PDF] Exergy of direct solar radiation
    Apr 30, 2018 · the flux of solar radiation is characterized by its own temperature (radiation), and its exergy ... The brightness temperature of the thermal ...
  41. [41]
    exergy of radiation, effective temperature of photon and entropy ...
    Therefore, if the photosynthesis is discussed by the spectral exergy efficiency, the theoretical results will be con- sistent with the experimental results, and ...
  42. [42]
    Exergy Dynamics of Systems in Thermal or Concentration Non ...
    Jun 8, 2017 · The non-equilibrium exergy is always larger than its equilibrium counterpart and constitutes the “real” total exergy content of the system, i.e. ...
  43. [43]
    Bioengineering thermodynamics of biological cells - PMC
    Dec 1, 2015 · Keywords: Entropy generation, Exergy ... thermodynamics applied to biosystems, biological cells, cancer and non equilibrium states.
  44. [44]
    Exergy transfer analysis of microwave heating systems - ScienceDirect
    Relation between microwave exergy and Poynting vector is analyzed. •. A general formulation of exergy transfer analysis for microwave heating systems is ...
  45. [45]
    [PDF] E xergy of undiluted thermal radiation - Sci-Hub
    Utilization of solar radiation, by an absorbing surface, is evaluated and the optimum temperature of this surface is determined. The above considerations were ...<|control11|><|separator|>
  46. [46]
    Energy, Exergy, Entropy Generation Minimization, and ... - Frontiers
    This brief review presents an overview of the essential elements of the methods of energy, exergy, entropy generation minimization, and exergoenvironmental ...
  47. [47]
  48. [48]
    [PDF] Exergy Analysis of a Vapor Absorption Refrigeration System Using ...
    3.4 Second Law of Thermodynamic (Exergy) Analysis. The analysis based on the first law of thermodynamics determines the amount of energy entering and leaving ...
  49. [49]
    Kinetic Exergy - an overview | ScienceDirect Topics
    ... exergy has four major parts: physical, chemical, potential and kinetic. Kinetic and potential exergy have the same meaning as the corresponding energy terms.
  50. [50]
    Exergy from Thermodynamics | McGraw-Hill Education
    Energy only covers quantity, while exergy covers both the quantity and quality of energy. This means that energy is related to the first law of thermodynamics.
  51. [51]
    Potential Exergy - an overview | ScienceDirect Topics
    Potential exergy is defined as the work obtainable from a system's state, such as its position or chemical composition, due to a disparity with its environment ...<|separator|>
  52. [52]
    Chemical Exergy - an overview | ScienceDirect Topics
    Example 3.1: Calculate the molar chemical exergy of H2O and O2 using the data on partial pressures in Table 3.1, for a reference state of T0 = 298.15 K and ...
  53. [53]
    The Tenuous Use of Exergy as a Measure of Resource Value or ...
    Until the limitations are addressed, exergy should only be used for its original purpose as a decision making tool for engineering systems analysis. Keywords:.
  54. [54]
    Second law analysis of a conventional steam power plant
    Nov 1, 1993 · The boiler feed pump turbine accounts for 0.25 percent of the total exergy destruction. Fluid flow mixing is responsible for 0.23 percent of ...
  55. [55]
    Energy and exergy analysis of the steam power plants
    From the result, it was reported that the combustion chamber in the boiler exhibited the highest exergy destruction (77%) with energy efficiency of 43.8%.
  56. [56]
    Energy and exergy analyses for a combined cycle power plant in ...
    The exergy destruction to the total exergy destruction percentage ratio was evaluated to be as high as 77%, and 13% in the boiler system and turbine, ...
  57. [57]
    Exergy analysis in combined heat and power systems: A review
    Dec 15, 2020 · In this study, a comprehensive literature review is conducted based on the published papers on exergy analysis and exergoeconomic evaluation of CHP systems.Missing: matching | Show results with:matching
  58. [58]
    (PDF) Optimal design of cogeneration system based on ...
    Aug 6, 2025 · In this article, a new procedure was introduced for the optimal design of utility system in process industries. This method was based on the ...
  59. [59]
    Innovative software for thermodynamic cycle processes
    EBSILON®Professional works with conventional power plants, nuclear and solar power plants, desalination plants, fuel-cell applications, and user-specific ...
  60. [60]
    ProSimPlus Energy software - chemical plant energy efficiency
    ProSimPlus Energy is a must-have software to improve process energy efficiency. Reducing operating costs and optimizing capital investment.
  61. [61]
    Exergy analysis in ProSimPlus® simulation software
    ProSimPlus® is a process engineering software that performs rigorous steady-state mass and energy balances for a wide range of industrial processing plants ( ...Exergy Analysis In... · Exergy Calculations · Exergy Analysis In A Process...
  62. [62]
    Energy, Exergy, and Emissions Analyses of Internal Combustion ...
    Aug 31, 2023 · This work presents longitudinal dynamics simulations of passenger vehicles to understand the magnitude of exergy destruction in ICEVs and BEVs.
  63. [63]
    An exergy analysis methodology for internal combustion engines ...
    Dec 15, 2019 · When combustion started, physical exergy increased slightly while chemical exergy reduced significantly due to exergy transfers associated with ...Missing: automotive | Show results with:automotive
  64. [64]
    Non-renewable and renewable levelized exergy cost of electricity ...
    Dec 30, 2024 · Table 3 shows that renewable energies (hydro, wind and photovoltaic) exhibit higher exergy costs for manufacturing (143–6.9 TJ/MW), compared to ...
  65. [65]
    Exergy Replacement Cost of Fossil Fuels: Closing the Carbon Cycle
    Cumulative Exergy Consumption (CExC) is a method used to account for the totality of exergy cost through a chain of production processes or the life cycle of a ...Missing: coefficient CEEx
  66. [66]
    Exergy as a Tool for Ecosystem Health Assessment - MDPI
    This paper reviews the application of exergy in ecology in the fields of ecological modeling and natural ecosystem monitoring.
  67. [67]
  68. [68]
    Exergy-based ecological indicators: From Thermo-Economics to ...
    Feb 1, 2019 · This paper presents a summary of the conceptual development and the practical applications of exergy-based Environmental Indicators. After a ...
  69. [69]
    Multi-scale extended exergy analysis of the “system Romania”
    Jan 1, 2025 · Multi-scale extended exergy analysis of the “system Romania”: A tool for monitoring the UN-2030 SDGs · Highlights · Abstract · Introduction.
  70. [70]
  71. [71]
    The role of exergy in energy policy making - ResearchGate
    Aug 5, 2025 · This paper deals with the utilization of exergy as an efficient tool for energy policy making applications since exergy is a measure of ...
  72. [72]
    Exergy Accounting: A Quantitative Comparison of Methods ... - MDPI
    First, exergy is a measure of the capacity of an energy flow to perform physical work and hence may better represent the economic value of energy flows than ...Missing: scarcity | Show results with:scarcity
  73. [73]
  74. [74]
  75. [75]
    Modelling entropy and exergy changes during a fluid self ...
    The self-organization process of a thermodynamic system has been investigated 'from within' by calculating system exergy and entropy changes due to ...
  76. [76]
    [PDF] prigogine-lecture.pdf - Nobel Prize
    Irreversible processes may lead to a new type of dynamic states of matter which I have called “dissipative structures”. Sections 2-4 are devoted to the ...Missing: exergy | Show results with:exergy
  77. [77]
    [PDF] Energy rate density as a complexity metric and evolutionary driver
    May 17, 2010 · Chaisson, E.J. Cosmic Evolution: The Rise of Complexity in Nature; Harvard University Press: Cambridge, 2001. 4. Chaisson, E.J. Complexity: An ...
  78. [78]
    A model for the cosmic creation of nuclear exergy - Nature
    Apr 1, 1982 · An old theme in cosmic thermodynamics is that the second law implies that the Universe is running down (running out of exergy) and approaching ...