Fact-checked by Grok 2 weeks ago

Polytropic process

A polytropic process is a thermodynamic process involving the expansion or compression of a gas or vapor, characterized by the relation pV^n = C, where p is the pressure, V is the volume, n is the polytropic index (a constant determined by the specific process), and C is a constant. This relation describes a reversible path where both heat transfer and work occur, generalizing several ideal processes for real-world approximations in systems like engines and compressors. The polytropic index n typically ranges from 0 to \infty, with specific values corresponding to familiar thermodynamic processes: n = 0 for isobaric (constant pressure), n = 1 for isothermal (constant temperature), n = \gamma (where \gamma = C_p / C_v) for reversible adiabatic (no heat transfer), and n = \infty for isochoric (constant volume). For an ideal gas undergoing a polytropic process, the work done is given by W = \frac{p_1 V_1 - p_2 V_2}{n-1} (for n \neq 1), and the heat transfer can be derived from the first law of thermodynamics as Q = \Delta U + W, where \Delta U is the change in internal energy. These processes assume constant specific heat, allowing for analytical solutions that bridge ideal and non-ideal behaviors, though deviations occur with real gases. Polytropic processes are widely applied in engineering contexts, such as modeling in vapor cycles (where $1 < n < \gamma, leading to heat rejection) or expansion in internal combustion engines (where heat losses make n < \gamma). They provide a practical framework for calculating efficiency and performance in turbomachinery, with the index n often determined experimentally to account for irreversibilities like friction or heat dissipation. Unlike purely ideal processes, polytropes capture intermediate behaviors, making them essential for thermodynamic analysis in power generation and HVAC systems.

Definition and Fundamentals

Definition

A polytropic process is a generalized thermodynamic path followed by a fluid, in which the pressure and volume are related by a power-law relationship parameterized by the polytropic index, allowing it to describe a wide range of expansion or compression behaviors. This process applies to ideal gases, real gases, and vapors, providing a flexible model for systems where heat transfer and work occur simultaneously. Unlike specific processes such as isobaric or isochoric, the polytropic path captures intermediate behaviors through variation in the index value, making it suitable for approximating real-world fluid dynamics in engineering applications. The term "polytropic" was coined in the 19th century by German engineer , who introduced it around 1866 to model non-ideal thermodynamic changes, particularly in steam engines and compressors where actual processes deviate from idealized assumptions. Zeuner's work focused on incremental state changes in gases and vapors, laying the foundation for using polytropic paths to represent practical approximations of complex heat and work interactions in machinery. This historical development emphasized the process's role in bridging theoretical ideals with empirical observations in early thermodynamic engineering. Polytropic processes are typically analyzed under the assumptions of a closed system, where no mass crosses the boundary, and quasi-static changes occur slowly enough to maintain near-equilibrium conditions at every stage, without chemical reactions altering the system's composition. These prerequisites ensure that the process can be treated as a sequence of equilibrium states, facilitating the application of thermodynamic principles like the first law. The utility of the polytropic model lies in its ability to interpolate between distinct processes—such as those with constant temperature or entropy—by adjusting the index, thus providing a versatile tool for analyzing efficiency and performance in devices like reciprocating engines and turbines.

Governing Equation

The polytropic process is mathematically characterized by the governing equation PV^n = C, where P denotes pressure, V denotes volume, n is the polytropic index, and C is a constant that remains invariant throughout the process. This equation encapsulates a wide range of thermodynamic behaviors by varying the value of n, providing a unified framework for processes involving heat and work transfer in gases or vapors. The equation is derived for an ideal gas from the first law of thermodynamics under the assumption of a constant effective specific heat. For N moles of gas, the heat transfer is \delta Q = C \, dT, where C is the total effective heat capacity. The first law gives dU = \delta Q - P \, dV, with dU = N C_V \, dT. Substituting yields N C_V \, dT = C \, dT - P \, dV, or P \, dV = (C - N C_V) \, dT. Using the ideal gas law PV = N R T, differentiate to P \, dV + V \, dP = N R \, dT, and solve for dT to substitute back, leading after algebraic manipulation and integration to P V^n = C, where the polytropic index n = 1 + \frac{N R}{C - N C_V}. This derivation assumes ideal gas behavior and quasi-equilibrium. For ideal gases, the governing equation extends to other thermodynamic variables, such as temperature and volume. Combining PV^n = C with the ideal gas law PV = N R T leads to T V^{n-1} = \constant, illustrating how temperature varies inversely with volume raised to the power n-1. Similarly, relations like T P^{(1-n)/n} = \constant can be derived for pressure-temperature paths. The polytropic index n is dimensionless, ensuring dimensional homogeneity in the equation PV^n = C, as the exponents on pressure and volume (both with dimensions of energy per mole or similar) balance without introducing units to n. This property allows n to be determined empirically from process measurements, maintaining consistency across different systems.

Polytropic Index

The polytropic index, denoted n, is a dimensionless parameter that characterizes the pressure-volume relationship in a polytropic process, influencing the balance between heat capacity effects and the degree of heat transfer relative to work done during compression or expansion. It determines how the process deviates from ideal cases like adiabatic or isothermal behavior, with higher values of n indicating less heat exchange and steeper pressure-volume curves on logarithmic diagrams. Specific values of n correspond to well-known thermodynamic processes: n = 0 for an at constant pressure, n = 1 for an at constant temperature, n = \gamma (where \gamma is the specific heat ratio C_p / C_v) for an with no heat transfer, and n \to \infty for an at constant volume. These assignments highlight n's role in parameterizing a continuum of processes between constant-pressure and constant-volume limits. The range of n is typically $0 < n < \infty for standard gas compression and expansion processes, reflecting varying levels of heat involvement; values between 1 and \gamma (e.g., 1.4 for air) are common in engineering applications like compressors, where partial heat loss occurs. In rare expansions involving significant cooling or unusual heat addition, n can be negative, leading to pressure increases with volume. To determine n, experimental methods involve fitting pressure-volume data from measured process paths or calculating from inlet and outlet temperatures and pressures using the relation n = \frac{\ln(P_2 / P_1)}{\ln(V_1 / V_2)}. Theoretical models based on heat transfer rates, such as those accounting for finite-rate heat exchange in dynamic systems, yield n iteratively by incorporating factors like thermal conductivity and surface area. The polytropic index relates to the specific heats through the effective polytropic specific heat C_n = C_v \frac{\gamma - n}{1 - n}, where C_v is the specific heat at constant volume and \gamma = C_p / C_v; this can be rearranged as n = \frac{\gamma - m}{1 - m}, with m = C_n / C_v representing a normalized measure of heat capacity influenced by heat loss or gain during the process. This connection underscores n's parameterization of non-adiabatic effects without implying reversibility.

Thermodynamic Properties

Work Done

In a polytropic process for a closed system, the boundary work is calculated as the integral of pressure with respect to volume, W = \int_{V_1}^{V_2} P \, dV, where the sign convention defines positive work as that done by the system during expansion and negative during compression. For a process following P V^n = C with n \neq 1, substituting P = C V^{-n} into the integral yields W = \frac{P_2 V_2 - P_1 V_1}{1 - n}, which simplifies equivalently to W = \frac{P_1 V_1 - P_2 V_2}{n - 1} using algebraic rearrangement. This formula applies to both ideal and non-ideal gases, provided the polytropic relation holds, and assumes quasistatic conditions for the boundary work evaluation. For the special case where n = 1, corresponding to an for an ideal gas, the relation becomes P V = C, and the work integral evaluates to W = P_1 V_1 \ln \left( \frac{V_2}{V_1} \right). This logarithmic form arises directly from integrating P = \frac{C}{V}, resulting in positive work for expansion (V_2 > V_1) and negative for compression. Graphically, the work done corresponds to the area beneath the process curve on a pressure-volume (P-V) diagram, where the polytropic path is steeper for larger n values. For expansion from a fixed initial state, increasing n (e.g., from 1 to \gamma \approx 1.4 for air) reduces the area under the curve, thereby decreasing the magnitude of work output compared to shallower paths like isobaric (n = 0). The work in a polytropic process relates to heat transfer through the first law of thermodynamics.

Heat Transfer

In a polytropic process for an , the Q is calculated using of , Q = \Delta U + W, where \Delta U is the change in and W is the done by the system. For an undergoing such a process, \Delta U = m c_v (T_2 - T_1), with m as the , c_v as the specific heat at constant volume, and T_2 - T_1 as the temperature change; the work W is given by W = \frac{m R (T_2 - T_1)}{1 - n}, where R is the specific and n is the polytropic index. Substituting these into yields the explicit expression for : Q = m c_v \frac{\gamma - n}{1 - n} (T_2 - T_1), where \gamma = c_p / c_v is the specific heat ratio. This formula equivalently appears as Q = m c_v \frac{n - \gamma}{n - 1} (T_2 - T_1), highlighting the dependence on n relative to \gamma. The can also be related directly to the work done: Q = \frac{\gamma - n}{\gamma - 1} W. This relation stems from and the expressions for \Delta U and W, emphasizing how the polytropic index modulates the exchange relative to mechanical work. Additionally, the process involves an effective c_n = c_v \frac{\gamma - n}{1 - n}, such that Q = m c_n (T_2 - T_1); this c_n can be positive, negative, or zero depending on n, reflecting the non-standard heat-temperature relationship in polytropic paths. The direction of heat flow is determined by the sign of Q, which depends on n compared to \gamma and the process direction (e.g., expansion or compression). For n = \gamma, Q = 0, indicating no as in an . When n < \gamma, is typically added to the system (positive Q) during , as the process deviates from adiabatic conditions toward requiring thermal input to follow the PV^n = constant path. Heat transfer Q is measured in SI units of joules (J) or, in imperial units, British thermal units (Btu). For a numerical illustration, consider 1 kg of air (\gamma = 1.4, c_v = 0.717 kJ/kg·K) undergoing expansion from T_1 = 300 K to T_2 = 250 K with n = 1.2. Here, \Delta T = -50 K, so Q = 1 \times 0.717 \times \frac{1.2 - 1.4}{1.2 - 1} \times (-50) = 0.717 \times (-1) \times (-50) = 35.85~\text{kJ}, indicating heat addition to the system during the expansion.

Internal Energy Change

For an ideal gas undergoing a polytropic process, the change in internal energy depends solely on the temperature change between the initial and final states, as the internal energy is a function of temperature only. The formula for the internal energy change is given by \Delta U = m C_v (T_2 - T_1), where m is the mass of the gas, C_v is the specific heat at constant volume, and T_1 and T_2 are the initial and final temperatures, respectively. The temperatures in a polytropic process are related to the volumes through the equation derived from the and the polytropic relation PV^n = \constant: \frac{T_2}{T_1} = \left( \frac{V_1}{V_2} \right)^{n-1}, where V_1 and V_2 are the initial and final volumes, and n is the polytropic index. This relation shows that the temperature variation—and thus \Delta U—is indirectly influenced by n through the process path, even though internal energy itself does not depend directly on n. The magnitude of \Delta U varies with n; for processes closer to isothermal (n = 1), where T_2 = T_1 and \Delta U = 0, the change is minimal, while it increases for values of n farther from 1, such as in (n = \gamma), leading to larger temperature differences. For real gases, the internal energy change includes additional contributions from intermolecular forces and volume effects, typically accounted for using departure functions that quantify deviations from .

Specific Cases

Isobaric Process

In the context of polytropic processes, the isobaric process occurs when the polytropic index n = 0, maintaining constant pressure throughout the thermodynamic change. This condition implies P = \text{constant}, distinguishing it from other polytropic cases where pressure varies with volume according to PV^n = \text{constant}. For an ideal gas, the constant pressure aligns with the ideal gas law PV = mRT, leading to a direct proportionality between volume and temperature: V/T = \text{constant}. This relationship holds as long as the gas behaves ideally, allowing temperature changes to drive proportional volume expansions or contractions without altering pressure. The work done by the system during an isobaric process is calculated as the product of the constant and the change in . Mathematically, this is expressed as W = P (V_2 - V_1), where P is the constant pressure, and V_1 and V_2 are the initial and final s, respectively. This formula arises from the integral W = \int_{V_1}^{V_2} P \, dV, simplifying under constant P. For expansion (V_2 > V_1), the work is positive (done by the system), representing the maximum possible for a given volume change among polytropic processes with the same initial , as pressure remains at its highest value throughout. Heat transfer in an for an is governed by of , Q = \Delta U + W, where the change in \Delta U = m C_v (T_2 - T_1) depends solely on temperature change. Substituting the isobaric work yields Q = m C_p (T_2 - T_1), with C_p as the at constant pressure (C_p = C_v + [R](/page/R), where R is the ). This heat input is the maximum required for a given \Delta T among polytropic processes, as the maximized work contribution amplifies the total energy needed beyond the fixed internal energy change. Isobaric processes are frequently applied in heating scenarios at constant pressure, such as in piston-cylinder devices where external pressure is balanced by atmospheric or reservoir conditions, facilitating controlled .

Isothermal Process

An represents a special case of the polytropic process with polytropic index n = 1, where the T remains constant throughout. For an , this condition implies the relation PV = constant, derived from the PV = mRT under constant T. This process occurs when the gas undergoes or while exchanging with its surroundings to precisely maintain the temperature, ensuring no net change in the system's thermal state. The work done by the during an isothermal expansion from initial volume V_1 to final volume V_2 is given by W = m R T \ln \left( \frac{V_2}{V_1} \right), where m is the mass of the gas and R is the specific . For an , the change \Delta U = 0 in an , as internal energy depends solely on temperature. Consequently, from the first law of , the Q equals the work done W (positive for expansion, indicating heat absorption by the system). In a reversible , which requires quasi-static execution to maintain at every stage, the change of the is \Delta S = m R \ln \left( \frac{V_2}{V_1} \right). This positive value for expansion reflects increased disorder due to greater volume availability for molecular motion. Such processes are fundamental to the Carnot cycle's isothermal legs, supporting the cycle's reversible operation and theoretical efficiency limit.

Adiabatic Process

The represents a specific instance of the polytropic process for an , characterized by the polytropic index n = \gamma, where \gamma = C_p / C_v is the ratio of the specific heat at constant to the specific heat at constant volume. In this process, no is transferred to or from the ([Q](/page/Q) = 0), leading to the governing relation PV^\gamma = \text{[constant](/page/Constant)}. This equation arises from the combination of of and the assumptions under adiabatic conditions. A key derived relation for the connects and as TV^{\gamma-1} = \text{[constant](/page/Constant)}. This follows from substituting the into the pressure- relation and integrating, highlighting how varies inversely with raised to the power \gamma - 1, with steeper changes for diatomic gases where \gamma \approx 1.4 compared to monatomic gases where \gamma = 5/3. Since Q = 0, simplifies to \Delta U = W, where W is the work done on the system. For an , the change in is \Delta U = m C_v (T_2 - T_1), yielding W = m C_v (T_1 - T_2) for expansion from state 1 to state 2 (where T_2 < T_1) when considering work done by the system as positive. These expressions hold precisely for reversible adiabatic processes, which are quasi-static and isentropic. Irreversible adiabatic processes, such as rapid expansions involving friction or non-equilibrium effects, deviate from the ideal n = \gamma behavior but can be approximated using a polytropic model with an effective index n \neq \gamma, fitted from experimental pressure-volume data to capture entropy generation and inefficiencies.

Isochoric Process

The isochoric process represents a limiting case of the polytropic process as the polytropic index n approaches infinity, where the volume remains constant throughout. In this scenario, the process adheres to the polytropic relation pV^n = \text{constant}, which simplifies to constant volume as n \to \infty. For an ideal gas undergoing an isochoric process, the ideal gas law implies that the ratio of pressure to temperature P/T remains constant, since PV = mRT with fixed V and mass m. Thus, any change in temperature directly scales the pressure, expressed as P_2 / P_1 = T_2 / T_1, where subscripts 1 and 2 denote initial and final states; this relation is essential for modeling constant-volume heating or cooling scenarios in ideal gases. Due to the absence of volume change, no boundary work is performed, so the work done W = 0. By the first law of thermodynamics, the heat transfer Q then equals the change in internal energy \Delta U, yielding Q = \Delta U = m c_v (T_2 - T_1), with c_v as the specific heat capacity at constant volume. This highlights that all energy input via heat directly alters the internal energy without mechanical work output.

Relationships and Comparisons

Equivalence to Energy Transfer Ratios

In a polytropic process for an ideal gas, the index n can be interpreted physically as arising from a constant ratio of infinitesimal heat transfer to infinitesimal work, \frac{\delta Q}{\delta W} = K, where K is a constant specific to the process. This energy transfer ratio distinguishes the polytropic process from other thermodynamic paths and directly links to the pressure-volume relation PV^n = \text{constant}. For an ideal gas, the first law of thermodynamics, \delta Q = dU + \delta W (with \delta W = P dV as work done by the system), implies K = 1 + \frac{dU}{\delta W}. Integrating over the process yields the overall ratio K = \frac{Q}{W} = 1 + \frac{\Delta U}{W}, where \Delta U is the change in internal energy and W is the net work done by the system. Deriving the explicit form of n from this ratio involves the ideal gas properties and specific heats. The boundary work for a polytropic process is W = \frac{R (T_1 - T_2)}{n - 1}, where R is the gas constant and T_1 > T_2 for expansion, while \Delta U = c_v (T_2 - T_1) = -c_v (T_1 - T_2) with c_v the specific heat at constant volume. Substituting into the first law gives Q = \Delta U + W, leading to \frac{Q}{W} = 1 - \frac{(n - 1) c_v}{R}. Since \frac{R}{c_v} = \gamma - 1 (where \gamma = \frac{c_p}{c_v} is the heat capacity ratio), rearranging yields n = 1 - (\gamma - 1) \frac{\Delta U}{W}. This equation demonstrates that n equates to a specific ratio of non-work energies (internal energy change relative to work), encapsulating how the process partitions energy between storage in the gas and mechanical output. Physically, the index n thus reflects the balance between compression or expansion work and heat dissipation or addition during the process. A higher n (closer to \gamma) indicates less relative to work, approaching adiabatic conditions where energy changes are dominated by internal energy adjustments without external heat exchange. For instance, when heat transfer is zero (Q = 0), \Delta U = -W and n = \gamma. This energy ratio perspective provides insight beyond the geometric P-V description, highlighting polytropic processes as those maintaining proportional energy exchanges throughout.

Comparison with Reversible Processes

Polytropic processes serve as approximations to ideal reversible processes in thermodynamic analyses, particularly for gases undergoing or with . In the limit of n = 1, a polytropic process for an becomes a reversible , where temperature remains constant due to slow, quasi-static that maintains with the surroundings. Conversely, when n = \gamma (the ratio of specific heats), the process aligns with a reversible adiabatic (, characterized by no and constant , as prevents any exchange with the environment. A key distinction arises in , where polytropic processes account for irreversibilities that reduce compared to their reversible counterparts. For compressors, the polytropic \eta_p, which measures the ratio of isentropic to actual work for stages, is given by \eta_p = \frac{\gamma - 1}{\gamma} \cdot \frac{n}{n - 1}. This value equals 1 for the reversible adiabatic case (n = \gamma) but falls below 1 for irreversible polytropic processes (n > \gamma), reflecting increased work input due to factors like and non-ideal . Irreversibilities in polytropic processes manifest as entropy generation, primarily from heat transfer occurring at finite rates across temperature gradients rather than infinitesimally. Unlike reversible processes, where entropy change is solely due to reversible heat transfer (\Delta S = \int \frac{\delta Q_{\text{rev}}}{T}), polytropic processes produce additional entropy through these gradients, quantifying the loss of available work. Modern validations using (CFD) confirm the accuracy of polytropic approximations in modeling real flows, such as in compressors and engines, by comparing simulated pressure-volume paths and against experimental data. These studies demonstrate that polytropic models effectively capture deviations from ideality in complex geometries, with errors typically under 5% for head and efficiency predictions.

Applications and Extensions

Engineering Applications

Polytropic processes are extensively applied in the and of compressors and turbines, particularly in multi-stage compression systems where real-gas behavior and necessitate a polytropic exponent n between 1 and \gamma (the adiabatic index). In centrifugal and axial compressors, polytropic modeling allows for the calculation of head, a key performance metric representing the imparted to the per unit , using the formula H = \frac{n}{n-1} R T \left[ \left( \frac{P_2}{P_1} \right)^{(n-1)/n} - 1 \right], where R is the gas constant, T is the inlet temperature, and P_1, P_2 are inlet and outlet pressures, respectively. This approach accounts for inefficiencies such as friction and cooling, enabling more accurate predictions of power requirements and stage efficiencies in turbomachinery. For turbines, similar polytropic expansions model the work extraction in gas and steam turbines, optimizing blade design and overall cycle performance. In internal combustion engines, polytropic processes approximate the expansion stroke during the power phase, where heat losses to cylinder walls and incomplete combustion lead to an effective polytropic exponent n \approx 1.3. This value captures the deviation from ideal adiabatic expansion (n = \gamma \approx 1.4), allowing engineers to estimate indicated work and more realistically in cycle simulations for spark-ignition and engines. Such modeling is crucial for predicting under varying loads and speeds. Refrigeration cycles, particularly vapor-compression systems, employ polytropic to represent the non-isentropic behavior in reciprocating or centrifugal compressors, incorporating cooling effects from jacket water or intercooling. By selecting an appropriate n (typically 1.1 to 1.3 for common refrigerants like R-134a), the process accounts for during compression, improving (COP) calculations and refrigerant selection. This is especially relevant in industrial chillers and units where efficiency gains from polytropic analysis reduce . The application of polytropic processes in traces back to the early , integrated into analyses of Rankine and Brayton cycles for and gas power plants to bridge ideal and real behaviors. Today, software tools like Aspen Plus facilitate detailed simulations of polytropic and in these cycles, using built-in models for multi-stage units and correlations to optimize plant design and operation.

Generalizations Beyond Ideal Gases

The polytropic process, originally formulated for ideal gases, can be generalized to real gases by incorporating the compressibility factor Z, which accounts for deviations from ideal behavior due to intermolecular forces and finite molecular volume. For real gases, the standard relation PV^n = \text{constant} is approximated using equations of state (EOS) such as the Peng-Robinson or Soave-Redlich-Kwong models to capture non-ideal effects during compression or expansion, though the constant n assumption may not hold accurately in high-pressure regimes where properties vary significantly, often requiring numerical methods or variable n. Alternatively, for van der Waals gases, the polytropic model incorporates attractive and repulsive intermolecular forces explicitly, leading to a modified EOS that influences the effective polytropic index n, often resulting in nonclassical gas dynamic behaviors such as anomalous wave speeds. These generalizations allow accurate prediction of work and in real gas systems, such as supercritical fluids in compressors. In multi-phase systems involving liquids and vapors, polytropic approximations are applied to model processes in devices like and condensers, where phase transitions and play key roles. For liquid pumping, due to low , the fluid is treated as incompressible with constant v, so pump work is calculated as v (p_2 - p_1); polytropic modeling is rarely applied, but if used, n approaches infinity. In condensers, vapor condensation is modeled as an (n = 0), where heat rejection occurs at nearly constant while decreases due to change, facilitating energy balance assessments in cycles. These adaptations extend the polytropic framework beyond single-phase gases, though they require empirical adjustments for effects. Beyond engineering, polytropic processes find interdisciplinary applications, such as in where stellar polytropes model self-gravitating spheres under . In this context, the Lane-Emden equation describes and profiles via P = K \rho^{1 + 1/n}, with n = 3 providing stability for convective stars like those on the , independent of central and approximating radiative envelopes. In acoustics, sound wave propagation in fluids is treated as a near-polytropic process with n approaching the adiabatic index \gamma \approx 1.4 for air, enabling predictions of wave speed and attenuation in stratified or bubbly media without significant generation. These uses highlight the versatility of polytropic models in scaling complex, large-scale phenomena. Despite these extensions, polytropic processes have limitations in scenarios involving phase changes or turbulent flows. During phase transitions, such as or , the abrupt release or absorption of violates the quasi-equilibrium assumption, causing deviations from the PV^n relation and requiring hybrid models like those combining polytropic and isenthalpic steps. In turbulent flows, viscous and chaotic mixing introduce non-reversible effects that render the constant n approximation invalid, often necessitating alternatives like isentropic processes for high-speed compressible flows where shocks dominate. For such cases, more advanced or numerical methods are preferred over polytropic simplifications. Recent advances in the leverage (CFD) and (ML) to determine effective polytropic exponents in complex flows, overcoming traditional analytical limitations. CFD simulations, using volume-of-fluid or multiphase models, extract n from pressure-volume data during transient events like hydraulic shocks, achieving for real gas and multi-phase interactions. ML techniques, such as , further enhance this by predicting variable n in turbulent or relativistic hydrodynamics, reducing computational costs by orders of magnitude while incorporating EOS corrections for non-ideal behaviors. These methods have enabled precise modeling in applications from pipeline flows to astrophysical simulations.

References

  1. [1]
    POLYTROPIC PROCESS - Thermopedia
    Polytropic was originally coined to describe any reversible process on any open or closed system of gas or vapor which involves both heat and work transfer.
  2. [2]
    Polytropic Process | Definition & Characteristics | nuclear-power.com
    A polytropic process is any thermodynamic process that can be expressed by the following equation: pVn = constant. The polytropic process can describe gas ...
  3. [3]
    [PDF] All About Polytropic Processes - SMU Physics
    Aug 26, 2022 · It turns out that these four processes are examples of a more general ideal-gas process, called a polytropic process, in which the specific heat ...
  4. [4]
    All About Polytropic Processes | The Physics Teacher - AIP Publishing
    Sep 1, 2022 · It turns out that these four processes are examples of a more general ideal-gas process, called a polytropic process, in which the specific heat ...Processes with constant... · Negative specific heat · Thermodynamics with...
  5. [5]
    Definition of polytropic process - Physics Stack Exchange
    Mar 14, 2017 · The term "polytropic" was firstly introduced by German physicist and engineer G. Zeuner probably about 1866. The starting point for him was ...Missing: history | Show results with:history
  6. [6]
    Centrifugal Compressor Polytropic Performance—Improved Rapid ...
    May 28, 2021 · 1.1. Polytropic History. A compressor polytropic performance calculation method was first documented in the 1860′s by Zeuner as discussed in the ...
  7. [7]
    Polytropic process in a closed system - Thermodynamics - tec-science
    Jul 11, 2021 · Polytropic processes obey p⋅V^n=constant, where n is the polytropic index. Special cases include isobaric, isothermal, isentropic, and ...Missing: quasi- static<|control11|><|separator|>
  8. [8]
    Simple realization of the polytropic process with a finite-sized reservoir
    Jul 1, 2023 · The ideal gas is compressed or expanded quasi-statically while in thermal contact with a finite reservoir with constant heat capacity, and the ...
  9. [9]
    [PDF] Polytropic Processes
    A polytropic process takes the general form of the common thermodynamic processes, and is defined by the equation. pV n = C or p1V n. 1 = p2V n. 2 where n is ...<|control11|><|separator|>
  10. [10]
    Polytropic Index - an overview | ScienceDirect Topics
    The polytropic index is defined as a parameter in a polytropic process that characterizes the relationship between pressure and volume, where it is not equal to ...<|separator|>
  11. [11]
    [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 ...
  12. [12]
    Derivation of the formulas for work and heat of a polytropic process
    Nov 7, 2021 · The aim of this article is to derive the formulas for calculating work and transferred heat for polytropic thermodynamic processes.
  13. [13]
    [PDF] Polytropic Process of an Ideal Gas
    Page 7. Remarks. For an adiabatic process p. 1. V. 1 k = p. 2. V. 2 k. Work done during an adiabatic process: W. 12. = (p. 1. V. 1. – p. 2. V. 2. )/(k-1).
  14. [14]
    [PDF] First Law of Thermodynamics Control Mass (Closed System)
    • polytropic processes: where PV n = C. • examples of polytropic processes include: Isobaric process: if n = 0 then P = C and we have a constant pressure ...
  15. [15]
    [PDF] Polytropic Process - WPI
    A polytropic process is a quasiequilibrium process described by pV^n = constant, where n can take any value from -∞ to +∞.
  16. [16]
    Constant Pressure - Physics
    A constant pressure process is called an isobaric process. An example is a gas in a container sealed with a piston that is free to slide up and down. If heat ...Missing: applications | Show results with:applications
  17. [17]
    [PDF] Notes on Thermodynamics, Fluid Mechanics, and Gas Dynamics
    Dec 15, 2021 · If n = 0, then the process is isobaric (constant pressure). If n = 소с, then the process is isometric (constant volume).
  18. [18]
    3.2 Work, Heat, and Internal Energy - UCF Pressbooks
    From Equation 3.4 and the ideal gas law, W = ∫ V 1 V 2 p d V = ∫ V 1 V 2 ( n R T V ) d V .
  19. [19]
    Thermodynamic processes
    An isobaric process is a process that occurs at constant pressure. We then have W = P(V2 - V1). If the pressure of an ideal gas is kept constant, then the ...Missing: applications | Show results with:applications
  20. [20]
    [PDF] Lecture 15
    This is a reversible process. 1) The entropy of a gas increases during a quasi-static isothermal expansion. What happens to the entropy of the environment? 2) ...
  21. [21]
    3.3 The Carnot Cycle - MIT
    A Carnot cycle is shown in Figure 3.4. It has four processes. There are two adiabatic reversible legs and two isothermal reversible legs.
  22. [22]
    [PDF] Chap. 3 Some general statements
    d(∆h) = (∂∆h/∂T)p dT + (∂∆h/∂p)T dp. For an isobaric process, only the second term vanishes and we have d(∆h)p ≡dL = ...
  23. [23]
    [PDF] Chap. 2 - Equation of State (Ideal Gas Law)
    One cannot assume that the ideal gas law is valid at. T = 0 K. As temperature increases, so does pressure, and vice versa. This assumes that V = const. (Does ...
  24. [24]
    [PDF] Ideal gas law • Quasi-static processes: isochoric, isobaric
    Absolute Zero and Absolute Temperature. • property changes linearly with temperature: e.g., pressure of constant-volume gas. • p = 0 for all gases at p due to ...
  25. [25]
    [PDF] PHYS 1220, Engineering Physics, Chapter 19 – The First Law of
    • For an ideal gas undergoes a isochoric process: ΔU =Q . dU=nCV dT. Page 4 ... - Equation of state for Ideal gas undergoes an Adiabatic process. • dQ=0.
  26. [26]
  27. [27]
  28. [28]
    4.3 Work – Introduction to Engineering Thermodynamics
    Consider an ideal gas undergoing a polytropic process. At the initial state: P1=200 kPa, v1=0.05 m3/kg. At the final state ...
  29. [29]
    [PDF] Polytropic Efficiency
    Nov 9, 2005 · γ-1 γ η the polytropic efficiency of the compressor is identical, i.e T. 3 is determined from. ∙. T3 ⎛ p3 ⎞ pc the polytropic efficiency ...
  30. [30]
    [PDF] Chapter 4 Entropy and a Move Towards Chaos
    More generally, we can also define a process to be polytropic in which the ... n = 0 for isobaric process (PV 0 = P = constant). 2. n = 1 for isothermal ...<|control11|><|separator|>
  31. [31]
    How to Estimate Compressor Efficiency - PetroSkills
    Jul 6, 2015 · When calculating the compressor head and discharge temperature the efficiency used will be isentropic or polytropic (isentropic efficiency is ...
  32. [32]
    Centrifugal compressor polytropic compression calculation
    The following formula can be used to calculate the polytropic compression power for a centrifugal compressor : Ppoly = (n*Z*R'*T1)/(n-1)*[(P2/P1)(n-1)/n-1]* ...
  33. [33]
    [PDF] Polytropic Exponents for Common Refrigerants - Purdue e-Pubs
    A polytropic process, where PV=constant, is a model for compression/expansion. This paper presents a method for determining the best polytropic exponent for ...
  34. [34]
    [PDF] THERMODYNAMICS OF THE REFRIGERATION CYCLE
    1 – 2 polytropic compression on the condensation pressure (for comparison. 1 – 2' isentropic compression). 2 – 2'' isobaric cooling, deheating of the.
  35. [35]
    Aspen Plus Unit Operations Technical Documentation - COMPR
    Nov 3, 2020 · For polytropic compressors the outlet temperature, if known, may be entered as a second parameter. In this case the polytropic efficiency will ...
  36. [36]
    [PDF] GPPS-TC-2024-0110
    Sep 6, 2024 · Using this oversimplified model for the polytropic process, the polytropic work and polytropic efficiency are calculated as follows: w̿p ...Missing: original | Show results with:original
  37. [37]
    Molecular interpretation of nonclassical gas dynamics of dense ...
    May 16, 2006 · The van der Waals polytropic gas model is used to investigate the role of attractive and repulsive intermolecular forces and the influence ...
  38. [38]
    Simplified modeling of scroll compressors with vapor injection
    Liquid Injection (LI) involves injecting liquid-phase refrigerant from the condenser outlet directly into the compressor (Ayub et al., 1992; Dutta et al., 2001; ...
  39. [39]
    [PDF] ME 418 Lecture 13 - Vapor Compression Equipment - I
    25 F with 5 F superheat and leaves the condenser as saturated liquid at 120 F. The polytropic exponent for the process is 1.2. Page 19. Pressure p e p c. V cl.
  40. [40]
    [PDF] POLYTROPES
    Polytropes are self-gravitating gaseous spheres that were, and still are, very useful as crude approximation to more realistic stellar models.
  41. [41]
    Investigation on propagation mechanism of leakage acoustic waves ...
    This paper establishes a theoretical calculation model of sound speed based on the bubble distribution pattern in the pipeline axial direction.
  42. [42]
    [PDF] Effects of Two-Phase Flow in a Multistage Centrifugal Compressor
    Sep 22, 2016 · A Calculation of Input Power, Adiabatic Efficiency and Polytropic Efficiency ... polytropic efficiency. 5.6 Discussions on the Compressor ...
  43. [43]
    Using CFD analysis to evaluate the performance of a natural gas ...
    Next, the polytropic exponent (n) is determined from the pressure and temperature data. Afterward, it is used to calculate the ideal work for a polytropic ...
  44. [44]
    Machine Learning-Driven Conservative-to-Primitive Conversion in ...
    We present a novel machine learning (ML)-based method to accelerate conservative-to-primitive inversion, focusing on hybrid piecewise polytropic and ...