Fact-checked by Grok 2 weeks ago

Compressibility

Compressibility is a fundamental property of that quantifies the relative change in of a substance in response to an applied , typically under isothermal conditions, and is defined as the negative reciprocal of the K, where K = -\frac{\Delta P}{\Delta V / V}. The isothermal compressibility \kappa_T is mathematically expressed as \kappa_T = -\frac{1}{V} \left( \frac{\partial V}{\partial P} \right)_T, measuring how much a material's volume decreases with increasing pressure at constant . In solids and liquids, compressibility is generally low, indicating high resistance to volume change; for example, water has an isothermal compressibility of approximately $4.6 \times 10^{-10} Pa^{-1} at 20°C and atmospheric pressure, making it nearly incompressible under typical conditions. Gases, however, exhibit high compressibility, often characterized by the compressibility factor Z, defined as Z = \frac{PV}{nRT}, which deviates from unity for real gases due to intermolecular forces and finite molecular volume, as seen in the van der Waals equation of state. Compressibility plays a critical role in various fields, including , , and ; in fluid flows, it determines whether a flow is treated as incompressible (Mach number M < 0.3, density variation <5%) or compressible (M > 0.3), affecting phenomena like shock waves in high-speed . Adiabatic compressibility \kappa_S, related to isothermal by \kappa_S = \kappa_T / \gamma where \gamma = C_P / C_V, is relevant in dynamic processes like sound propagation. Understanding compressibility is essential for applications ranging from reservoir modeling to under extreme pressures.

Fundamental Concepts

Definition

Compressibility refers to the ability of a to undergo a relative decrease in volume in response to an increase in applied , a observed across solids, liquids, and gases. This measure quantifies how much the volume of a substance diminishes proportionally when subjected to compressive forces, reflecting the internal structure and intermolecular forces within the . In practical terms, it describes the susceptibility of to being compacted under , with the extent of change depending on the state of the substance. The formal concept of compressibility in was developed in the amid investigations into the behavior of gases, influencing later thermodynamic studies. Units for compressibility are typically given in inverse pascals (Pa⁻¹), reflecting the relative volume change per unit pressure, or alternatively per atmosphere for contexts. Compressibility serves as the reciprocal of the , providing an inverse measure of a material's resistance to . Gases demonstrate high compressibility due to the significant empty space between molecules, allowing substantial volume reduction; for instance, air at can compress markedly under moderate pressures like those in pneumatic systems. In contrast, liquids exhibit much lower compressibility, as seen in , where volume decreases by only about 0.5% even at depths of 1,000 meters in the , owing to closer molecular packing. Solids possess the least compressibility among the states of ; metals like require extreme pressures, such as those in geological deep conditions, to achieve measurable volume changes. A broad distinction exists between static and dynamic compressibility: static compressibility characterizes the steady-state volume response to gradually applied in conditions, while dynamic compressibility pertains to transient responses under rapid fluctuations, such as . This differentiation highlights how the rate of pressure application can influence observed behavior in materials.

Mathematical Formulation

Compressibility is mathematically defined as the relative volume change induced by a pressure change, expressed in finite difference form as \beta = -\frac{1}{V} \frac{\Delta V}{\Delta P}, where V is the initial , \Delta V is the change in , and \Delta P is the change in . The negative sign accounts for the typical decrease in volume with increasing pressure. This definition quantifies the material's resistance to uniform under applied . For continuous and small pressure variations, the definition adopts an infinitesimal form using : \beta = -\frac{1}{V} \left( \frac{\partial V}{\partial P} \right), where the partial derivative assumes other variables, such as , are held constant unless specified otherwise. This form arises from the differential relation dV = \left( \frac{\partial V}{\partial P} \right) dP, rearranged to express the fractional volume change \frac{dV}{V} = -\beta \, dP. The assumption of small changes ensures in the volume-pressure relationship, approximating the response as proportional within the elastic limit. In derivations from volume-pressure relations, compressibility emerges as the of the volume-pressure for small perturbations. Starting from an empirical or theoretical relating V and P, differentiation yields the local compressibility at a given state point. For larger pressure ranges, the response may deviate from , leading to nonlinear compressibility where \beta varies with P, requiring of the rather than simple proportionality. This nonlinearity is particularly relevant in high-pressure regimes, such as deep geophysical contexts, though the suffices for modest changes. Compressibility is the reciprocal of the K, defined as K = -\frac{\partial P}{\partial (\ln V)} = \frac{1}{\beta}, linking the concept directly to elastic material properties. For gases, a dimensionless extension is the Z, defined as Z = \frac{PV}{nRT}, where n is the number of moles, R is the , and T is . For ideal gases, Z = 1; deviations indicate behavior, with Z < 1 at moderate pressures due to intermolecular attractions and Z > 1 at high pressures from repulsive forces. This factor provides a normalized measure of compressibility relative to ideal conditions.

Physical Properties and Relations

Bulk Modulus Connection

The bulk modulus K, a measure of a material's resistance to uniform compression, is defined as K = -V \left( \frac{\partial P}{\partial V} \right)_T, where V is volume, P is pressure, and the subscript T denotes constant temperature. This definition directly quantifies the stiffness of a material under hydrostatic pressure, contrasting with more compressible substances that exhibit larger volume changes for a given pressure increment. The reciprocal relationship between the and compressibility \beta arises from their fundamental definitions. Compressibility is given by \beta = -\frac{1}{V} \left( \frac{\partial V}{\partial P} \right)_T, which rearranges to \left( \frac{\partial V}{\partial P} \right)_T = -V \beta. Substituting into the bulk modulus expression yields K = -V \left( \frac{\partial P}{\partial V} \right)_T = -V / \left( \frac{\partial V}{\partial P} \right)_T = -V / (-V \beta) = \frac{1}{\beta}, establishing [K](/page/K) as the inverse of compressibility and highlighting how low compressibility corresponds to high stiffness. Bulk modulus is typically measured using ultrasonic techniques, which determine elastic wave speeds in the material to compute the adiabatic K_S = \rho c^2 (where \rho is and c is the speed of longitudinal ), or through hydrostatic compression experiments that directly apply uniform and monitor changes to obtain the isothermal K_T. These methods provide accurate values for both fluids and solids, with ultrasonic approaches favored for high-pressure scenarios due to their non-destructive nature. For liquids like , K_S and K_T differ minimally since \gamma \approx 1.01. Representative values illustrate the range of material stiffness: for , K \approx 2.2 GPa, indicating relatively low compressibility suitable for hydraulic applications, while for , K \approx 160 GPa reflects high stiffness essential in . The exhibits dependence on and , generally decreasing with increasing temperature due to enhanced atomic vibrations that soften the material, and increasing with pressure as interatomic distances compress and bonding strengthens. In , for instance, the bulk modulus of metals like decreases by about 0.3–0.5% per rise near but rises nonlinearly with pressure, following relations like K(P) = K_0 + K_0' P where K_0' is the (typically 4–7 for solids). These variations are critical for predicting behavior in extreme environments, such as deep-earth conditions or high-temperature processing.

Speed of Sound Relation

The speed of sound in a fluid medium is fundamentally linked to its compressibility through the adiabatic bulk modulus K_S, which measures resistance to uniform compression under adiabatic conditions. The propagation velocity c of longitudinal pressure waves is given by c = \sqrt{\frac{K_S}{\rho}}, where \rho is the mass density of the fluid. Since the adiabatic compressibility \beta_S = 1/K_S (with K_S = \gamma K_T and \gamma = C_P / C_V), this formula demonstrates an inverse square root dependence of the sound speed on compressibility: higher compressibility (larger \beta_S) results in slower wave propagation. This relation arises from the one-dimensional for small-amplitude disturbances in a compressible . Consider a element of length \Delta x subjected to a \Delta p, leading to a \Delta \xi. Applying Newton's second to the yields the \partial^2 \xi / \partial t^2 = (1/\rho) \partial p / \partial x. Relating the change to volume strain via the adiabatic (\Delta p = -K_S \Delta V / V) and assuming small , the wave speed emerges as c = \sqrt{K_S / \rho}, confirming the foundation of propagation. In , the situation differs due to the ability to support stresses, resulting in both longitudinal and transverse () waves. For longitudinal waves, the speed is c_L = \sqrt{\frac{K + \frac{4}{3} \mu}{\rho}}, where \mu is the , extending the fluid case by incorporating resistance to deformation. waves propagate at c_S = \sqrt{\mu / \rho}, which is absent in fluids where \mu = 0. This distinction highlights how compressibility alone governs fluid acoustics, while involve coupled compressional and effects. Representative values illustrate these differences: in air at (20°C, 1 atm), c \approx 343 m/s, reflecting high compressibility; in at 20°C, c \approx 1480 m/s, due to lower compressibility. In solids like aluminum, c_L \approx 6420 m/s and c_S \approx 3040 m/s, showcasing the role of . Experimental verification of these relations relies on compressibility measurements to predict and confirm speeds in acoustics and ultrasonics. For instance, ultrasonic pulse-echo techniques use time-of-flight data to compute c, which matches predictions from independently measured \beta_S and \rho, validating the formulas across frequencies from audible to gigahertz ranges. Such methods are essential in and geophysical probing, where discrepancies inform material inhomogeneities. These derivations and formulas assume isotropic and homogeneous , limiting applicability to polycrystalline or composite materials with directional variations in properties.

Thermodynamic Compressibility

Isothermal Compressibility

Isothermal compressibility, often denoted as \kappa_T or \beta_T, quantifies the relative change in of a substance in response to a change while maintaining constant temperature. It is formally defined by the equation \kappa_T = -\frac{1}{V} \left( \frac{\partial V}{\partial P} \right)_T, where V is the volume, P is the pressure, and the subscript T indicates that the temperature is held constant. The negative sign ensures that \kappa_T is positive, as volume typically decreases with increasing pressure. This property is fundamental in thermodynamics for describing equilibrium processes where heat exchange with the surroundings is permitted. Thermodynamic identities link isothermal compressibility to other key properties, such as the isobaric thermal expansion coefficient \alpha = \frac{1}{V} \left( \frac{\partial V}{\partial T} \right)_P. A central relation derived from Maxwell's relations and the cyclic rule is C_P - C_V = \frac{T V \alpha^2}{\kappa_T}, where C_P and C_V are the heat capacities at constant pressure and volume, respectively. This equation highlights how \kappa_T influences the distinction between heat capacities and underscores its role in thermodynamic stability and response functions. For instance, rearranging yields \kappa_T = \frac{T V \alpha^2}{C_P - C_V}, providing a pathway to compute \kappa_T from measurable quantities like \alpha and heat capacities. For an ideal gas, the isothermal compressibility takes a simple form: \kappa_T = \frac{1}{P}, reflecting the direct proportionality of volume to pressure at fixed temperature via the ideal gas law PV = nRT. This inverse pressure dependence arises because intermolecular forces are negligible, allowing straightforward volume adjustments without thermal barriers. In contrast, real gases exhibit deviations captured by equations like the van der Waals model, (P + \frac{a n^2}{V^2})(V - n b) = n R T, where a accounts for attractions and b for molecular volume. These corrections introduce a compressibility factor Z = \frac{P V}{n R T}, such that \kappa_T = \frac{1}{P} \left[ 1 - \left( \frac{\partial \ln Z}{\partial \ln P} \right)_T \right], leading to values of \kappa_T that can exceed or fall below the ideal case depending on conditions near the critical point. Isothermal compressibility plays a vital role in applications involving , such as , where it helps model changes during measurements at constant , ensuring accurate balance calculations. In equilibria studies, \kappa_T is essential for predicting stability in systems under controlled laboratory pressures; for example, measurements of liquid compressibility guide the design of high-pressure experiments to map boundaries and mixing behaviors in solutions. Unlike adiabatic compressibility, which applies to rapid processes without , \kappa_T governs slower, isothermal scenarios typical in these lab settings.

Adiabatic Compressibility

Adiabatic compressibility, denoted as \beta_S or \kappa_S, quantifies the relative change in of a substance under variation while maintaining constant , characteristic of reversible adiabatic processes where no is exchanged with the surroundings. It is formally defined as \beta_S = -\frac{1}{V} \left( \frac{\partial V}{\partial P} \right)_S, where V is , P is , and the subscript S indicates the isentropic condition. This measure is particularly relevant for dynamic phenomena, such as rapid compressions, where cannot be established due to the absence of . The adiabatic compressibility relates to the isothermal compressibility \beta_T through the \gamma = C_P / C_V, where C_P and C_V are the heat capacities at constant pressure and volume, respectively: \beta_S = \beta_T \frac{C_V}{C_P} = \frac{\beta_T}{\gamma}. Since \gamma > 1 for most substances, \beta_S < \beta_T, indicating that substances are less compressible under adiabatic conditions than isothermal ones, as the lack of heat transfer leads to temperature changes that resist volume alteration. This relation can be derived using thermodynamic potentials and Maxwell relations. Starting from the differential of entropy dS = \left( \frac{\partial S}{\partial T} \right)_P dT + \left( \frac{\partial S}{\partial P} \right)_T dP, where \left( \frac{\partial S}{\partial T} \right)_P = \frac{C_P}{T} and, by the Maxwell relation from the Gibbs free energy, \left( \frac{\partial S}{\partial P} \right)_T = -\left( \frac{\partial V}{\partial T} \right)_P = -V \alpha_P (with \alpha_P the isobaric thermal expansion coefficient). For constant entropy (dS = 0), \left( \frac{\partial T}{\partial P} \right)_S = \frac{V \alpha_P T}{C_P}. The volume differential is dV = \left( \frac{\partial V}{\partial P} \right)_T dP + \left( \frac{\partial V}{\partial T} \right)_P dT, so at constant S, \left( \frac{\partial V}{\partial P} \right)_S = \left( \frac{\partial V}{\partial P} \right)_T + \left( \frac{\partial V}{\partial T} \right)_P \left( \frac{\partial T}{\partial P} \right)_S. Substituting yields \left( \frac{\partial V}{\partial P} \right)_S = \left( \frac{\partial V}{\partial P} \right)_T + V^2 \alpha_P^2 T / C_P, and thus \beta_S = \beta_T - V \alpha_P^2 T / C_P. Relating to \gamma via C_P - C_V = V \alpha_P^2 T / \beta_T gives \gamma = \beta_T / \beta_S. For ideal gases, the relation simplifies further, as \beta_T = 1/P and \beta_S = 1/(\gamma P). In dry air at standard conditions, primarily composed of diatomic molecules like N₂ and O₂, \gamma \approx 1.4, so \beta_S \approx 0.714 / P, making air stiffer under adiabatic compression than isothermal. This difference is crucial for sound propagation, where rapid pressure oscillations occur adiabatically, causing small temperature fluctuations that enhance wave speed compared to an isothermal scenario. Adiabatic compressibility is typically measured indirectly through the adiabatic bulk modulus K_S = 1/\beta_S, often using ultrasonic pulse-echo techniques that determine sound velocity c = \sqrt{K_S / \rho} (with \rho density), from which K_S is derived. These methods involve sending acoustic pulses through the sample and analyzing echo timings and amplitudes, suitable for liquids and solids under controlled conditions.

Applications in Geosciences

Rock and Mineral Compressibility

Rocks and minerals, particularly silicates and oxides prevalent in the Earth's crust and mantle, exhibit low compressibility under ambient conditions due to their tightly bonded crystal structures. For instance, quartz (SiO₂), a common silicate mineral, has an isothermal compressibility β of approximately 2.7 × 10^{-11} Pa^{-1} at room temperature and atmospheric pressure, corresponding to a bulk modulus K_T of about 37 GPa. This low value reflects the resistance of framework silicates to volume reduction, with similar trends observed in oxides like periclase (MgO), where K_T exceeds 160 GPa at ambient conditions. Under increasing pressure, compressibility typically decreases as the bulk modulus rises, with the pressure derivative K' often around 4-6 for these materials, enhancing their stability in deep Earth environments. To quantify compressibility at high pressures relevant to the geosphere, pressure-volume equations of state (EOS) are employed, with the being widely used for fitting experimental data on minerals. This third-order formulation relates volume V to pressure P through parameters including the zero-pressure volume V₀, bulk modulus K₀, and its pressure derivative K'₀, as derived from finite strain theory and validated against hydrostatic compression experiments. For quartz, diamond anvil cell (DAC) measurements yield EOS parameters V₀ = 112.98 ų, K₀ = 37.1 GPa, and K'₀ = 6.0, accurately describing compression up to several tens of GPa where structural distortions like Si-O-Si bond angle reductions occur. These DAC experiments, pioneered in the 1980s, provide precise P-V-T data essential for modeling mineral behavior in the mantle. Porosity and fractures significantly elevate the compressibility of rocks compared to their constituent minerals, with effects most pronounced in sedimentary formations. Sedimentary rocks, such as sandstones and limestones, often possess porosities of 10-30%, leading to bulk compressibilities 10-100 times higher than crystalline values (β ≈ 10^{-10} to 10^{-9} Pa^{-1}), as pore collapse and crack closure dominate under differential stress. In contrast, igneous rocks like granite and basalt, with low porosities (<5%) and minimal fractures, display compressibilities closer to those of individual minerals (β ≈ 10^{-11} Pa^{-1}), behaving more rigidly due to their dense, interlocking crystal matrices. This distinction arises from the four key compressibilities in porous media—bulk, pore, skeletal, and grain—where higher porosity amplifies the bulk response to confining pressure. In geophysical modeling, mineral and rock compressibility plays a critical role in simulating mantle convection and deep Earth dynamics. Accurate EOS data inform density profiles and phase transitions, influencing convective vigor and heat transport; for example, compressible formulations reveal asymmetries in boundary layers and more realistic seismic anomalies when integrated with thermochemical models. Self-consistent mineral physics calculations, incorporating compressibility alongside thermal expansivity, enhance predictions of mantle structure, linking laboratory-derived properties to global-scale processes like plume dynamics. Data from high-pressure experiments since the 1980s, archived in mineral physics resources such as those from the Consortium for Materials Properties Research in Earth Sciences (COMPRES), underpin these models.

Seismic Wave Propagation

In seismic wave propagation through the Earth, compressibility plays a central role in determining the velocity of primary (P) waves, which are compressional waves that rely on the material's resistance to volume change. The adiabatic compressibility, denoted as \kappa_S = 1/K_S where K_S is the adiabatic bulk modulus, governs this process because seismic wave frequencies are high enough that heat exchange is negligible, approximating adiabatic conditions. The P-wave velocity V_P in an isotropic, homogeneous medium is given by V_P = \sqrt{\frac{K_S + \frac{4}{3} G}{\rho}}, where G is the shear modulus and \rho is the density; this formula highlights how lower compressibility (higher K_S) increases V_P by enhancing the material's stiffness against compression. As seismic waves travel deeper into the Earth, compressibility decreases with increasing pressure, leading to higher wave speeds in the mantle. In the upper mantle, pressures around 1-5 GPa result in V_P values of approximately 8 km/s, but by the lower mantle at depths exceeding 2000 km and pressures over 100 GPa, compressibility drops significantly due to the compression of mineral lattices, pushing V_P to about 13-14 km/s. This depth-dependent reduction in compressibility, combined with rising temperature effects, produces the observed increase in seismic velocities with depth, as pressure dominates over thermal softening in solid-state mantle rocks. Anelasticity introduces attenuation to seismic waves, where partial melting zones exhibit higher effective compressibility due to the presence of low-viscosity melt pockets that enhance dissipative mechanisms like grain boundary sliding. In regions such as the asthenosphere or mid-ocean ridge axes with 1-5% partial melt, this increased effective \kappa_S lowers V_P by up to 5-10% and amplifies attenuation (measured by quality factor Q), as the melt's high compressibility (e.g., \kappa_S \approx 10^{-9} Pa^{-1} for basaltic melt versus $10^{-11} Pa^{-1} for solids) dominates the bulk response during wave passage. Seismic tomographic models leverage variations in compressibility to map subsurface structures, revealing cold, low-compressibility slabs in subduction zones where V_P anomalies exceed +2% due to eclogite-phase minerals with elevated K_S. At the core-mantle boundary (CMB), ultra-low velocity zones (ULVZs) show compressibility enhancements from partial melting or iron enrichment, reducing V_P by 10-30% over patches spanning hundreds of kilometers, as inferred from joint inversions of wave speeds and mineral elasticity data. These models integrate compressibility contrasts to delineate slab penetration to the CMB and chemical heterogeneities. Historically, the Adams-Williamson equation formalized the integration of compressibility into Earth's density profile under hydrostatic equilibrium and adiabatic conditions, expressed as \frac{d\rho}{dr} = -\frac{\rho g}{K_S} + \frac{\rho \alpha \gamma g}{C_P}, where the first term accounts for self-compression via \kappa_S = 1/K_S, \alpha is thermal expansivity, \gamma the Grüneisen parameter, g gravity, r radius, and C_P heat capacity at constant pressure; the second term corrects for adiabatic decompression. Derived in 1923, this equation enabled early density models of the mantle by combining seismic velocities with measured compressibilities of silicates, predicting a core-mantle density jump of about 4-5 g/cm³.

Applications in Fluid Mechanics

Compressible Flow Regimes

In compressible fluid dynamics, flow regimes are primarily distinguished by the Mach number, defined as the ratio of the flow velocity v to the local c, denoted as Ma = v / c. Flows with Ma < 0.3 exhibit negligible density variations and can be approximated as incompressible, simplifying analysis by assuming constant . In contrast, when Ma > 0.3, compressible effects become significant, leading to substantial changes in , , and that must be accounted for in modeling. This threshold arises because the represents the propagation speed of disturbances, influencing how compressibility alters behavior. A key distinction in compressible regimes involves the , which enforces mass conservation. For steady , it takes the form \nabla \cdot (\rho \mathbf{v}) = 0, where \rho is the variable and \mathbf{v} is the vector, reflecting density fluctuations due to or . In incompressible approximations, is , reducing the equation to \nabla \cdot \mathbf{v} = 0, which implies a divergence-free field and ignores density changes. These variations highlight how compressible flows require tracking evolution, particularly in scenarios with pressure gradients that induce volumetric changes. Compressible regimes also necessitate the inclusion of an energy equation to capture alterations from heating. As fluid parcels accelerate or decelerate, work done by forces converts to , raising and further influencing via thermodynamic relations. This coupling between mechanical and thermal processes is absent in incompressible models, where temperature effects are often decoupled. Representative examples illustrate these regimes. In compressible flow through pipelines, such as natural gas transport, Mach numbers remain below 1 but exceed 0.3, causing reductions along the line due to frictional drops, which affect capacity and require compressible corrections for accurate predictions. Conversely, supersonic flows in converging-diverging nozzles, like those in rocket engines, achieve Ma > 1, where area changes drive acceleration and pronounced shifts, enabling efficient thrust generation. Numerical modeling of compressible regimes often employs finite volume methods, which discretize the conservation equations over control volumes to handle variable and shock-like features robustly. These methods conserve , , and locally, making them suitable for simulating density gradients and heating effects in complex geometries. High-resolution schemes within finite volume frameworks, such as those using flux limiters, ensure stability and accuracy for both and supersonic conditions.

Equations of State for Fluids

Equations of state (EOS) for fluids provide essential relations between P, volume V, and T, enabling the prediction of compressibility effects in various applications. For liquids, isothermal EOS often approximate the isothermal compressibility \beta_T = -\frac{1}{V} \left( \frac{\partial V}{\partial P} \right)_T as nearly constant over moderate ranges. The Tait exemplifies this approach, expressing volume change as \frac{V_0 - V}{V_0} = C \log_{10} \left( \frac{B + P}{B + P_0} \right), where V_0 is the reference volume at P_0 (typically atmospheric), and C and B are empirical constants fitted to experimental data. This form, originally developed for and in 1888, accurately models density variations under hydrostatic compression, with B often approximating the and C near unity for many liquids. Differentiating the Tait yields \beta_T \approx \frac{C}{(B + P) \ln 10}, confirming its utility for fluids where compressibility decreases weakly with . Polytropic processes extend these concepts to dynamic scenarios, described by P V^\gamma = \text{constant}, where \gamma is the polytropic index. For adiabatic conditions, \gamma equals the ratio of specific heats C_P / C_V, linking directly to adiabatic compressibility \beta_S = -\frac{1}{V} \left( \frac{\partial V}{\partial P} \right)_S = \frac{1}{\gamma P} for ideal gases, though applicable to dense fluids via thermodynamic relations. This EOS captures reversible processes without , providing closure for modeling in fluids where \beta_S < \beta_T. For real fluids like hydrocarbons, cubic EOS such as the Peng-Robinson model incorporate a Z = PV / RT to account for non-ideal behavior. The equation takes the form P = \frac{RT}{V - b} - \frac{a \alpha}{V(V + b) + b(V - b)}, where a and b are substance-specific parameters derived from critical properties, and \alpha adjusts for temperature. Solving for Z yields a cubic , with the largest real root used for vapor phases; this enables prediction of phase equilibria and compressibility deviations in fluids, improving accuracy over van der Waals models for pressures up to critical points. In high-pressure environments, such as ocean depths exceeding 1000 m (pressures ~100 ), water's compressibility influences acoustic . The isothermal compressibility of decreases from ~4.5 × 10^{-10} Pa^{-1} at surface conditions to lower values at depth, per the TEOS-10 formulation, which uses a Gibbs to compute \beta_T = -\frac{1}{\rho} \left( \frac{\partial \rho}{\partial P} \right)_{S_A, T}. This variation raises speed c = 1 / \sqrt{\beta_S \rho} (using adiabatic \beta_S) by up to 20 m/s over 4000 m depth, critical for modeling. Empirical fits like the Tait validate these trends for up to 100 . Validation of these EOS relies on laboratory shock compression experiments, which probe extreme states. For instance, plate-impact experiments on up to 200 GPa confirm Mie-Grüneisen forms integrated with thermal models, matching Hugoniot data within 1% for pressure-volume relations and validating compressibility predictions under . Similar gas-gun tests on and hydrocarbons refine parameters in cubic and empirical EOS, ensuring reliability for predictive simulations.

Applications in Aerodynamics

High-Speed Aerodynamic Effects

In high-speed aerodynamics, compressibility effects become prominent as aircraft approach transonic speeds (Mach numbers around 0.8 to 1.2), profoundly influencing lift, drag, and overall stability. At these regimes, local airflow over wings and fuselages can accelerate to supersonic speeds, leading to shock wave formation that disrupts smooth flow and causes boundary layer separation. This results in a dramatic rise in drag, primarily due to wave drag and induced separation, which can degrade aircraft performance by increasing fuel consumption and reducing control authority. For instance, near Mach 1, shock-induced separation often significantly increases total drag, necessitating design adjustments to maintain efficiency and stability. A key theoretical tool for analyzing subsonic compressibility effects is the Prandtl-Glauert , which approximates how modifies incompressible solutions. Derived from linearizing the equations, it scales pressures and forces by a factor of \beta = \frac{1}{\sqrt{1 - M_a^2}}, where M_a is the . This effectively increases the apparent thickness and of airfoils, amplifying but also contributing to earlier onset of drag rise as approaches . For example, at M_a = 0.8, \beta \approx 1.67, meaning lift coefficients from low-speed tests must be adjusted upward by this factor to predict compressible behavior accurately. The method, originally proposed by Prandtl in 1921 and refined by Glauert, remains foundational for preliminary design despite its limitations near conditions. To mitigate in flight, the —developed by Richard Whitcomb at NACA in 1952—guides fuselage and wing integration for smoother cross-sectional area distribution along the aircraft axis. This principle minimizes shock strength by avoiding abrupt area changes, which would otherwise generate interfering wave systems. Iconic applications include the "Coke-bottle" fuselage shape, where the midsection is narrowed near the to compensate for wing thickness, reducing rise by up to 30-50% in early tests. Aircraft like the benefited from this redesign, achieving supersonic dash capability after initial failures to meet performance goals. Historical insights into these effects emerged from 1940s wind tunnel tests during , when high-altitude dives revealed unexpected buffeting and control loss in fighters like the . NACA experiments at , using pressurized tunnels to simulate altitude, quantified compressibility burble—early shock-induced separation—pushing designs toward thinner wings and sweep to delay critical numbers. These findings, documented in reports from 1941 onward, informed post-war supersonic programs and underscored the need for integrated high-speed testing. Today, modern simulations via (CFD) resolve these phenomena by solving the compressible Navier-Stokes equations, capturing viscous effects, shocks, and in complex geometries. Tools like structured overset grids enable accurate prediction of rise and stability derivatives for configurations, reducing reliance on costly iterations. For example, ’s FUN3D solver has been validated against flight data for fighters, showing drag predictions within 5% of experiments while incorporating real-gas effects at high speeds. This approach supports optimization, ensuring like the F-35 maintain performance across speed regimes.

Shock Wave Formation

In compressible flows exceeding the speed of sound, shock waves form as abrupt discontinuities where the fluid properties—such as pressure, density, and temperature—change rapidly across a thin region, driven by the finite compressibility of the medium that prevents smooth adjustment to sudden changes in flow conditions. These discontinuities arise because information about downstream obstacles cannot propagate upstream faster than the flow itself, leading to a buildup of compression that manifests as a shock. The structure of a shock wave is governed by conservation laws applied across the discontinuity, ensuring mass, momentum, and energy balance. The fundamental relations describing these jumps are the Rankine-Hugoniot equations, derived from the integral form of the Euler equations for one-dimensional flow, which relate the states upstream (denoted as 1) and downstream (denoted as 2) of the shock. For a normal shock perpendicular to the flow direction, the pressure jump ΔP and density jump Δρ satisfy ΔP = ρ₁ u₁² (1 - ρ₁/ρ₂), where u₁ is the upstream velocity, but in the limit of weak shocks (Mach number M₁ approaching 1 from above), this simplifies to ΔP/Δρ ≈ c², with c being the speed of sound, akin to the isentropic acoustic wave relation. Oblique shocks, inclined at an angle β to the upstream flow, occur when supersonic flow encounters a wedge or ramp, allowing partial turning of the flow while keeping it supersonic downstream if the deflection angle θ is below a critical value; the shock angle β is determined by the θ-β-M relation, tan θ = 2 cot β (M₁² sin² β - 1) / (M₁² (γ + cos 2β) + 2). For reflection off a wall, regular reflection produces two oblique shocks of equal strength, but at higher incidence angles, Mach reflection occurs, forming a Mach stem (normal shock) and reflected oblique shock, with detachment happening when the deflection exceeds the maximum for attached oblique shocks (θ_max ≈ 20° for γ=1.4 and M₁=2). Across any , the process is irreversible, resulting in an increase Δs = c_v ln[(2γ M₁² - (γ-1))((γ-1) M₁² + 2)/((γ+1) M₁²)] + R ln[(γ+1) M₁² / ((γ-1) M₁² + 2)], which quantifies the dissipation due to the shock's finite thickness involving and heat conduction, though idealized as discontinuous. This links directly to adiabatic compressibility, as the post-shock state follows the Hugoniot curve rather than the isentrope, with significant heating: the temperature ratio is given by \frac{T_2}{T_1} = \frac{[2 \gamma M_1^2 - (\gamma - 1)] [(\gamma - 1) M_1^2 + 2]}{(\gamma + 1)^2 M_1^2}, where γ is the specific heat ratio, illustrating how compressibility amplifies thermal effects in supersonic flows. For example, at M₁=2 and γ=1.4, T₂/T₁ ≈ 1.687, highlighting the rapid heating. In applications, shock waves in supersonic inlets of aircraft engines, such as mixed-compression designs, position oblique shocks to slow and compress incoming air efficiently while minimizing total pressure loss from entropy rise, optimizing thrust. Similarly, blast waves from explosions propagate as strong shocks with spherical expansion, where initial overpressure decays as r^{-3} in the adiabatic phase, causing downstream heating and damage through compressibility-driven compression. Experimental visualization of these phenomena in wind tunnels employs Schlieren imaging, which detects density gradients via light refraction, revealing shock structures as sharp dark-light boundaries; for instance, in supersonic facilities, it captures normal shocks in nozzles or oblique patterns around models at Mach 2-5.

Special and Advanced Cases

Negative Compressibility Phenomena

Negative compressibility phenomena refer to rare cases where materials exhibit an apparent decrease in or in under increasing hydrostatic , corresponding to a negative isothermal compressibility β = -(1/V)(∂V/∂P)_T < 0. While negative linear compressibility— along specific directions—is more commonly observed, negative compressibility affecting overall is rarer and typically signals underlying instabilities or transitions rather than stable equilibrium states. The primary mechanisms driving negative compressibility involve structural phase changes, electronic transitions, or multi-phase interactions that counteract conventional . In structural phase changes, can trigger reconfiguration of arrangements, leading to along specific directions or overall increase. In multi-phase systems, like partially saturated porous media, forces in bridges within microcracks can generate effective negative under certain conditions. These mechanisms often manifest transiently during dynamic processes rather than in static equilibria. Representative examples include certain perovskite structures under high pressure, where electronic transitions induce negative compressibility. For instance, in two-dimensional hybrid perovskites like (C8H17NH3)2PbBr4, compression triggers a configuration transition and rotation of organic carbon chains, resulting in a giant volume expansion of up to 12.9%, attributed to enhanced interlayer spacing and structural reconfiguration. In geophysical contexts, partially saturated rocks such as Mancos shale demonstrate apparent negative compressibility during laboratory experiments, where equal increases in confining stress and gas pore pressure lead to extensional volumetric strain and reduced shear wave velocities due to hysteresis in liquid bridges within microcracks. These effects highlight domain-specific behaviors in natural materials. Thermodynamically, negative compressibility violates the stability criterion that the bulk modulus K = 1/β must be positive for , as it implies a concave-down free energy landscape prone to or collapse. Consequently, such states are inherently and cannot persist in isolated systems without external constraints; observed cases are often effective or transient, occurring near critical points, during phase transitions, or in constrained geometries like composites. This instability underscores the need for careful interpretation in experimental contexts. Research on negative compressibility in geophysical contexts dates back to the , particularly in geothermal simulations where apparent negative responses arise in multi-phase systems. These findings have implications for modeling subsurface processes, such as carbon storage and rock behavior under . Ongoing investigations emphasize the role of such phenomena in enhancing our understanding of planetary interiors.

Compressibility in Metamaterials

Metamaterials are artificially engineered structures designed to exhibit effective mechanical properties not found in natural materials, including tailored compressibility that can be negative or zero. These properties arise from the periodic arrangement of subwavelength unit cells, allowing control over response under hydrostatic . In particular, effective compressibility β, defined as β = -(1/)(∂/∂), can be tuned below zero, leading to expansion under compression, or to zero for incompressible behavior. Lattice structures in metamaterials achieve effective negative compressibility (β < 0) through auxetic geometries featuring negative (ν < 0), where under uniaxial compression, transverse expansion occurs, enabling overall volume expansion under pressure via coordinated or rotation of rigid elements. For instance, re-entrant or chiral lattices deform such that the effective K = 1/β becomes negative in specific stress regimes, as demonstrated in experimental quasistatic tests on polymer-based architectures. This auxetic mechanism contrasts with conventional materials by leveraging geometric nonlinearity rather than intrinsic material response. Zero compressibility in metamaterials is realized through designs like photonic crystals or cellular foams incorporating mechanisms that resist volumetric change under hydrostatic loading, maintaining near-constant volume via constrained deformation paths in the unit cells. Topology-optimized orthotropic lattices, for example, can exhibit exactly zero effective compressibility by balancing positive and negative contributions from anisotropic elements, as shown in computational models validated against finite element simulations. These structures approximate incompressibility (β ≈ 0) without relying on high-stiffness base materials, instead using instability thresholds to lock volume. Fabrication of such metamaterials has advanced since the 2010s with additive manufacturing techniques like and nanoscale assembly, enabling precise control over complex geometries. Polymer lattices, such as those produced via , demonstrate negative effective compressibility under uniaxial compression, where applied pressure induces transverse expansion and negative bulk response due to auxetic unit cells, with experimental measurements confirming β < 0 up to strains of 10-20%. Nanomaterial integration further enhances tunability, allowing multifunctional responses in lightweight composites. Applications of compressibility-tailored metamaterials span multiple fields, including acoustic cloaking devices where negative or zero β enables wave manipulation without reflection, such as in pentamode structures that guide sound around obstacles. In , these materials provide superior shock absorption through energy dissipation in buckling modes, reducing impact loads in composite panels. Biomedical uses include auxetic stents that expand radially under deployment stress while maintaining vessel patency, leveraging negative compressibility for better conformability and reduced restenosis risk. The theoretical foundation for these effective properties relies on homogenization theory, which averages microscopic deformations to derive macroscopic medium parameters like β for periodic structures. This approach, applied via asymptotic expansions or finite element methods, predicts how local geometries yield global compressibility, guiding design for desired responses in the long-wavelength limit. Negative compressibility concepts in metamaterials extend natural instabilities but are stabilized through architecture.

References

  1. [1]
    Bulk Elastic Properties - HyperPhysics
    The reciprocal of the bulk modulus is called the compressibility of the substance. The amount of compression of solids and liquids is seen to be very small. The ...Missing: definition | Show results with:definition
  2. [2]
    Isothermal Compressibilities | PNG 520: Phase Behavior of Natural ...
    The isothermal compressibility of a fluid is defined as follows: cf=−1V(∂V∂ρ)T. (18.15)
  3. [3]
    Water Compressibility | U.S. Geological Survey - USGS.gov
    If the water compressed, it wouldn't "push back" out of the straw. Incompressibility is a common property of liquids, but water is especially incompressible.Missing: physics engineering
  4. [4]
    [PDF] Notes on Thermodynamics, Fluid Mechanics, and Gas Dynamics
    (3) The compressibility factor, Z, is defined as,. Z := pv. RT. = p. ρRT . (3.81). If Z ⇡ 1 for a gas, then it can be modeled well with the ideal gas model. The ...
  5. [5]
    Types of Fluid Flows – Introduction to Aerospace Flight Vehicles
    Compressible Versus Incompressible Flows ... The term “compressibility” applied to a fluid means that a fluid can be compressed, squeezing the fluid and bringing ...
  6. [6]
    [PDF] Isothermal vs. adiabatic compressibility
    And in this form we immediately recognize the adiabatic and isothermal compressibilities: γ ≡. Cp. CV. = κT. κS . The argument above came from students in the ...
  7. [7]
    14.1: Compressibility - Chemistry LibreTexts
    Mar 20, 2025 · Compressibility is the measure of how much a given volume of matter decreases when placed under pressure. If we put pressure on a solid or a ...
  8. [8]
    Compressibility - an overview | ScienceDirect Topics
    Compressibility is a measure of the change in volume resulting from the external pressure applied to the surface of an object.
  9. [9]
    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 ...
  10. [10]
    Compressibility: Definition, How It Works, Calculation, and ... - Xometry
    Oct 18, 2023 · Compressibility is a fundamental property that characterizes a material's ability to change in volume when subjected to external forces.
  11. [11]
    Bulk Modulus and Fluid Elasticities - The Engineering ToolBox
    Stainless steel with Bulk Modulus 163×109 Pa is aprox. 80 times harder to compress than water with Bulk Modulus 2.15×109 Pa . Example - Density of Seawater in ...
  12. [12]
    27.5: Compressibility of a Fluid - Physics LibreTexts
    Jul 20, 2022 · The SI unit for bulk modulus is the pascal. If the bulk modulus of a material is very large, a large pressure change will result in only a ...
  13. [13]
    What Exactly is The Compressibility of Fluids? - Mensor Blog
    Compressibility of fluids is the ability to change density with pressure. Liquids are more compressible than solids, and gases are highly compressible.
  14. [14]
    Comparison of static and dynamic measurements of compressibility ...
    The good agreement of the two sets of compressibilities at pressures greater than 2 kb implies that other elastic properties calculated from velocities will be ...
  15. [15]
    SPE/ISRM 47222 Static and Dynamic Compressibility of ... - OnePetro
    As for the mechanical behavior of any elastic body, cavern com- pressibility is apparent both through static and dynamic tests. (i.e., involving waves triggered ...
  16. [16]
    Compressibility - an overview | ScienceDirect Topics
    Compressibility is a measure of the relative volume change of a solid or a fluid in response to a pressure change.
  17. [17]
  18. [18]
  19. [19]
    12.3 Stress, Strain, and Elastic Modulus - UCF Pressbooks
    Compressibility of an object or medium is the reciprocal of its bulk modulus. Shear strain is the deformation of an object or medium under shear stress. The ...
  20. [20]
    Compressibility - Galileo
    In an ideal gas, dE = CVdT and pV = NkBT, so that dH = dE + d(pV) = (CV + NkB)dT and Cp = CV + NkB.Missing: formula | Show results with:formula
  21. [21]
    High Pressure Measurements of Density, Velocity of Sound ... - NIH
    The adiabatic bulk modulus, Bs, is calculated from Bs = ρc2 where ρ is the density and c is the velocity of sound. The low pressure dilatomer [1] was used to ...
  22. [22]
    [PDF] Comparison of dynamic and static bulk moduli of reservoir rocks
    Based on traditional ultrasonic measurement under hydrostatic stress, relation between dynamic and static bulk modulus is investigated. The pore stiffness ...
  23. [23]
    [PDF] Ultrasonic measurement of cylinder expansion at pressures to 40 ...
    The present paper describes how ultrasonic measurements made in solid-filled piston and cylinder devices to 45 kilobars are combined with measurements of the ...
  24. [24]
    Incompressibility of Water | Harvard Natural Sciences Lecture ...
    The bulk modulus of water is about 2.2 x 109 Pa, which means that a change of 1 N/m2 of external pressure on the liquid is able to change a given volume of it ...
  25. [25]
    [PDF] Elastic Modulus = Stress
    Young's Modulus, Shear Modulus,. Bulk Modulus,. Material. E (N/m²). G (N/m²). B (N/m²). Solids. Iron, cast. 100 × 109. 40 × 109. 90 × 109. Steel. 200 × 109. 80 ...
  26. [26]
    [PDF] Elastic Properties of Metals and Alloys. II. Copper
    Oct 29, 2009 · Effects on the elastic constants of tf'mperature, pressure, and mechanical (plastic) deformation are in· cluded. The Cauchy (central·force) ...
  27. [27]
    [PDF] Elasticity of single-crystal periclase at high pressure and temperature
    All of the elastic moduli increased with increasing pressure but decreased with increasing temperature, except the off-diagonal modulus C12, which remained ...
  28. [28]
    [PDF] Temperature Effects on the Universal Equation of State of Solids
    These predictions will Include high temperature pressure-volume curves, thermal expansion curves, and the temperature dependence of the Isothermal bulk modulus ...
  29. [29]
  30. [30]
    [PDF] FORMULAS FOR CALCULATING THE SPEED OF SOUND Revision G
    Jul 13, 2000 · The c term is referred to as the bulk or plate speed of longitudinal waves. The shear modulus can be expressed as. )1(2. E. G ν+. =.
  31. [31]
    Solids and Metals - Speed of Sound - The Engineering ToolBox
    Solids and Metals - Speed of Sound. Solid, Speed of Sound (m/s). Longitudinal, Shear, Extensional. Aluminum, rolled, 6420, 3040, 5000. Antimony, 3400.
  32. [32]
    Speed of sound for understanding metals in extreme environments
    Oct 7, 2024 · Metrology advancements enable the speed of sound to be measured at extreme conditions up to 15 000 K and 600 GPa. This comprehensive review ...
  33. [33]
    Isothermal Compressibility. - an overview | ScienceDirect Topics
    For ideal gas, we can define the compressibility as: (2.76) c g = 1 P. Whereas, for nonideal gas, compressibility is defined as: (2.77) c g = 1 P − 1 Z ( ...
  34. [34]
    [PDF] Isothermal compressibility studies of binary solution of cycloheptane ...
    Isothermal compressibility, Internal pressure are the key parameters in molecular thermodynamics of fluid phase equilibria. These parameters can be deduced ...
  35. [35]
    Isothermal Compressibility - an overview | ScienceDirect Topics
    Isothermal compressibility refers to the measure of the relative volume change of a fluid or solid in response to the application of pressure at constant ...
  36. [36]
    8: Thermodynamics - Physics LibreTexts
    Aug 16, 2020 · The isothermal compressibility: κT=−1V(∂V∂p)T; The isobaric volume coefficient: γp=1V(∂V∂T)p; The adiabatic compressibility: κS=−1V(∂V∂p)S.
  37. [37]
    Calculation of isentropic compressibility and sound velocity in two ...
    Apr 25, 2010 · They are related by βT = βSγ, where γ = CP/CV is the isentropic exponent (the ratio of constant pressure and constant volume specific heats) ...
  38. [38]
    Specific Heat - Cp & Cv | Glenn Research Center - NASA
    Jun 30, 2025 · “Gamma” is just a number whose value depends on the state of the gas. For air, gamma = 1.4 for standard day conditions.
  39. [39]
    5.8: Adiabatic Compressibility - Chemistry LibreTexts
    Jan 15, 2023 · There are small temperature oscillations that occur due to the adiabatic compression followed by expansion of the gas carrying the sound waves.
  40. [40]
    Note: Measurement method for sound velocity of melts in large ...
    Aug 19, 2011 · Many thermophysical parameters may be derived from sound velocity, such as adiabatic bulk modulus, ratio of isobaric to isochoric heat ...
  41. [41]
    Structure and elastic properties of quartz at pressure
    Mar 2, 2017 · Unit cells and crystal structures were determined on a single crystal of quartz at seven pressures from 1 atm to 61.4 kbar.
  42. [42]
    None
    Nothing is retrieved...<|separator|>
  43. [43]
    (PDF) Compressibility of Porous Rocks - ResearchGate
    Aug 10, 2025 · Four compressibilities are defined for porous rock, relating changes in the bulk and pore volumes to changes in the pore and confining pressures.
  44. [44]
    Incorporating self‐consistently calculated mineral physics into ...
    Mar 7, 2009 · In conclusion, this combined approach of mantle convection and self-consistently calculated mineral physics is a powerful and useful technique ...
  45. [45]
    [PDF] High-Pressure Mineral Physics - COMPRES
    The properties of silicate minerals such as perovskite at core-mantle boundary layer conditions have been investigated both experimental- ly and theoretically.Missing: databases | Show results with:databases
  46. [46]
    Seismic Wave Propagation - an overview | ScienceDirect Topics
    For a homogeneous, isotropic medium, the seismic P-wave velocity VP is given by. V P = ( 4 / 3 ) μ + k ρ ,. where μ is the shear modulus, k is the bulk modulus, ...
  47. [47]
    Xenolith constraints on seismic velocities in the upper mantle ...
    Jan 3, 2004 · The isothermal bulk modulus, KT, is related to the adiabatic bulk modulus by the relationship KT = KS (1/(1 + αγT)), where α and γ are the ...
  48. [48]
    Seismic Waves and Earth's Interior
    Temperature tends to lower the speed of seismic waves and pressure tends to increase the speed. Pressure increases with depth in Earth because the weight of ...
  49. [49]
    Single-Crystal Elasticity of MgSiO 3 Bridgmanite to Mid-Lower ...
    Apr 14, 2021 · These elastic parameters are included in a self-consistent thermodynamic model to calculate seismic wave velocities along a lower mantle adiabat ...
  50. [50]
    Seismic Attenuation - an overview | ScienceDirect Topics
    The presence of partial melt in the mantle has the potential to reduce seismic velocity and attenuation, although the magnitude of this effect depends ...
  51. [51]
    [PDF] Effective seismic wave velocities and attenuation in partially ... - HAL
    Jul 20, 2021 · derived compressibility of the rock and melt assemblage depends on the characteristics of the. 295 external wave-induced forcing. Wave ...
  52. [52]
    The Rock Physics behind 4D Seismic | GeoScienceWorld Books
    gas and water, oil and water, and heavy oil and ...
  53. [53]
    Compositional and thermal state of the lower mantle from joint 3D ...
    Compositional and thermal state of the lower mantle from joint 3D inversion with seismic tomography and mineral elasticity ... bulk modulus (24). Then it ...
  54. [54]
    [PDF] Compositional and seismic structure of the mantle
    The last equation is called the Adams-Williamson equation, and describes the change in density purely due to adiabatic self-compression. From a seismic ...
  55. [55]
    Mach Number
    Mach number is object speed divided by the speed of sound, and it is the ratio of an object's speed to the speed of sound in the gas.Missing: criterion | Show results with:criterion
  56. [56]
    [PDF] Compressible Flow Regimes
    Any flow can be considered incompressible if the Mach number is less than about 0.3. ‐ Even gases can flow incompressibly ‐ The incompressible assumption ...Missing: criterion | Show results with:criterion
  57. [57]
    Compressible Flow vs Incompressible Flow in Fluid Mechanics
    Aug 11, 2023 · Compressible flow is a flow that changes in density under pressure, whereas incompressible flow does not. A good indicator is Mach Number.
  58. [58]
    Incompressible vs. Compressible Flow - CFD Land
    This blog explores key concepts including compressible flow regimes, the role of Mach number in identifying flow type, and how fluid properties influence ...
  59. [59]
    [PDF] Conservation of Energy Equation
    For problems involving heat transfer and compressible flows, this equation will permit us to obtain the temperature of the fluid, thereby adding one more ...
  60. [60]
    [PDF] Review Compressible Flow
    • To model/analyze compressible flow. – must include energy conservation equation. (in addition to mass and momentum conservation). Page 3. 3. Compressible Flow ...
  61. [61]
    [PDF] A Polytropic Approximation of Compressible Flow in Pipes with ...
    Subsonic compressible flow is a complex phenomenon, and in order to simplify the required analysis and to allow broadly applicable techniques to be developed, ...
  62. [62]
    Converging/Diverging (CD) Nozzle
    The expansion of a supersonic flow causes the static pressure and temperature to decrease from the throat to the exit, so the amount of the expansion also ...
  63. [63]
    [PDF] Compressible Flow - TME085 - Chapter 12
    Finite Volume Method (FVM). Practical examples. Available CFD codes. Time integration. Numerical schemes. Spatial discretization. Governing equations. Quasi-1D ...
  64. [64]
    [PDF] Finite Volume Solution of the Compressible Boundary-Layer ...
    In the present study we have attempted only to show that a finite volume method is as accurate as corresponding finite difference methods. We have not attempted ...
  65. [65]
    The Tait equation: 100 years on
    The “Tait equation,” which is now widely used to fit liquid density data over wide pressure ranges, is a modification of the original equation of Tait.
  66. [66]
    Tait equation in the extended corresponding states framework
    In this study, we describe a method based on the Tait equation which allows accurate estimation of density and isothermal compressibility of non-polar and ...
  67. [67]
    All About Polytropic Processes | The Physics Teacher - AIP Publishing
    Sep 1, 2022 · Equation (5) looks very much like the familiar equation pVγ = constant for an adiabatic process, where γ = CP/CV is the specific heat ratio.
  68. [68]
    A New Two-Constant Equation of State | Industrial & Engineering ...
    ... Peng–Robinson Equation of State. Industrial & Engineering Chemistry Research 2024, 63 (24) , 10795-10811. https://doi.org/10.1021/acs.iecr.4c01076. Hang Yang ...
  69. [69]
    [PDF] The international thermodynamic equation of seawater – 2010:
    Mar 20, 2017 · ... Isothermal compressibility ..…….………………………………………………... 21. 2.16 ... (Sea Water) library which is already in wide circulation. Both the.
  70. [70]
    The isothermal compressibility of seawater near one atmosphere
    The isothermal compressibility of artificial seawater has been measured by a piezometric technique from ‰ ‰ intervals and from 0 to 40°C in 5° intervals.Missing: ocean depths
  71. [71]
    Complete equation of state for shocked liquid nitrogen - AIP Publishing
    Aug 2, 2016 · These calculations provide the first comparison of EOS developments with recently measured P-T states under multiple shock compression. The ...
  72. [72]
    Research in Supersonic Flight and the Breaking of the Sound Barrier
    The large drag rise due to compressibility effects normally occurs at a freestream Mach number slightly above the critical Mach number; this is called the ...
  73. [73]
    Transonic Aircraft
    Drag increases dramatically as an aircraft approaches Mach 1, so airliners use high thrust gas turbine propulsion systems. On the slide we show a DC-8 airliner ...Missing: rise | Show results with:rise
  74. [74]
    [PDF] Linear Subsonic Flow - MIT OpenCourseWare
    This is the Prandtl-Glauert rule. It is a similarity rule that relates incompressible flow over a given two-dimensional profile to subsonic compressible ...
  75. [75]
    [PDF] sak12e5.tmp - NASA Technical Reports Server (NTRS)
    The principal assumptions of the Prandtl-Glauert method for subsonic compressible flow are as follows: (1) The influence of the viscosity and conductivity. of ...
  76. [76]
    The Whitcomb Area Rule: NACA Aerodynamics Research ... - NASA
    The area rule revolutionized how engineers looked at high-speed drag and impacted the design of virtually every transonic and supersonic aircraft ever built.
  77. [77]
    [PDF] advanced transonic aerodynamic technology
    The wave itself causes some increase in drag, but usually the principal effect is separation of the boundary layer with a significant increase in drag, ...Missing: compressibility | Show results with:compressibility
  78. [78]
    [PDF] HIGH-SPEED AERODYNAMIC
    This compression of the air caused the formation of shock waves and substantial increases in drag. The increase in transonic drag seemed insurmountable for ...
  79. [79]
    [PDF] Lattice Boltzmann and Navier-Stokes Cartesian CFD Approaches for ...
    Lattice Boltzmann (LB) and Navier-Stokes (NS) CFD approaches are compared for simulating airframe noise, using the same Cartesian mesh and focusing on high- ...
  80. [80]
  81. [81]
    Normal Shock Wave Equations
    On this slide we have listed the equations which describe the change in flow variables for flow across a normal shock. The equations presented here were derived ...
  82. [82]
    [PDF] Shock Waves
    Figure 7: Graphs of p2/p1, T2/T1 and ρ2/ρ1 against inlet Mach number, M1. In addition, it is important to evaluate the entropy change across the shock which can ...
  83. [83]
    [PDF] AA210A Fundamentals of Compressible Flow
    Oct 25, 2020 · For most flow situations (outside of shock waves) the stress and heat conduction ... 9.3.1 The Rankine Hugoniot relations. The jump conditions can ...
  84. [84]
    [PDF] Equations, Tables and Charts for Compressible Flow
    Many relations for normal shock waves are conveniently expressed in ... The Rankine-Hugoniot relations are the same as for normal shock waves: www. 1.
  85. [85]
    Oblique Shock Waves
    The Mach number and speed of the flow also decrease across a shock wave. If the shock wave is perpendicular to the flow direction, it is called a normal shock.
  86. [86]
    Shock reflection and oblique shock waves - AIP Publishing
    Dec 27, 2007 · As a shock front hits a planar wall with an incident angle α ⁠, an oblique reflected shock wave is produced. For small incident angle α ⁠, the ...
  87. [87]
    [PDF] 4.3 Hydrodynamical shocks
    T2. T1. = [2M. 2. 1. (. 1)] [2 + (. 1)M2. 1]. ( + 1)2M2. 1. ,. (377) and hence if M1 > 1, the post-shock temperature is higher than the pre-shock temper- ature, ...
  88. [88]
  89. [89]
    Blast Wave - an overview | ScienceDirect Topics
    A blast wave is defined as a shock wave that propagates through an explosive mixture at supersonic velocity, resulting from the almost instantaneous increase ...
  90. [90]
    Schlieren Flow Visualization
    Aerodynamicists use wind tunnels to test models of proposed aircraft and engine components. During a test, the model is placed in the test section of the ...
  91. [91]
    Negative linear compressibility in common materials - AIP Publishing
    Jun 10, 2015 · Negative linear compressibility occurs when a material behaves contrary to expectation by conventional elasticity theory under the application of hydrostatic ...
  92. [92]
    Evidence of Pressure-Induced Phase Transitions and Negative ...
    Nov 22, 2021 · ... Transitions and Negative Linear Compressibility in Formamidinium Manganese-Hypophosphite Hybrid Perovskite ... Transitions and Electronic ...
  93. [93]
  94. [94]
    Giant Negative Compressibility in 2D Hybrid Perovskites via ...
    Jun 22, 2025 · A 2D perovskite constructed by inorganic -[PbBr 6 ]- connected by C chains, displays an abnormal volume expansion ratio of 12.9% under external compression.
  95. [95]
    Negative compressibility and global instability in two-phase systems
    Apr 2, 2018 · Negative compressibility of a liquid with an inclusion of solid seed is predicted theoretically and verified experimentally with helium-4 ...
  96. [96]
    Experiments on Metamaterials with Negative Effective Static ...
    Dec 8, 2017 · We demonstrate experimentally, for what we believe is the first time, a metamaterial with a negative effective compressibility under quasistatic conditions.
  97. [97]
    Auxetic mechanical metamaterials - RSC Publishing
    Jan 17, 2017 · Auxetic mechanical metamaterials are identified by a negative Poisson's ratio and are perhaps the most widely studied type of mechanical metamaterials.
  98. [98]
    Designing orthotropic materials for negative or zero compressibility
    In this paper we propose a general design technique for finding metamaterials with negative or zero compressibility by using a topology optimization approach.
  99. [99]
    3D printing of active mechanical metamaterials: A critical review
    Today's 3D-printed mechanical metamaterials exhibit impressive properties such as high strength density ratio, negative compressibility, negative or zero ...
  100. [100]
    Micro-Structured Two-Component 3D Metamaterials with Negative ...
    Jan 12, 2017 · We fabricate three-dimensional (3D) two-component polymer micro-lattices by using gray-tone laser lithography. We perform cross-correlation analysis of optical ...
  101. [101]
    Metamaterials with negative compressibility highlight evolving ...
    Oct 3, 2024 · Negative compressibility, as discussed above, is a longitudinal effect: the volume, area, or length deformation is measured along the dimensions ...
  102. [102]
    Mechanical Metamaterials: Auxetic Metamaterials - IntechOpen
    The excellent mechanical properties of auxetic metamaterials are expected to overcome some common problems in aerospace, naval, biomedical, and railway ...1. Introduction · Figure 2 · 4. New Auxetic Metamaterials
  103. [103]
    Auxetic 3D printed metastructure stents for enhanced mechanical ...
    This study introduces a hybrid auxetic stents designed for vascular applications, a significant advancement over traditional stent materials.Missing: shock | Show results with:shock