Fact-checked by Grok 2 weeks ago

Bulk modulus

The bulk modulus, often denoted as B or K, is a fundamental mechanical property of a substance that quantifies its resistance to uniform (hydrostatic) under applied . It is defined as the of the infinitesimal increase in to the corresponding relative decrease in , mathematically expressed as B = -\frac{dP}{dV/V} = -V \left( \frac{\partial P}{\partial V} \right), where V is the and P is the ; the negative sign ensures a positive value since reduces . This modulus has units of , typically measured in pascals (Pa) or gigapascals (GPa), reflecting its equivalence to over in volumetric terms. As one of the primary elastic constants—alongside (for uniaxial tension), (for distortion), and (for lateral contraction)—the bulk modulus specifically addresses volumetric elasticity under isotropic stress, distinguishing it from directional deformations. It is particularly important in for evaluating in high-performance alloys, ceramics, and composites, where high values indicate strong interatomic bonding and resistance to deformation under extreme conditions. In and , the bulk modulus informs models of Earth's interior and speeds, as it correlates with a material's and molecular interactions, with softer volatiles exhibiting lower values compared to rigid minerals. The bulk modulus also finds applications in and , such as predicting the behavior of liquids under in hydraulic systems, where values for are around 2.2 GPa at , far lower than metals like (approximately 160 GPa), highlighting the relative incompressibility of solids. Its often involves hydrostatic experiments, and relations to other moduli, such as B = \frac{E}{3(1-2\nu)} (where E is and \nu is ), enable comprehensive of isotropic materials.

Fundamentals

Definition and Basic Properties

The bulk modulus, often denoted as K, quantifies a material's to under hydrostatic . It is mathematically defined as the negative ratio of an infinitesimal increase in to the corresponding relative decrease in , evaluated at constant temperature: K = -V \left( \frac{\partial P}{\partial V} \right)_T Here, V represents the , P the , and the subscript T indicates the isothermal condition. The negative sign accounts for the fact that increasing causes , ensuring K remains positive. The units of the bulk modulus match those of , commonly expressed in gigapascals (GPa) in the system. This property specifically measures the response to isotropic , where equal acts from all directions, resulting in pure volumetric without deformation or shape change. In contrast, uniaxial stresses induce both longitudinal and lateral deformations, as captured by other moduli. High values of K indicate low , meaning substantial is required to achieve even a small fractional volume change, which is particularly relevant for fluids and solids under high-pressure conditions. The concept of volume elasticity, akin to the modern bulk modulus, emerged in the through experimental studies on gas and , with notable measurements conducted by Henri Victor Regnault around 1860 that informed early understandings of pressure-volume relations. The specific term "bulk modulus" first appeared in in the early . For isotropic materials like fluids, K is a scalar, uniformly describing resistance to in all directions due to the lack of preferred orientations. In anisotropic solids, elastic behavior is governed by a fourth-rank tensor, but under hydrostatic loading, an effective bulk modulus can be derived as the of the sum of the elements of the compliance matrix corresponding to normal stresses, specifically K = [S_{11} + S_{22} + S_{33} + 2(S_{12} + S_{13} + S_{23})]^{-1} for materials with orthotropic symmetry, focusing on the average volumetric response while acknowledging directional variations.

Relation to Other Elastic Moduli

In isotropic materials, the bulk modulus K is related to the Lamé constants [\lambda](/page/Lambda) (the first Lamé parameter) and [\mu](/page/MU) (the , also known as the second Lamé parameter) by the expression K = \lambda + \frac{2}{3} \mu. This relation arises from the general form of the elastic stress-strain tensor for isotropic media, where [\lambda](/page/Lambda) governs volumetric response and [\mu](/page/MU) contributes to the resistance against uniform compression. The bulk modulus connects to Young's modulus E and Poisson's ratio \nu through K = \frac{E}{3(1 - 2\nu)}. This formula derives from considerations of strain energy or, equivalently, by superposing uniaxial strains to achieve hydrostatic stress, yielding the volumetric strain \Delta V / V = -P / K, where P is pressure, and expressing it in terms of E and the lateral contraction captured by \nu. Similarly, K relates to Young's modulus E and the shear modulus G (equivalent to \mu) via K = \frac{E G}{3(3G - E)}. This expression follows by substituting the relation G = \frac{E}{2(1 + \nu)} into the formula for K in terms of E and \nu, highlighting how shear resistance influences . In the limit of incompressibility, where \nu \to 0.5, K \to \infty, as E \to 3G, rendering volume changes negligible under pressure. These relations extend to wave : in fluids, lacking , the P-wave speed is c_p = \sqrt{K_s / \rho}, where K_s is the adiabatic bulk modulus and \rho is , whereas in , it incorporates effects as c_p = \sqrt{(K_s + \frac{4}{3} G) / \rho}.

Thermodynamic Considerations

Isothermal and Adiabatic Forms

The bulk modulus can be defined under different thermodynamic constraints, leading to distinct isothermal and adiabatic forms that reflect the role of exchange during or processes. The isothermal bulk modulus, denoted K_T, characterizes the material's response to volume changes at constant and is given by K_T = -V \left( \frac{\partial P}{\partial V} \right)_T, where V is the and P is the . This form applies to static or quasi-static processes where is maintained constant, allowing with the surroundings. In contrast, the adiabatic bulk modulus, K_S, describes volume changes under constant conditions, with no exchange, and is expressed as K_S = -V \left( \frac{\partial P}{\partial V} \right)_S. This is particularly relevant for rapid, dynamic processes such as the propagation of sound waves, where the occurs too quickly for significant thermal equilibration. The two moduli are related through the \gamma = C_P / C_V, where C_P and C_V are the heat capacities at constant and volume, respectively, yielding \frac{K_S}{K_T} = \gamma. Since \gamma > 1 for most materials due to C_P > C_V, it follows that K_S > K_T, indicating greater resistance to under adiabatic conditions. In limiting cases, such as when \gamma = 1, the isothermal and adiabatic moduli become equal; this occurs, for example, in gases at very low frequencies where processes approach isothermal despite being nominally adiabatic.

Derivation from Thermodynamic Potentials

The isothermal bulk modulus K_T can be derived from the F(T, V), which is the appropriate for constant and . The pressure is given by the relation P = -\left( \frac{\partial F}{\partial V} \right)_T. Differentiating with respect to at constant yields \left( \frac{\partial P}{\partial V} \right)_T = -\left( \frac{\partial^2 F}{\partial V^2} \right)_T. Since the isothermal bulk modulus is defined as K_T = -V \left( \frac{\partial P}{\partial V} \right)_T, it follows that K_T = V \left( \frac{\partial^2 F}{\partial V^2} \right)_T. This represents the curvature of the with respect to , quantifying the material's resistance to uniform at constant . Maxwell relations from the thermodynamic potentials further connect these derivatives. Specifically, from the differential form dF = -S \, dT - P \, dV, the equality of mixed partial derivatives implies relations that link pressure-volume behavior to and . Applying this to the pressure derivative gives the consistent form \left( \frac{\partial P}{\partial V} \right)_T = -\left( \frac{\partial^2 F}{\partial V^2} \right)_T, reinforcing the expression for K_T. This derivation ties directly to the U(S, V), as F = U - TS, where the second derivative of U at constant contributes to adiabatic stiffness, though the isothermal case emphasizes . Similarly, the adiabatic bulk modulus K_S can be derived from the H(S, P), where dH = T \, dS + V \, dP. The volume is given by V = \left( \frac{\partial H}{\partial P} \right)_S, and \left( \frac{\partial V}{\partial P} \right)_S = \left( \frac{\partial^2 H}{\partial P^2} \right)_S, yielding K_S = -V / \left( \frac{\partial^2 H}{\partial P^2} \right)_S. The \gamma, which bridges thermal and elastic properties, emerges from these potentials as \gamma = \frac{V \alpha K_T}{C_V}, where \alpha = \frac{1}{V} \left( \frac{\partial V}{\partial T} \right)_P is the coefficient and C_V is the at constant volume. This relation, derived using Maxwell identities from F or G, quantifies anharmonic effects in the volume dependence of vibrational frequencies, connecting the second derivatives of free energies to thermal expansion.

Measurement Techniques

Experimental Determination

The experimental determination of the bulk modulus involves techniques that measure a material's resistance to uniform compression, typically under hydrostatic conditions to ensure isotropic pressure application. Static methods operate at low frequencies and yield the isothermal bulk modulus, while dynamic methods probe high-frequency responses to obtain the adiabatic bulk modulus, which relates to the via c = \sqrt{K_S / \rho}, where K_S is the adiabatic bulk modulus and \rho is the . Static methods for solids commonly employ a piston-cylinder apparatus, where the sample is subjected to hydrostatic pressure and the resulting volume change is measured to compute the bulk modulus as K = -V \frac{\Delta P}{\Delta V}. This setup confines the sample in a cylindrical chamber with a pressure-transmitting medium, allowing pressures up to 2–5 GPa, or higher in specialized designs. Such measurements are particularly useful for ductile materials like metals or polymers, where direct volumetric compression can be tracked using displacement sensors or diffraction for lattice parameter changes. Dynamic methods, such as ultrasonic techniques, utilize the pulse-echo method to determine sound velocities in the . In this approach, ultrasonic pulses are transmitted through the sample, and the time-of-flight between echoes is used to calculate the longitudinal speed c_L and shear speed c_T, from which the adiabatic bulk modulus is derived as K_S = \rho (c_L^2 - \frac{4}{3} c_T^2) after measuring the \rho. This technique is non-destructive and applicable to a wide range of solids, including polycrystalline materials, with accuracies typically better than 1% for well-characterized samples. For transparent materials, Brillouin scattering provides a contactless dynamic measurement of the high-frequency adiabatic bulk modulus by analyzing the of light from acoustic s. The frequency shift of the scattered light yields phonon velocities in various directions, enabling derivation of the full tensor and thus K_S, often under extreme conditions like in diamond anvil cells. This method excels in single crystals or glasses where optical access is feasible, offering resolution down to micrometer scales. Key challenges in these experiments include maintaining hydrostaticity, especially in , where non-uniform stresses can distort measurements; this is addressed by using pressure-transmitting fluids like , which remain hydrostatic up to about 7 GPa. Additionally, sample introduces errors by contributing compliant void spaces that close under pressure, leading to an underestimation of the effective bulk modulus at low pressures; corrections often involve accounting for crack or using dry, low-porosity samples.

Computational and Theoretical Methods

Computational and theoretical methods provide predictive tools for determining the bulk modulus of materials, particularly when experimental data are scarce or under extreme conditions. These approaches range from semi-empirical equation-of-state (EOS) fittings based on thermodynamic assumptions to fully quantum-mechanical simulations and classical molecular dynamics. Such methods enable the exploration of bulk modulus variations with pressure, temperature, and composition, often achieving accuracies comparable to experiments for well-characterized systems. One foundational technique involves fitting equations of state to energy-volume data obtained from simulations or experiments, allowing extraction of the bulk modulus and its pressure derivative. The Birch-Murnaghan EOS, a third-order finite-strain model, is widely used for high-pressure predictions in solids, assuming a dependence of the bulk modulus on . It is expressed as: P(V) = \frac{3K_0}{2} \left[ \left( \frac{V_0}{V} \right)^{7/3} - \left( \frac{V_0}{V} \right)^{5/3} \right] \left\{ 1 + \frac{3}{4} (K_0' - 4) \left[ \left( \frac{V_0}{V} \right)^{2/3} - 1 \right] \right\} where P is pressure, V is volume, V_0 is the reference volume, K_0 is the zero-pressure bulk modulus, and K_0' is its pressure derivative. This EOS, derived from Eulerian finite strain theory, accurately describes compression in cubic crystals and minerals up to gigapascal pressures, with K_0 obtained by minimizing the fit to computed or measured P-V points. Density functional theory (DFT) offers an ab initio approach to compute the bulk modulus by evaluating the stress-strain response in periodic crystal structures. In DFT, the total energy is minimized as a functional of the electron density, yielding ground-state properties including equilibrium volume and elastic constants from finite distortions or analytical derivatives of the energy with respect to volume. For instance, the isothermal bulk modulus K_T is derived from the second derivative of the energy per unit volume at the equilibrium lattice constant: K_T = V \frac{\partial^2 E}{\partial V^2}\big|_{V=V_0}. Software packages such as VASP and Quantum ESPRESSO facilitate these calculations using plane-wave basis sets and pseudopotentials, with typical accuracies within 5-10% of experimental values for semiconductors and metals when employing generalized gradient approximation (GGA) functionals. These methods are particularly valuable for novel materials where interatomic interactions must be quantum-mechanically resolved. Molecular dynamics (MD) simulations predict the bulk modulus through statistical mechanics in the isothermal-isobaric (NPT) ensemble, where volume fluctuations under constant pressure and temperature directly relate to compressibility. The isothermal bulk modulus is extracted from the variance of volume fluctuations via the fluctuation-dissipation theorem: K_T = \frac{\langle V \rangle k_B T}{\langle (\Delta V)^2 \rangle}, where \langle V \rangle is the average volume, k_B is Boltzmann's constant, T is temperature, and \langle (\Delta V)^2 \rangle is the mean-squared volume deviation. This approach, implemented in codes like LAMMPS or GROMACS, captures thermal effects and anharmonicities, providing K_T values for liquids, polymers, and nanomaterials with statistical precision improving with simulation length, though it requires reliable force fields for accuracy. Empirical potentials, such as the Lennard-Jones (LJ) model, approximate interatomic interactions for efficient large-scale simulations of bulk modulus in simple systems like or molecular crystals. The LJ potential, V(r) = 4\epsilon \left[ \left( \frac{\sigma}{r} \right)^{12} - \left( \frac{\sigma}{r} \right)^6 \right], yields the bulk modulus from the pressure-volume relation in MD or runs, often matching experimental values for van der Waals solids within 20%. However, its pairwise additive form neglects many-body effects and directional bonding, leading to significant inaccuracies—up to 50% errors—for metals where embedded-atom or tight-binding potentials are more appropriate. These limitations highlight the need for potential selection based on material type, linking back to microscopic interatomic models.

Material Values and Applications

Selected Bulk Moduli for Common Materials

The bulk modulus varies widely among materials, spanning orders of depending on , bonding type, and . Gases exhibit extremely low values owing to their sparse molecular arrangement, liquids intermediate values reflecting moderate intermolecular forces, and solids high values due to strong atomic interactions. These differences highlight the material's resistance to , with values typically measured under ambient conditions unless noted. Representative examples are summarized in the table below, drawn from established thermophysical data compilations.
Material CategoryMaterialBulk Modulus (GPa)ConditionsSource
GasesAir0.00014STP, adiabatic
LiquidsWater2.220°C
Solids (metals)Steel160Room temperature
Solids (covalent)Diamond442Room temperature
Across metals, bulk moduli generally increase with cohesive energy per unit volume, which can correlate with within similar groups, as higher mass often accompanies denser packing and stronger . In crystalline solids, arises from directional bonding, leading to variations in along different axes; for polycrystals, effective isotropic values are bounded by the Voigt (upper) and Reuss (lower) averages to account for aggregate behavior. Under elevated , the bulk modulus typically strengthens nonlinearly, but a common is K(P) = K_0 + K_0' P, where K_0 is the ambient bulk modulus and K_0' (the pressure derivative) averages approximately 4 for many , indicating progressive stiffening with compression. Temperature generally softens materials through anharmonic vibrations and , causing a decrease in bulk modulus; for metals, this often amounts to roughly a 10% reduction between 300 K and 1000 K, as observed in elastic constant measurements.

Role in Engineering and Natural Phenomena

In engineering applications, the bulk modulus is essential for designing , where it quantifies a fluid's to and directly influences , , and . Hydraulic fluids like oils exhibit high bulk moduli, typically around 1.5–2 GPa, which minimize volumetric changes under , ensuring precise and rapid transmission of in actuators and machinery. This near-incompressibility reduces energy losses and vibrations, critical for such as excavators and aircraft . In pressure vessel design, engineers rely on the to predict material deformation under internal pressure using the relation \frac{\Delta V}{V} = -\frac{\Delta P}{K}, where \Delta V / V is the relative volume change, \Delta P is the pressure increment, and K is the . This formula allows assessment of containment integrity for gases or liquids in tanks and pipelines, guiding material selection to prevent rupture; for instance, vessels with K \approx 160 GPa maintain minimal expansion under high pressures up to several hundred . Low materials would lead to excessive swelling, compromising safety in applications like chemical processing or fuel storage. Geophysicists use the bulk modulus to model Earth's interior structure and dynamics, particularly in the where the inner core has an estimated bulk modulus of about 1340 GPa at the , consistent with properties of iron under those extreme conditions, enabling rapid propagation through the high-pressure environment. This property helps interpret seismic data to infer core and state, as variations in K affect P-wave velocities observed in global recordings. In mantle convection studies, the bulk modulus integrates with and thermal expansivity to simulate material flow and heat transfer, reproducing observed seismic anomalies like low-velocity zones that drive . The bulk modulus governs acoustic phenomena in natural fluids, such as the in , where K \approx 2.3 GPa contributes to velocities around 1500 m/s, optimizing signal propagation for underwater navigation and detection. This compressibility influences wave refraction in ocean layers, affecting long-range communication in marine environments. Additionally, in processes, the bulk modulus sets thresholds for bubble formation under negative acoustic pressures; fluids with higher K require greater intensity to initiate collapse, mitigating damage in propellers and pumps while informing models of natural phenomena like oceanic bubble dynamics. In biological contexts, the effective bulk modulus of compliant structures like vessels is notably low, around 0.01 GPa, enabling elastic expansion to buffer pulsatile flow and regulate without excessive wall . This low K facilitates and , essential for cardiovascular and adapting to physiological demands such as exercise. In ultrasound imaging of tissues, bulk modulus variations underpin techniques, where differences in compressibility between healthy and pathological soft tissues—such as tumors with elevated K—allow non-invasive stiffness mapping to aid in organs like the liver or .

Microscopic Foundations

Interatomic Interactions

The bulk modulus emerges from the microscopic interactions between atoms, particularly through the curvature of the interatomic potential energy landscape. In models using central pair potentials U(r), where r is the interatomic distance, the bulk modulus K at equilibrium is proportional to the combination of the second derivative of the potential and a term involving the first derivative, specifically K \propto \frac{1}{r} \frac{d^2 U}{dr^2} + \frac{1}{r} \frac{dU}{dr} evaluated at the equilibrium distance r_0. This expression arises because uniform compression changes the volume by scaling all distances equally, linking the energy change to these derivatives in a lattice sum over pair interactions. Such pair potential models, like the Lennard-Jones potential, provide a foundational understanding of how short-range repulsive and long-range attractive forces contribute to resistance against volumetric strain. The magnitude of the bulk modulus varies significantly with the type of atomic bonding, reflecting the strength and nature of interatomic forces. Covalent bonds, characterized by strong, directional sharing of electrons, yield high bulk moduli; for instance, exhibits a bulk modulus of approximately 442 GPa due to its rigid tetrahedral network. Metallic bonds, involving delocalized electrons, result in intermediate values, typically 50–200 GPa for common metals like (140 GPa), as the bonding is less directional but still cohesive. Ionic bonds, formed by electrostatic attractions between charged ions, produce variable bulk moduli (around 20–100 GPa, e.g., 24 GPa for NaCl) that depend on and ion size, with higher coordination often enhancing stiffness. Beyond the harmonic approximation, anharmonicity in the interatomic potential introduces nonlinear effects that cause the bulk modulus to depend on pressure and temperature. Higher pressures steepen the potential well, generally increasing K, but in some materials, anharmonic coupling leads to softening under compression due to mode Grüneisen parameters exceeding unity. Thermal softening occurs as anharmonic vibrations allow atoms to explore higher-energy regions of the potential, reducing the effective curvature and thus K with rising temperature; this is evident in materials like silicates where phonon anharmonicity correlates with negative pressure dependence. To account for temperature effects without full anharmonic treatment, the quasi-harmonic approximation treats vibrational frequencies as volume-dependent, enabling prediction of and modulus variation. In this framework, the bulk modulus decreases with because volume expansion shifts the equilibrium to a softer part of the potential landscape, as seen in calculations for metals like where quasi-harmonic contributions match experimental softening up to several hundred . This approximation captures essential thermodynamic links, such as the relation between and modulus via Grüneisen parameters, without invoking explicit higher-order .

Connection to Linear Elasticity

The generalized Hooke's law in three dimensions expresses the linear relationship between the components of the stress tensor \sigma_{ij} and the infinitesimal strain tensor \epsilon_{kl} via the fourth-rank elastic stiffness tensor C_{ijkl}: \sigma_{ij} = C_{ijkl} \epsilon_{kl}. This formulation extends the one-dimensional spring constant to anisotropic solids, where the 81 components of C_{ijkl} reduce to 21 independent ones due to symmetries in stress and strain. For crystals with cubic symmetry, only three independent constants remain: C_{11}, C_{12}, and C_{44}. Under hydrostatic pressure, which induces uniform volumetric strain \epsilon_{kk} = 3\epsilon (where \epsilon is the linear strain), the bulk modulus K—defined as the ratio of hydrostatic stress to volumetric strain—emerges as K = \frac{C_{11} + 2C_{12}}{3}. This expression captures the resistance to uniform compression without involving shear (C_{44}), as shear deformations do not contribute to volume change in the linear regime. In the isotropic limit, applicable to polycrystalline aggregates or amorphous solids, the stiffness tensor simplifies further, depending on just two parameters: the first Lamé constant \lambda (relating to lateral strain effects) and the shear modulus \mu. The bulk modulus then takes the form K = \lambda + \frac{2}{3} \mu, derived by taking the trace of the stress-strain relation under hydrostatic conditions, where the deviatoric (shape-changing) components vanish, isolating the volumetric response. This relation highlights how K governs the material's incompressibility, with \mu contributing to overall stiffness but weighted less in pure dilation. Mechanical stability in linear elasticity demands that the stiffness tensor be positive definite, ensuring strains increase energy under applied stress. A key condition is K > 0, which prevents catastrophic collapse under compression and forms part of the Born stability criteria for lattice dynamical stability. For cubic crystals, this translates to C_{11} + 2C_{12} > 0 (or equivalently $3K > 0), alongside shear stability conditions like C_{44} > 0 and C_{11} > |C_{12}|; violations signal phase transitions or instabilities. Real materials deviate from ideal due to defects, which introduce local softening. Vacancies reduce atomic coordination, weakening interatomic bonds and decreasing K by 1-10% at typical concentrations (e.g., ~0.1-1% in annealed metals or semiconductors); in , a monovacancy induces ~0.34% softening, scaling to ~6% reduction at 1% vacancy fraction via linear response. Dislocations similarly localize , contributing comparable reductions through core distortions, though their effects are often screened in dense networks. These perturbations arise from microscopic interactions but manifest in the effective constants.

References

  1. [1]
    [PDF] Lecture 7 - Elasticity
    • Bulk modulus B = - V dP/dV = V d2E/dV2 (same. units as pressure ) • Compressibility K = 1/B. Physics 460 F 2006 Lect 7. 6.
  2. [2]
    12.3 Stress, Strain, and Elastic Modulus - UCF Pressbooks
    Bulk strain is the response of an object or medium to bulk stress. Here, the elastic modulus is called the bulk modulus. Bulk stress causes a change in the ...
  3. [3]
    [PDF] Predictive analytics for crystalline materials: bulk modulus - cucis
    Sep 27, 2016 · The bulk modulus is one of the important parameters for designing advanced high-performance and thermoelectric materials.
  4. [4]
    [PDF] 3 Fundamentals of Planetary Materials - CalTech GPS
    The bulk modulus is a measure of the strength of interaction among the molecules in the material—soft (weakly- bound/volatile) materials compress easily, while ...
  5. [5]
    Fluid Properties - Copyright
    Hence, the value of K for liquids is typically large. For example, the bulk modulus for water at 20 C is K=2.24 GPa. For mercury at 20 C, K=28.5 GPa.
  6. [6]
    [PDF] 2. Definitions and Methods - Stanford University
    For example, the bulk modulus is measured in the hydrostatic compression experiment. The shear modulus is measured in the shear deformation experiment. Stress ...
  7. [7]
    Equations of State (EOS) - SERC (Carleton)
    Jan 19, 2012 · The bulk modulus is a measure of the ability of a material to withstand changes in volume when under uniform compression. This equation is only ...<|control11|><|separator|>
  8. [8]
    Bulk Modulus - an overview | ScienceDirect Topics
    Bulk modulus is defined as the ratio of hydrostatic pressure to volume strain, indicating a material's resistance to uniform compression.
  9. [9]
    Bulk Modulus and Fluid Elasticities - The Engineering ToolBox
    Bulk Modulus is a fluid's compressibility, how easily a unit volume changes with pressure. It's the ratio of pressure change to volume change per unit volume.
  10. [10]
    [PDF] Heat and Thermodynamics - LibreTexts
    In that case we would find that the adiabatic bulk modulus is less than the isothermal bulk modulus – but that ... Regnault's experiments of around 1860.
  11. [11]
    BULK MODULUS Definition & Meaning - Dictionary.com
    Bulk modulus definition: a coefficient of elasticity of a substance ... Word History and Origins. Origin of bulk modulus. First recorded in 1905–10.Missing: earliest mention
  12. [12]
    [PDF] Anisotropic Elasticity
    Mar 15, 2020 · Even in cubic materials, 3 numbers/coefficients/moduli are required to describe elastic properties; isotropic materials only require 2. • ...
  13. [13]
    [PDF] Relation between elastic constants
    Feb 13, 2020 · The constants referred to below are. λ – Lamé coefficient, has no name. µ – Lamé coefficient, shear modulus ν – Poisson's ratio. E – Young's ...
  14. [14]
    Relation between Young's modulus and shear modulus - DoITPoMS
    We obtained the relation between Young's modulus \( E \) and shear modulus \( G \). We can also relate the bulk modulus \( K \) to Young's modulus \( E \).
  15. [15]
    [PDF] 332: Mechanical Behavior of Materials
    3) Bulk Modulus for an Isotropic Material. The bulk modulus, Kb, of a ... 3(3G−E). E. 2G. − 1. G(4G−E). 3G−E. E, ν. E. 3(1−2ν). E. 2(1+ν). E(1−ν). (1+ν)(1−2ν ...
  16. [16]
    [PDF] Thermodynamics and Equations of State
    Isothermal bulk modulus. Adiabatic bulk modulus. The pressure is obtained by differentiation: Р. = 'au av T. The isothermal bulk modulus, KT, is. K₁ = -V(ƏPƏV)T.
  17. [17]
    [PDF] Change in the vibrational properties of bulk metal glass with time
    Jun 15, 2006 · To elaborate, the adiabatic-to- isothermal bulk modulus ratio, Ks /KT =Cp /Cv, where Cp is the heat capacity at constant pressure and Cv is ...
  18. [18]
    Bulk Modulus
    The two most important bulk moduli are the isothermal bulk modulus, ... Note that the isentropic bulk modulus of an ideal gas is greater than its isothermal bulk ...
  19. [19]
    Causal-Constraint Broadband Sound Absorption under Isothermal ...
    Jun 10, 2025 · In the high-frequency limit, the bulk modulus approaches the adiabatic value, B 0 = γ P , where γ is the specific heat ratio. This transition ...
  20. [20]
    [PDF] Pressure—Volume—Temperature Equation of State
    Isothermal bulk modulus. Thermal expansion parameter. Grüneisen parameter. = V ... Isothermal bulk modulus. Thermal expansion parameter. KT(V, T) = ✓. P. T.
  21. [21]
    [PDF] PDF - School of Physics and Astronomy
    Using H in place of U gives Cp > 0, and using F gives positive isothermal bulk modulus (KT ). ... Gibbs-Helmholtz relation for calculating free energy and entropy ...
  22. [22]
    On the importance of the free energy for elasticity under pressure
    ... Gibbs free energy G, rather than the energy E. ... Thus, the bulk modulus K V is the second derivative ... second derivative of energy ψ, rather than of G.
  23. [23]
    Elastic properties of studied with an ultrasonic pulse-echo technique
    Nov 6, 2006 · The sound speeds were measured by the ultrasonic pulse-echo overlap technique ... Note that the value of the bulk modulus determined by the ...
  24. [24]
    Direct measurement and calculation of rubber bulk modulus by ...
    The method uses a piston-cylinder apparatus to measure rubber bulk modulus by applying load to a sample and water, measuring displacement, and calculating the  ...
  25. [25]
    [PDF] The pressure-volume relation of ytterbium up to 9 GPa
    Further, these measurements were made with piston-cylinder apparatus which had ~ 4 GPa as the upper limit of attainable pressure. As a result, no measurements.<|separator|>
  26. [26]
    Brillouin Scattering and its Application in Geosciences
    Jan 1, 2014 · 2000). Surface Brillouin scattering is in general not observable in transparent materials, where bulk effects dominate the scattered intensity.
  27. [27]
    Experimental Investigation on Static and Dynamic Bulk Moduli of Dry ...
    Sep 29, 2020 · We examined experimentally the evolution of static and dynamic bulk moduli for porous Bentheim sandstone with increasing confining pressure up to about 190 MPa.<|control11|><|separator|>
  28. [28]
    Density functional theory predictions of the mechanical properties of ...
    Jul 2, 2021 · Here we introduce the concepts behind DFT, and highlight a number of case studies and methodologies for predicting the elastic constants of materials.
  29. [29]
    [PDF] Revisting Lennard Jones, Morse, and N-M potentials for metals
    Feb 2, 2022 · The. Lennard Jones N-M/Mie potential is an excellent example in that it cannot physically model materials with small ratios of bulk modulus to.
  30. [30]
    Metals and Alloys - Bulk Modulus Elasticity - The Engineering ToolBox
    Stainless steel with Bulk Modulus 163×109 Pa is approximate 80 times harder to compress than water with Bulk Modulus 2.15×109 Pa.
  31. [31]
    Properties of Diamond - Steve Sque
    Mechanical and Thermal Properties ; Bulk modulus, 442.3 GPa [3] ; Linear expansion coefficient (300 K), 1.05×10-6 K-1 [2] ; Melting point, 3773 K [2] 4027 °C [10]
  32. [32]
    Relations between the cohesive energy, atomic volume, bulk ...
    Aug 10, 2025 · By analysing the experimental data available in the literature, it has been found that the bulk modulus B of metals is proportional to the cohesive energy ...
  33. [33]
    First-principles calculations of the bulk modulus of diamond
    The bulk modulus can be obtained from first-principles theories. In this case, the total energy is calculated as a function of volume and then fitted to some ...
  34. [34]
    (PDF) Temperature dependence of the elastic constants of aluminum
    Aug 7, 2025 · Over the temperature range investigated the isothermal bulk modulus and the two shear moduli have a simple exponential dependence on isobaric ...
  35. [35]
    Bulk Modulus: What is it? When is it Important?
    the measure of a fluid's resistance to compression —of a hydraulic fluid if position, response time, and stability are
  36. [36]
    Novel bulk modulus model to estimate stiffness in fluid power systems
    This study presents a novel model for the bulk modulus. The representation does not rely on the tangent or the secant definition of the bulk modulus.
  37. [37]
    [PDF] PRESSURE VESSELS - MIT
    Aug 23, 2001 · The bulk modulus. K, also called the modulus of compressibility, is the ratio of the hydrostatic pressure p needed for a unit relative ...
  38. [38]
    Bulk Modulus: Definition, How it Works, Formula, Examples, and ...
    Apr 15, 2023 · Bulk modulus is a measure of a material's resistance to uniform compression. It describes how much a material decreases in volume when pressure ...
  39. [39]
    Properties of iron at the Earth's core conditions - Oxford Academic
    It is thus concluded that the Earth's inner core is very likely to be virtually pure iron in its hexagonal close packed (hcp) phase.
  40. [40]
    An equation of state for liquid iron and implications for the Earth's core
    The equation of state parameters, centered at 1 bar and 1811 K (the normal melting point of iron), are density, ρ0=7019 kg/m3, isentropic bulk modulus, KS0= ...<|separator|>
  41. [41]
    Thermal expansivity, heat capacity and bulk modulus of the mantle
    We derive exact expressions for the thermal expansivity, heat capacity and bulk modulus for assemblages with arbitrarily large numbers of components and phases.
  42. [42]
    PRINCIPLES OF UNDERWATER SOUND Chapter 8
    Although seawater is almost a thousand times denser than air, the enormous bulk modulus of water is the more important factor determining sound speed. Of ...
  43. [43]
    Numerical simulation of cavitation threshold in water and ...
    The cavitation threshold P and frequency f can be expressed as P = Afα + B. •. As the number of bubbles increases, the cavitation threshold indicates a non- ...
  44. [44]
    From compliance to moduli: clarifying basic mechanical properties of ...
    Compliance is said to be the inverse of stiffness (13), the elastic characteristic of the arterial system (5), or the elastic behavior of a hollow vessel or ...
  45. [45]
    Introduction to ultrasound elastography - PMC - NIH
    This article describes the physical basis of both elastographic methods: compression elastography and shear wave elastography.
  46. [46]
    [PDF] From the interatomic potential, we can immediately obtain a number ...
    The bulk modulus requires the stress P to be applied in all directions, and ... We conclude by calculating Young's modulus using this version of ...
  47. [47]
    Full article: Interatomic potentials: achievements and challenges
    Here, we review empirical interatomic potentials designed to reproduce elastic properties, defect energies, bond breaking, bond formation, and even redox ...
  48. [48]
    [PDF] Determination of the Bulk Modulus of a Lennard-Jones Solid
    The quantum Monte Carlo method is used to determine the bulk modulus of a Lennard-Jones solid. The quantum. Monte Carlo method is discussed in detail and is ...
  49. [49]
    Bulk modulus for polar covalent crystals | Scientific Reports - Nature
    Oct 29, 2013 · Bulk modulus measures material's resistance to uniform compression, and is highly related to the chemical composition and crystal structure.
  50. [50]
    Giant Phonon Anharmonicity and Anomalous Pressure Dependence ...
    Jul 19, 2016 · In this paper, we predict an abnormally negative trend dκ L /dP < 0 in Y 2 Si 2 O 7 silicate using density functional theoretical calculations.
  51. [51]
    Pressure-induced softening as a common feature of framework ...
    Jun 26, 2013 · The correlation between the negative thermal expansion and the pressure-induced softening is investigated on the basis of thermodynamics. Figure ...
  52. [52]
    Pressure and temperature dependent ab-initio quasi-harmonic ...
    Apr 15, 2024 · We present the ab-initio temperature and pressure dependent thermoelastic properties of body-centered cubic tungsten.
  53. [53]
    Quasi harmonic approximation — Phonopy v.2.43.6
    Using phonopy results of thermal properties, thermal expansion and heat capacity at constant pressure can be calculated under the quasi-harmonic approximation.
  54. [54]
    Appendix D: Calculation of the Elastic Moduli - Wiley Online Library
    For a cubic crystal the bulk modulus B D (C11 C 2C12)/3. Bulk modulus is deter- mined from the curvature of the E(V) curve (see Section 10.3).
  55. [55]
    Lamé parameters of common rocks in the Earth's crust and upper ...
    Jun 24, 2010 · Interestingly and most notably, only λ and μ appear in Hooke's law, but not Young's modulus (E), the bulk modulus (K), or Poisson's ratio (υ).
  56. [56]
    Reliably Modeling the Mechanical Stability of Rigid and Flexible ...
    In contrast, the first Born stability criterion to fail for the LP phase of MIL-53(Al) is C66 – P. This corresponds to a shear deformation on the yz plane along ...Introduction · Theoretical Methods To... · Closing Theoretical Guidelines...
  57. [57]
    Mechanical stability conditions for 3D and 2D crystals under ... - arXiv
    Nov 24, 2024 · In the literature there are at least four stability criteria for a loaded material, which differ in the choice of appropriate stiffness tensors ...
  58. [58]
    [PDF] Elastic constants of defected and amorphous silicon with the ... - MIT
    Oct 29, 2004 · Coordination and ring statistics changes compared to the percent change in bulk modulus for vacancy defects and the. FFCD. Net changes are for ...
  59. [59]
    Impact of vacancies on structure, stability and properties of ...
    For NbB2 and TaB2 we find a decreased bulk modulus with increasing vacancy concentration. For both MoB2 and WB2 with M-vacancies we see a clear decrease in the ...