Fact-checked by Grok 2 weeks ago

Density

Density is a fundamental of that quantifies the of a substance per , serving as a measure of how compact or concentrated the material is. It is mathematically expressed by the formula ρ = m / V, where ρ (the Greek letter rho) represents density, m is the , and V is the . The SI of density is kilograms per cubic meter (kg/m³), though other common units include grams per cubic centimeter (g/cm³) for solids and liquids. As an intensive property, density remains constant for a given substance regardless of the sample , distinguishing it from extensive properties like or that scale with . This makes density invaluable for identifying and characterizing materials, such as distinguishing metals like lead (density ≈ 11.34 g/cm³) from lighter substances like aluminum (density ≈ 2.70 g/cm³). In fluids, density plays a critical role in phenomena like , where objects denser than the surrounding fluid sink, while less dense ones float, as explained by . Density varies with and , particularly for gases and liquids, due to changes in volume while mass remains constant; for instance, most substances expand when heated, decreasing their density. In scientific applications, density measurements enable quality control in industries like and pharmaceuticals—for example, assessing content in beverages or purity in oils—and support , such as tracking through density gradients. It also informs designs, from to ensure flotation to in for structural integrity.

Definition and Fundamentals

Core Definition

Density, denoted by the Greek letter ρ, is the m of a substance divided by its V, expressed as \rho = \frac{m}{V}. This quantity, known as mass density, characterizes how compactly the matter in a substance is packed. Mass density is distinct from other types of density, such as (the number of particles per unit volume) or (the per unit volume), which apply to specific contexts in physics like or . As an intensive , density remains constant regardless of the sample size or ; for example, a small or a large block of the same material will exhibit the same density value under identical conditions. In everyday scenarios, density determines whether objects sink or float in fluids: lead sinks in due to its higher density, while floats because of its lower density relative to . This principle underlies effects observed in fluids.

Mathematical Formulation

The mass density \rho of is fundamentally defined as the ratio of to in the as the approaches zero, ensuring a pointwise characterization suitable for continuous . Specifically, for a point at \mathbf{r}, the density is given by \rho(\mathbf{r}) = \lim_{\Delta V \to 0} \frac{\Delta m}{\Delta V}, where \Delta m is the contained within an infinitesimal \Delta V surrounding \mathbf{r}. This limiting process arises from the in , which posits that matter can be modeled as a continuous distribution despite its atomic structure, allowing density to capture local concentration without regard to microscopic discreteness. For materials with non-uniform , density extends to a position-dependent \rho(\mathbf{r}), representing the local value at each point \mathbf{r} via the same : \rho(\mathbf{r}) = \lim_{\Delta V \to 0} \frac{\Delta m}{\Delta V} as \Delta V shrinks around \mathbf{r}. In , this local density is treated as a \rho: \mathbb{R}^3 \to \mathbb{R}, varying continuously across the to describe heterogeneous distributions. The total m of a body occupying volume V is then obtained by integrating this field: m = \int_V \rho(\mathbf{r}) \, dV, which follows directly from the additivity of over disjoint subvolumes and the definition of the as a of sums. In the framework of , density acquires additional complexity due to spacetime , manifesting as components of the stress-energy tensor T_{\mu\nu}. Here, proper density refers to the invariant measured in the local of the , while coordinate density denotes the value in a specific , which transforms non-trivially under changes of coordinates. This distinction underscores the tensorial nature of density in curved spacetime, where it contributes to the geometry via Einstein's field equations without altering the core limit-based definition.

Historical Development

Early Observations and Concepts

Early observations of density-like phenomena emerged from intuitive understandings of why objects sink or float in , laying the groundwork for more formalized concepts. In ancient civilizations, such as and , practical knowledge of material weights relative to volume informed construction and metallurgy, though these were not systematically theorized. The ancient advanced this into philosophical and empirical , distinguishing between concepts akin to modern density and mere weight. Aristotle (384–322 BCE) conceptualized natural elements as having inherent "light" or "heavy" properties that determined their tendency to move upward or downward in a medium like air or water. In his work On the Heavens, he described lighter elements like fire and air as rising, while heavier ones like earth and water sank, attributing this to their natural places in the cosmos rather than a quantitative ratio of mass to volume. This qualitative framework influenced Western thought for centuries, emphasizing elemental affinities over precise measurement. A pivotal empirical breakthrough occurred with (c. 287–212 BCE), who linked to the displacement of fluid, providing an early method to assess density. Legendarily, while tasked by King Hieron II to verify if was pure or adulterated, Archimedes realized in his bath that the volume of displaced water equaled the object's submerged volume, allowing density calculation as mass divided by volume without damaging the crown—famously exclaiming "!" This principle, detailed in his , established that an object's density relative to the fluid determines flotation, forming the basis of . During the medieval period, these ideas persisted in and Islamic scholarship, where scholars like (c. 1115–1130) explored balances for comparing material densities in works like The Book of the Balance of Wisdom, integrating Greek principles with experimental refinements. In the , (1564–1642) built on in (1638), analyzing why bodies of the same material but different float differently, emphasizing and shape over intrinsic heaviness alone. The transition to quantitative density occurred with (1643–1727), who in (1687) treated density as the ratio of to , using it to describe gravitational attraction and fluid equilibrium. Newton's formulation, such as in Proposition 19 of Book II, applied density to explain hydrostatic pressure variations, shifting from qualitative to mathematical rigor. This marked the evolution from observational concepts to a foundational .

Evolution of Measurement Techniques

During the , the era marked a pivotal shift toward quantitative in scientific measurements, including density, through refinements in balances and volumetric devices. , leveraging advanced analytical balances, conducted meticulous determinations that facilitated accurate density computations when integrated with or volumetric techniques, emphasizing the balance's role in establishing conservation laws in chemistry. Lavoisier further contributed by improving designs, such as those used to assess the density of mineral waters, thereby enhancing the reliability of evaluations for liquids. These instruments, often hydrostatic balances adapted for buoyancy-based measurements, allowed chemists to control for purity in metals and solutions with greater accuracy than previous qualitative methods. The 19th century brought further innovations, particularly in s and specialized apparatus for both liquids and solids. William Nicholson's constant-volume , introduced in 1784 but widely adopted and refined in the early 1800s, enabled precise specific gravity measurements for liquids and by maintaining a fixed displaced while varying weights. For liquids, Henri Victor Regnault's pycnometers, designed in 1843, represented a breakthrough with their calibrated flasks that minimized errors from and fluctuations, achieving densities with uncertainties below 0.1%. Regarding solids, oscillation-based methods emerged toward the century's end, such as those involving torsional pendulums or vibrating systems to infer density from resonant frequencies in fluids, though via spring balances like Philipp von Jolly's 1881 apparatus became standard for routine specific gravity assessments without destructive sampling. The introduced non-invasive and microscopic techniques, expanding density measurements beyond macroscopic samples. , developed by and his son William Lawrence Bragg starting in 1912, permitted atomic-level density calculations by determining volumes and atomic arrangements; early applications in the 1920s, such as for , yielded theoretical densities matching experimental values within 1%, revolutionizing solid-state analysis. Ultrasound methods for non-destructive testing gained traction from the 1940s, using pulse-echo propagation to measure acoustic velocities in materials like metals and composites, from which bulk densities could be derived via empirical correlations with wave attenuation and impedance, offering in-situ assessments without material alteration. In contemporary practice, computational approaches have complemented experimental methods, particularly through (DFT) in . The foundational Hohenberg-Kohn theorems of 1964 established that the ground-state uniquely determines all molecular properties, enabling DFT algorithms to predict material densities ab initio; for instance, calculations for semiconductors and polymers now achieve accuracies rivaling measurements, with errors often under 5% for bulk moduli and densities.

Units and Standards

Common Units and Systems

The (SI) defines density as per unit volume, with the standard unit being the per cubic meter (/m³), derived from the base units of (, ) and length (meter, m), where volume is expressed as cubic meters (m³). This unit is universally adopted in scientific and international contexts for its coherence with other SI-derived quantities. In and customary systems, density is commonly expressed in s per (lb/ft³), where the (lb) represents , though slugs per (slug/ft³) is used in contexts requiring consistency with force units like the pound-force (lbf). These units have historical prevalence in engineering applications, particularly in the and , for structural and calculations predating widespread metric adoption. Other frequently used units include grams per cubic centimeter (g/cm³), which is convenient for measuring the density of and liquids due to its alignment with typical scales. For dilute aqueous solutions, such as in environmental or chemical , milligrams per liter (mg/L) serves as a practical , approximating mass concentration where the solution's volume is close to that of the . In , density has the formula [M L⁻³], where M denotes and L denotes , reflecting its role as mass distributed over . This dimension connects density to derived quantities like ([M L T⁻²]) through relations such as weight (force equals times ) and ([M L⁻¹ T⁻²]), as seen in hydrostatic equations where depends on density, , and depth.
System/UnitUnit SymbolTypical Application
SIkg/m³General scientific and engineering use
/US Customarylb/ft³ or slug/ft³,
CGS/Metricg/cm³Solids and liquids in laboratories
Environmental/Chemicalmg/LDilute solutions and concentrations
, or specific gravity, is a dimensionless comparing a substance's density to a reference (often ), avoiding unit dependencies.

Conversions and Relative Density

Density units can be converted between different systems using , a method that ensures consistency by treating units as algebraic quantities. For instance, to convert from grams per cubic centimeter (g/cm³) to kilograms per cubic meter (kg/m³), multiply the value by the factor 1000, derived from the relations 1 kg = 1000 g and 1 m³ = 10^6 cm³, yielding 1 g/cm³ = 1000 kg/m³. This approach applies broadly: identify the conversion factors for and separately, then combine them to cancel undesired units step by step. Relative density, also known as specific gravity, is defined as the unitless ratio of the density of a substance to the density of a reference substance under specified conditions. For liquids and solids, the reference is typically at , where its density is maximized at approximately 1000 kg/m³; for gases, the reference is dry air at . This ratio, denoted as SG, is expressed mathematically as \text{SG} = \frac{\rho}{\rho_\text{reference}} where \rho is the density of the substance and \rho_\text{reference} is the density of the reference material, often (\rho_\text{H_2O}) for liquids. The concept of specific gravity has historical roots in chemistry, notably through the developed by French chemist Antoine Baumé in 1768, which used hydrometers to measure relative densities of liquids for industrial applications like and acid production. This scale provided two versions—one for liquids denser than and one for those less dense—calibrated against specific gravity to simplify comparisons without needing absolute density values. The primary advantage of relative density lies in its unitless nature, enabling direct, standardized comparisons of material properties across substances and conditions, independent of the unit system used for absolute density. However, specific gravity values are temperature-dependent, as both the substance's density and the reference density vary with ; thus, measurements must specify the temperature to ensure accuracy and comparability.

Measurement Methods

For Homogeneous Solids and Liquids

For homogeneous solids, density is commonly measured using , which relies on the buoyant force experienced by an object submerged in a . The principle states that the buoyant force F_b equals the weight of the displaced by the object, given by F_b = \rho_f V [g](/page/G), where \rho_f is the density, V is the of the displaced (equal to the object's for full submersion), and g is the . To apply this, the solid is weighed in air (W_{\text{air}} = m [g](/page/G), where m is the ) and then submerged in a of known density, such as , to obtain the W_{\text{fluid}}. The is calculated as V = \frac{W_{\text{air}} - W_{\text{fluid}}}{\rho_f [g](/page/G)}, and the density follows as \rho = \frac{m}{V} = \frac{W_{\text{air}} \rho_f}{W_{\text{air}} - W_{\text{fluid}}}. This is particularly suitable for regular-shaped homogeneous solids like metal cubes or cylinders, as it directly determines without needing geometric measurements. Modern implementations often use analytical balances for precise weighing, achieving accuracies on the order of 0.1% for densities around 1–10 /cm³. For liquids, pycnometry provides a standard volumetric technique by employing a pycnometer—a flask of precisely known , typically calibrated with . The process involves weighing the empty pycnometer (m_{\text{empty}}), filling it completely with the sample and weighing again (m_{\text{liquid}}), and sometimes accounting for the flask's calibration using a reference . The density is then computed as \rho = \frac{m_{\text{liquid}} - m_{\text{empty}}}{V_{\text{flask}}}, where V_{\text{flask}} is the internal at a specified . This method ensures high precision for homogeneous liquids like oils or solvents, as the pycnometer's design minimizes air bubbles and evaporation errors. standards, such as those from the International Union of Pure and Applied Chemistry (IUPAC), recommend using deionized at 20°C for determination, yielding results traceable to units. Typical laboratory pycnometers achieve resolutions better than 0.001 g/cm³ for volumes around 10–50 mL. A more advanced approach for both solids (via immersion) and liquids is the digital density meter, which utilizes the oscillating principle invented in the by researchers at the company. In this device, a U-shaped tube filled with the sample vibrates at its , where the T is proportional to the of the sample's density: T \propto \sqrt{\rho}. The meter's measure T and, combined with the tube's constants, compute \rho directly. This technique offers rapid, non-destructive measurements with , making it ideal for in homogeneous materials. Since its development, refinements have included to ±0.01°C, ensuring density accuracy to 0.0001 g/cm³ in standard lab environments. For solids, samples can be dissolved in a or measured via in the meter's fluid chamber, though primary use remains for liquids.

For Gases and Fluids

Measuring the density of gases and fluids requires techniques that account for their compressibility and potential flow dynamics, distinguishing these methods from those for solids. For gases under low-pressure conditions where ideal behavior predominates, density is commonly calculated using the ideal gas law rearranged as \rho = \frac{P M}{R T}, where \rho is the mass density, P is the pressure, M is the molar mass, R is the universal gas constant, and T is the absolute temperature. This formula derives from the standard ideal gas equation PV = nRT, substituting mass m = nM and density \rho = m/V, yielding a direct relation between measurable thermodynamic properties and density without requiring direct volume measurement of the gas sample. At higher pressures, where intermolecular forces cause deviations from ideality, the provides corrections to this formula by incorporating virial coefficients into the equation of state. The second virial coefficient B, the leading correction term, accounts for pairwise molecular interactions and modifies the Z in \rho = \frac{P M}{R T Z}, where Z = 1 + B \rho + \ higher\ order\ terms, with \rho here as ; this adjustment is essential for accurate density predictions up to moderate pressures around 100 atm. Higher-order coefficients like the third virial C may be included for denser conditions, but the second-order term often suffices for many applications involving compressed gases. For fluids, particularly liquids, the sinker (or ) method employs of a calibrated reference body to determine density indirectly through effects. A sinker of known volume V_s (typically or metal) is weighed in air to obtain its m_a, then immersed in the sample and weighed again to get the apparent mass m_f; the fluid density \rho_f is then calculated as \rho_f = \frac{m_a - m_f}{V_s} + \rho_{air}, where \rho_{air} accounts for air (often negligible). This technique is precise for homogeneous liquids, requiring only about 40 ml of sample and achieving accuracy to three decimal places, provided no air bubbles adhere to the sinker during immersion. In dynamic scenarios involving flowing fluids or gases, Coriolis flow meters offer a non-invasive approach to density measurement by leveraging the Coriolis effect on a vibrating tube. Density is measured from the resonant frequency of the tube, which depends on the fluid's density. The is obtained from phase shifts between sensors due to the on the flowing fluid, allowing computation of volumetric flow if needed. The meter's ability to simultaneously capture density, mass flow, and temperature without needing separate corrections for fluid properties makes it ideal for industrial processes like pipeline monitoring, where and flow variations are prominent.

For Heterogeneous and Irregular Materials

Measuring the density of heterogeneous materials, which exhibit varying composition or structure, and irregular materials, which lack uniform shapes, presents unique challenges compared to homogeneous substances, as voids, porosity, and non-uniformity can skew results if not accounted for properly. These materials include soils, powders, composites, and oddly shaped solids like rocks or artifacts. Accurate assessment requires methods that either isolate the solid phase or average over representative volumes to capture effective density. For irregular solids, the water displacement method, based on , is widely used to determine by submerging the object in a and measuring the rise in level. The density \rho is then calculated as \rho = \frac{m}{\Delta V}, where m is the of the solid and \Delta V is the displaced . This approach is effective for non-porous or low-porosity items, such as stones or metal castings, as it avoids direct geometric measurement. In powders and non-compact granular materials, a distinction is made between bulk density and true density to address inter-particle voids. Bulk density \rho_\text{bulk} is computed as \rho_\text{bulk} = \frac{m}{V_\text{container}}, where m is the mass and V_\text{container} is the total volume occupied by the powder in a container, including air spaces; this metric is crucial for handling and storage applications. True density, excluding voids, is measured via gas pycnometry, typically using helium to fill accessible pores and determine the skeletal volume of the solid phase. For example, in metal powders for additive manufacturing, true densities around 4.5 g/cm³ for titanium alloys are obtained this way, contrasting with bulk densities of 2.5–3.0 g/cm³. Heterogeneous materials like soils or composites require careful sampling to ensure representativeness, as local variations can lead to biased density estimates. Techniques involve collecting multiple subsamples from different locations or depths and averaging their densities, often using composite sampling with at least eight cores per site to minimize variability. In soils, for instance, this approach accounts for rock fragments and , yielding bulk densities typically ranging from 1.0 to 1.6 g/cm³ depending on . For materials with internal voids or complex microstructures, computed tomography () scanning provides a non-destructive solution to quantify effective density. X-ray generates density maps from voxel data, identifying voids as regions of near-zero density and computing overall density as the mass per volume. This method is particularly valuable for porous composites, revealing void fractions up to 10–20% that influence mechanical properties.

Variations and Influences on Density

Temperature and Pressure Effects

Density, defined as mass per unit volume, varies with and due to changes in the volume occupied by a fixed of . For most substances, an increase in causes , where the volume increases while remains constant, thereby decreasing density. The relative volume change is given by \Delta V / V = \beta \Delta T, where \beta is the of volume and \Delta T is the change. Consequently, the density at T can be approximated as \rho(T) = \rho_0 / (1 + \beta \Delta T), where \rho_0 is the density at the reference , valid for small \Delta T where \beta is approximately constant. Pressure influences density through , which quantifies how decreases under applied at constant . The isothermal \kappa is defined as \kappa = -\frac{1}{V} \left( \frac{\partial V}{\partial P} \right)_T. For small changes, the density increases approximately as \rho(P) \approx \rho_0 (1 + \kappa \Delta P), reflecting the inverse relationship between and density. This effect is more pronounced in gases and liquids than in , where \kappa is typically small. Water exhibits anomalous behavior in its temperature-density relationship, reaching maximum density of approximately 0.99984 g/cm³ at under standard ; above or below this temperature, density decreases due to . This property arises from hydrogen bonding and has significant implications for ecosystems, as it allows to float while denser sinks. In contrast, air density at is about 1.225 kg/m³ under standard conditions of 15°C and 1013.25 , decreasing with altitude primarily due to reduced , though variations also contribute; for instance, density drops to roughly half at 5.5 km elevation. For ideal gases, the equation of state PV = nRT directly relates density to environmental conditions, where density \rho = \frac{m}{V} = \frac{PM}{RT} (with M as molar mass), implying \rho \propto P / T. This proportionality explains why gas density increases with pressure and decreases with temperature at constant pressure, or vice versa.

Phase Transitions and Composition Changes

Phase transitions in materials often involve abrupt changes in density due to rearrangements in atomic or molecular structures, distinguishing them from gradual variations with temperature or pressure. In first-order phase transitions, such as those between solid, liquid, and gas states, density exhibits discontinuities because the transition occurs at a specific temperature and pressure where two phases coexist, accompanied by the absorption or release of latent heat. This latent heat represents the energy required to overcome intermolecular forces without changing temperature, leading to a sudden volume change and thus a jump in density. For instance, during the melting of ice into liquid water at 0°C and standard pressure, the density increases discontinuously from approximately 0.917 g/cm³ for ice Ih to 1.000 g/cm³ for liquid water, as the open hexagonal structure of ice collapses into a more compact liquid arrangement due to weakened hydrogen bonding. Allotropic forms of elements, which are different crystal structures of the same pure substance, can exhibit significant density differences arising from variations in atomic packing efficiency. Carbon provides a classic example: , with its rigid tetrahedral sp³-hybridized structure, has a density of 3.51 g/cm³, while , featuring layered sp²-hybridized sheets held by weak van der Waals forces, has a lower density of 2.26 g/cm³. This disparity stems from the denser close-packing in compared to the interstitial spaces between layers, influencing stability under where is favored. Isotopic substitution can also alter density through subtle changes in without affecting the overall , leading to measurable differences in materials like . Heavy (D₂O), where protium is replaced by , has a density of 1.105 g/cm³ at 20°C, compared to 0.998 g/cm³ for ordinary (H₂O), a roughly 10.6% increase attributable to the higher mass of atoms while maintaining similar molecular volumes. This isotopic effect is particularly relevant in studies of hydrogen bonding and phase behavior, as it influences properties like without inducing phase changes. In metals, polymorphic transitions between crystal structures often occur upon heating or under pressure, resulting in density shifts due to changes in and packing. For iron, the transition from body-centered cubic (BCC) α-phase to face-centered cubic (FCC) γ-phase at around 912°C involves a slight density increase from approximately 7.53 g/cm³ to 7.63 g/cm³, owing to the higher packing efficiency of the FCC structure (74% vs. 68% for BCC), which leads to a more compact atomic arrangement outweighing the effects of at that temperature. This discontinuity accompanies absorption and impacts applications in , such as processing.

Density in Solutions and Mixtures

Calculation Methods for Solutions

The density of a can be calculated under the assumption of ideal mixing, where the volumes of the solute and are additive. In this case, the total mass of the solution is the sum of the masses of the solute and , and the total volume is the sum of their individual volumes. The density \rho is then given by \rho_\text{solution} \approx \frac{m_\text{solute} + m_\text{solvent}}{V_\text{solute} + V_\text{solvent}}, where m denotes and V denotes . This approximation holds for dilute solutions or systems with minimal interactions between components. For non-ideal solutions, where solute-solvent interactions lead to volume changes, the apparent molar volume V_\phi of the solute accounts for deviations from additivity. It is defined as V_\phi = \frac{V_\text{solution} - n_1 V_1^\circ}{n_2}, where V_\text{solution} is the total volume of the solution, n_1 is the number of moles of solvent, V_1^\circ is the molar volume of the pure solvent, and n_2 is the number of moles of solute. This quantity, derived from experimental density measurements, quantifies the effective volume contribution of the solute, often revealing contractions or expansions due to molecular associations. Electrolyte solutions frequently require empirical equations to capture concentration-dependent density variations, as ionic and electrostatic effects cause non-linearity. A common linear approximation for low concentrations is \rho = \rho_\text{solvent} + k c, where c is the solute concentration (e.g., in ) and k is an empirical constant specific to the - pair. More comprehensive models, such as those fitted to extensive experimental data, extend this to multicomponent systems with parameters for individual ions, achieving accuracies within 0.1 /m³ for single electrolytes. Seawater exemplifies density, where dissolved salts (primarily NaCl at ~35 g/kg) increase the density from that of pure (1.00 g/cm³ at ) to approximately 1.025 g/cm³ at the surface, reflecting the additive mass effect of ions with minimal change. In contrast, alcohol- mixtures like - demonstrate upon mixing, where the is less than the sum of pure component , resulting in higher densities than predicted by additivity—for instance, a maximum occurs near 0.2 , with density deviations up to several percent at ambient temperatures.

Behavior in Alloys and Composites

In alloys, which are homogeneous mixtures of metals, the density is typically calculated using the inverse based on mass fractions to account for the additive nature of volumes in solid solutions. The formula is given by \rho = \frac{1}{\sum (w_i / \rho_i)}, where w_i is the mass fraction of component i and \rho_i is its pure density; this approach assumes no significant volume change upon mixing and is widely used for predicting densities in metallic . For instance, , an iron-carbon alloy with approximately 0.2-2% carbon by weight, exhibits a density of about 7.85 g/cm³, slightly lower than pure iron's 7.87 g/cm³ due to the lower density of carbon (2.26 g/cm³) and minor distortions. In fiber-reinforced composites, density behavior follows the direct rule of mixtures using volume fractions, expressed as \rho_c = \sum v_i \rho_i, where v_i is the volume fraction of phase i; this reflects the parallel addition of masses in distinct phases like fibers and matrix. This model assumes perfect bonding without voids and is foundational for designing lightweight materials, such as those in applications. For example, composites, consisting of fibers (density ~2.5 g/cm³) embedded in a matrix (density ~1.2 g/cm³), achieve an overall density of 1.5-2.0 g/cm³, significantly lower than pure due to the lower-density matrix dominating the volume. Porosity introduces voids that reduce the effective density in both alloys and composites, with the apparent density related to the true material density by \rho_{\text{apparent}} = \rho_{\text{true}} (1 - \phi), where \phi is the fraction. In metal matrix composites, from imperfect infiltration can lower density by 5-10%, compromising mechanical integrity, while in composites, it arises from trapped air during curing and similarly diminishes load-bearing capacity. Controlling \phi below 2% is critical for achieving theoretical densities in high-performance applications.

Examples and Data

Densities of Elements and Common Substances

The densities of pure elements provide fundamental reference values for understanding material properties and are typically measured under standard conditions. For gases, densities are reported at standard temperature and pressure (STP, 0°C and 1 atm) in grams per liter (g/L), while for liquids and solids, values are given in grams per cubic centimeter (g/cm³) at 20°C unless otherwise noted. These measurements reflect the inherent packing efficiency and atomic masses of the elements. The following table lists densities for a selection of 20 elements, spanning gases, liquids, and solids, drawn from compiled standard data:
ElementSymbolAtomic NumberDensityState/Condition
HydrogenH10.0899 g/LGas, STP
HeliumHe20.1785 g/LGas, STP
LithiumLi30.534 g/cm³Solid, 20°C
BerylliumBe41.848 g/cm³Solid, 20°C
CarbonC62.26 g/cm³Solid (graphite), 20°C
NitrogenN71.2506 g/LGas, STP
OxygenO81.429 g/LGas, STP
NeonNe100.9 g/LGas, STP
SodiumNa110.971 g/cm³Solid, 20°C
AluminumAl132.702 g/cm³Solid, 20°C
CalciumCa201.55 g/cm³Solid, 20°C
TitaniumTi224.54 g/cm³Solid, 20°C
IronFe267.874 g/cm³Solid, 20°C
CobaltCo278.9 g/cm³Solid, 20°C
SilverAg4710.5 g/cm³Solid, 20°C
GoldAu7919.32 g/cm³Solid, 20°C
MercuryHg8013.546 g/cm³Liquid, 20°C
PlatinumPt7821.45 g/cm³Solid, 20°C
LeadPb8211.35 g/cm³Solid, 20°C
OsmiumOs7622.6 g/cm³Solid, 20°C
These values are sourced from standard compilations and illustrate the wide range, from the low densities of light gases to the high densities of heavy transition metals. Densities of common substances, which often include compounds or composites, also serve as practical references in and daily applications. For instance, varies by and content, typically ranging from 0.4 to 0.8 g/cm³ for seasoned hardwoods and softwoods at 20°C. , a of , aggregates, and , has a standard density of about 2.4 g/cm³ at 20°C, reflecting its reinforced structure. Mercury, a , possesses a density of 13.6 g/cm³ at 20°C, making it one of the densest common substances. All listed densities are at 20°C and 1 unless specified otherwise, consistent with conventions in authoritative references such as the CRC Handbook of Chemistry and Physics. Across the periodic table, elemental densities generally increase with , particularly among metals, due to tighter atomic packing and higher atomic masses that enhance mass per unit volume without proportionally increasing volume.

Density Profiles in Air and Water

In Earth's atmosphere, air density decreases exponentially with increasing altitude, primarily due to the reduction in pressure under . This variation is described by the , approximated as \rho(z) \approx \rho_0 e^{-z/H}, where \rho(z) is the density at altitude z, \rho_0 = 1.225 /m³ is the sea-level density at standard conditions (15°C and 101.325 kPa), and H \approx 8.5 km is the , representing the altitude over which density falls by a factor of e. The arises from the balance of gravitational forces and in the , with H = RT / (Mg), where R is the , T is , M is , and g is . For instance, at 30 km altitude, density decreases to approximately 0.018 /m³, reflecting about 1.5% of sea-level value and influencing and atmospheric dynamics. The following table presents air density from the U.S. Standard Atmosphere model for selected altitudes between 0 and 100 km, based on geometric height and standard temperature-pressure profiles. Values beyond 80 km are extrapolated from extended models, showing the transition to the where densities approach negligible levels.
Altitude (km)Density (kg/m³)
01.225
100.414
200.089
300.018
400.0040
500.0010
600.00031
700.000083
800.000018
1000.00000056
In the , density profiles vary with temperature, , and , shaping oceanic . Pure achieves maximum density of 999.97 kg/m³ at under standard , decreasing at higher or lower temperatures due to and molecular structuring effects. , with average around 35 parts per thousand (ppt), has a typical surface density of about 1025 kg/m³ at 20°C, but this increases with (adding roughly 0.8 kg/m³ per ppt) and depth-induced (about 0.017 kg/m³ per meter in the deep ocean). Oceanic density profiles exhibit distinct layers: the surface (0–100 m) has lower density (~1020–1025 kg/m³) due to solar warming and freshwater inputs; the (100–1000 m) shows rapid density increase from cooling; and the deep ocean (>1000 m) reaches higher densities (~1028 kg/m³) from compressive pressure and uniform cold temperatures around 2–4°C, driving . These gradients maintain , with density anomalies of 1–3 kg/m³ sufficient to separate layers and influence global heat transport. The table below summarizes water density for pure (salinity 0 ) and typical (salinity 35 ) across temperatures from 0 to 100°C at (1 ), illustrating thermal effects; deep-ocean values would be 1–3% higher due to . Data are interpolated from standard equations of state for .
Temperature (°C)Density at 0 ppt (kg/m³)Density at 35 ppt (kg/m³)
0999.841028.0
4999.971028.0
10999.701027.1
20998.201024.8
30995.651022.2
40992.221019.2
50988.041015.8
60983.201012.0
80971.791002.6
100958.36990.2

Molar Volumes of Phases

The of a , denoted V_m, is the volume occupied by one of a substance in that and is calculated using the relation V_m = \frac{M}{\rho}, where M is the (g/) and \rho is the mass density (g/cm³). This yields V_m in cm³/, providing a measure of or molecular spacing at the , which links macroscopic density to microscopic . For , molar volumes differ between and due to changes in packing efficiency during . In solid phases, molar volumes reflect crystal lattice types; for instance, loosely packed body-centered cubic (bcc) metals like the alkali metals exhibit larger V_m values (e.g., 13.02 cm³/mol for solid lithium at 298 K), while transition metals with denser hexagonal close-packed or face-centered cubic structures show smaller values (e.g., 7.09 cm³/mol for solid iron at 298 K). Liquid phases typically display slightly expanded molar volumes owing to thermal disorder, with ratios of liquid to solid V_m often 1.02–1.05 for metals, though some elements like gallium exhibit anomalous contractions upon melting (solid 11.81 cm³/mol vs. liquid 11.44 cm³/mol at 302.91 K). Non-metallic anomalies are pronounced in water, where the liquid phase at 298 K has V_m = 18.07 cm³/mol, exceeding that of ice Ih at 273 K (19.65 cm³/mol) due to the expansive hydrogen-bonded framework in the solid. The following table presents molar volumes for over 50 elements in their solid phases at 298 K (unless noted) and liquid phases where data are available, typically at or near the melting point. Values for gases are omitted, as the focus is on condensed phases; elements without liquid data at standard or accessible conditions (e.g., high melting points) are marked N/A. Data are derived from standard compilations, with phase notes indicating conditions. Liquid values updated from reliable density measurements at melting point.
ElementSolid V_m (cm³/mol)Liquid V_m (cm³/mol)Phase Notes
Li13.0213.56Solid at 298 K; liquid at 453 K (m.p.)
Be4.885.33Solid at 298 K; liquid at 1560 K (m.p.)
B4.395.00Solid at 298 K; liquid at 2348 K (m.p.)
C5.297.46Graphite solid at 298 K; liquid at 4765 K (est. m.p.)
Na23.7524.80Solid at 298 K; liquid at 371 K (m.p.)
Mg13.9815.35Solid at 298 K; liquid at 923 K (m.p.)
Al10.0011.35Solid at 298 K; liquid at 931 K (m.p.)
Si12.0512.72Solid at 298 K; liquid at 1687 K (m.p.)
P16.99N/AWhite phosphorus solid at 298 K; high m.p.
S15.53N/ARhombic solid at 298 K; high m.p.
K45.6855.25Solid at 298 K; liquid at 337 K (m.p.)
Ca25.8626.68Solid at 298 K; liquid at 1115 K (m.p.)
Sc15.06N/ASolid at 298 K; high m.p.
Ti10.6211.80Solid at 298 K; liquid at 1941 K (m.p.)
V8.34N/ASolid at 298 K; high m.p.
Cr7.287.94Solid at 298 K; liquid at 2180 K (m.p.)
Mn7.358.12Solid at 298 K; liquid at 1519 K (m.p.)
Fe7.098.00Solid at 298 K; liquid at 1811 K (m.p.)
Co6.627.41Solid at 298 K; liquid at 1768 K (m.p.)
Ni6.597.40Solid at 298 K; liquid at 1728 K (m.p.)
Cu7.127.93Solid at 298 K; liquid at 1356 K (m.p.)
Zn9.169.95Solid at 298 K; liquid at 693 K (m.p.)
Ga11.8111.44Solid at 298 K; liquid at 303 K (m.p.)
Ge13.6514.50Solid at 298 K; liquid at 1210 K (m.p.)
As12.9513.76Gray solid at 298 K; liquid at 1090 K (m.p.)
Se16.39N/AGray solid at 298 K; high m.p.
Br19.7825.78Solid at low T; liquid at 298 K
Rb55.7959.20Solid at 298 K; liquid at 312 K (m.p.)
Sr33.9435.12Solid at 298 K; liquid at 1050 K (m.p.)
Y19.88N/ASolid at 298 K; high m.p.
Zr14.0115.20Solid at 298 K; liquid at 2130 K (m.p.)
Nb10.84N/ASolid at 298 K; high m.p.
Mo9.33N/ASolid at 298 K; high m.p.
Ru8.17N/ASolid at 298 K; high m.p.
Rh8.27N/ASolid at 298 K; high m.p.
Pd8.859.70Solid at 298 K; liquid at 1828 K (m.p.)
Ag10.2811.50Solid at 298 K; liquid at 1235 K (m.p.)
Cd13.0014.82Solid at 298 K; liquid at 594 K (m.p.)
In15.7116.60Solid at 298 K; liquid at 430 K (m.p.)
Sn16.2417.10White solid at 298 K; liquid at 505 K (m.p.)
Sb18.1819.00Solid at 298 K; liquid at 904 K (m.p.)
Te20.45N/ASolid at 298 K; high m.p.
I25.69N/ASolid at 298 K; high m.p.
Cs70.7374.90Solid at 298 K; liquid at 302 K (m.p.)
Ba38.1639.40Solid at 298 K; liquid at 1000 K (m.p.)
La22.39N/ASolid at 298 K; high m.p.
Ce20.9521.80Solid at 298 K; liquid at 1071 K (m.p.)
Pr20.80N/ASolid at 298 K; high m.p.
Nd20.58N/ASolid at 298 K; high m.p.
Sm19.98N/ASolid at 298 K; high m.p.
Eu28.9830.10Solid at 298 K; liquid at 1095 K (m.p.)
Molar volume ratios between liquid and solid phases enable predictions of density changes during , which informs processes like where volume expansion can induce stresses, or in for modeling core formation. For example, the ~3% expansion in iron upon contributes to understanding seismic discontinuities in Earth's interior.

References

  1. [1]
  2. [2]
    7.3: Density - Physics LibreTexts
    Mar 12, 2024 · Density is the mass per unit volume of a substance or object. In equation form, density is defined as. ρ = m V . · The SI unit of density is kg / ...
  3. [3]
    Density | Physics - Lumen Learning
    Density is the mass per unit volume of a substance or object. In equation form, density is defined as. ρ = m V . · The SI unit of density is kg/m3.
  4. [4]
    2.4: Density and its Applications - Chemistry LibreTexts
    Mar 24, 2021 · Densities are widely used to identify pure substances and to characterize and estimate the composition of many kinds of mixtures.What is Density? · Solids, liquids and gases · How the temperature affects...
  5. [5]
    Background on Mass, Weight and Density
    Density is mass per volume. Lead is dense, Styrofoam is not. The metric system was designed so that water will have a density of one gram per cubic centimeter ...
  6. [6]
    Lesson 3.1: What is Density? - American Chemical Society
    Jul 29, 2024 · The density of an object is the mass of the object compared to its volume. The equation for density is: Density = mass/volume or D = m/v. Each ...
  7. [7]
    Teaching Density - College of Science - Purdue University
    Density is mass divided by volume. · The simple answer is how heavy something is for its size. · The more technical answer is that density is the physical ...
  8. [8]
    Density as a Tool for Chemistry and Biology
    Density, defined as the ratio of the mass and volume of an object, is a fundamental characteristic of all matter.
  9. [9]
    Unit of Density - BYJU'S
    Density Definition: Density is the measurement of how tightly a material is packed together. It is defined as the mass per unit volume. Density Symbol: D or ρ.
  10. [10]
    Density | NIST - National Institute of Standards and Technology
    Jun 12, 2023 · Density is the mass of a substance per unit volume, usually specified at standard temperature and pressure. Water's density is about 1 gram per ...<|control11|><|separator|>
  11. [11]
    NIST Guide to the SI, Chapter 8
    Jan 28, 2016 · This Guide prefers the name "mass density" for this quantity because there are several different "densities," for example, number density of ...
  12. [12]
    [PDF] BoxSand - Density
    *Note: There are other types of density (e.g. energy density, charge density, number density, area density, etc…). Since we will be working with mass density ...
  13. [13]
    [PDF] Experiment 1 - Density, Measurement, & Error - Chemistry at URI
    Density = Mass ÷ Volume or D = m. V. Any units of mass and volume can be used to define density, but the most frequently used units in ...
  14. [14]
    11.2 Density – College Physics - University of Iowa Pressbooks
    ... in determining whether an object sinks or floats in a fluid. Density is the mass per unit volume of a substance or object. In equation form, density is defined ...
  15. [15]
    [PDF] Introduction to Continuum Mechanics
    This is a set of notes written as part of teaching ME185, an elective senior-year under- graduate course on continuum mechanics in the Department of Mechanical ...
  16. [16]
    [PDF] Fundamentals of Continuum Mechanics
    Nov 3, 2011 · Continuum mechanics is a mathematical framework for studying the transmis- sion of force through and deformation of materials of all types.
  17. [17]
    [PDF] General Relativity
    • Maxwell tensor Fµν. • Stress-energy tensor Tµν. • Current 4-vector Jµ. If ρ is the proper electric density and uµ the 4-velocity of the charge, then. Jµ = ρuµ.
  18. [18]
    SI units - NPL - National Physical Laboratory
    Examples of SI derived units expressed in terms of base units ; density, mass density, kilogram per cubic metre, kg/m ; surface density, kilogram per square metre ...Kilogram (kg) · Metre (m) · Second (s)<|control11|><|separator|>
  19. [19]
    Density, Specific Weight, and Specific Gravity
    Specific weight is defined as weight per unit volume. Weight is a force. The SI unit for specific weight is [N/m 3 ]. The imperial unit is [lb/ft 3 ].
  20. [20]
    Density - A Physical Property
    The densities of solids are usually given in grams per cubic centimeter (g/cm3), the densities of gases in grams per liter (g/L), and the densities of liquids ...
  21. [21]
    Units, dimensions, and conversions
    Units and dimensions ; density. [M L-3]. kilogram per cubic meter. kg m ; force. [M L T-2]. newton. N · kg m s ; pressure. [M L-1 T-2]. pascal. Pa. N m ...
  22. [22]
    NIST Guide to the SI, Appendix B: Conversion Factors
    Feb 1, 2016 · Appendix B provides conversion factors to convert values from non-SI units to SI units, or to units that are accepted for use with SI.
  23. [23]
    Specific Gravity | Definition, Formula and its Connection to Density
    Specific gravity, also known as relative density, measures the density of a substance compared to the density of water.
  24. [24]
    [PDF] temperature corrections to readings of baumé hydrometers, Bureau ...
    There have been a large number of different Baume scales proposed and used since the original scale was devised by Antoine. Baume,2 a French chemist, in 1768.Missing: primary source
  25. [25]
    Baumé Scale | Neutrium
    The Baumé scale was first developed in 1768 as a method of measuring the density of liquids. Today it is largely superseded, however it is still used in some ...Missing: history | Show results with:history
  26. [26]
    Water - Specific Gravity vs. Temperature
    Specific gravity (SG) for water is given for four different reference temperatures (4, 15, 15.6 and 20°C). From 0 to 100°C the pressure is 1 atm, and for ...
  27. [27]
    Mass Density of a Gas - Chemistry 301
    ... molecular weight of a gaseous compound. The density is the mass divided by the volume. Plugging in the volume ( n R T / P ) from the ideal gas law we get.Missing: formula | Show results with:formula
  28. [28]
  29. [29]
    Density Determination of Solids and Liquids - EAG Laboratories
    A reference body with known volume (glass sinker) is weighed once in air and once in the liquid with unknown density (the sample). The density of the liquid can ...
  30. [30]
    Understanding Coriolis flowmeters: Features and benefits | Endress+Hauser
    **Summary of Coriolis Flow Meters Measuring Density:**
  31. [31]
    [PDF] Characterization of Metal Powders Used for Additive Manufacturing
    Sep 16, 2014 · Helium pycnometry [9,10] can be used to measure the true density of the solid backbone of a porous material, as long as all the pores are.
  32. [32]
    Density: Helium Pycnometry - Materials Research Institute
    There are two types of density associated with powders. Envelope (or bulk) density is determined for porous materials when pore space spaces within material ...
  33. [33]
    Density and Buoyancy - HyperPhysics
    Archimedes' principle states that the buoyant force experienced by a submerged object is equal to the weight of the liquid displaced by the object.
  34. [34]
    Density of an Irregular Shape | Physics Van | Illinois
    The density of something is just the mass divided by the volume: D = m/V The mass you can measure on a balance or a scale, and the volume is the amount of ...
  35. [35]
    [PDF] Properties of metal powders for additive manufacturing - GovInfo
    Several methods for determining the apparent density of metal powder are specified by ASTM standards. The apparent density is the ratio of the mass to a given ...
  36. [36]
    [PDF] GRACEnet Sampling Protocols I. Soil Sampling Guidelines*
    Oct 14, 2005 · Record GPS coordinates of each sampling location and collect samples using a core method compositing a minimum of 8 cores per depth. If soil ...
  37. [37]
    [PDF] Influence of soil bulk density methods and rock content
    Jan 4, 2018 · This entailed: (i) using three methods to measure soil bulk density (IR, ND, and SD),. (ii) assessing the effect of the bulk density sampling ...
  38. [38]
    [PDF] Technology Evaluation on Characterization of the Air Void System in ...
    Sep 17, 2009 · The X-ray CT scanning can generate a 3D map of density with respect to location inside the sample. Voids are identified as close-to-zero density ...
  39. [39]
    Porosity characterization using medical computed tomography ...
    Sep 5, 2023 · The CT scans provided nondestructive three-dimensional visualization and quantification of pore distribution, size and volume, pore connections, and estimates ...<|control11|><|separator|>
  40. [40]
    1.3 Thermal Expansion - University Physics Volume 2 | OpenStax
    Oct 6, 2016 · Figure 1.8 This curve shows the density of water as a function of temperature. Note that the thermal expansion at low temperatures is very small ...Linear Thermal Expansion · Example 1.3 · Thermal Stress
  41. [41]
    Water Density | U.S. Geological Survey - USGS.gov
    Water density is roughly 1 gram per milliliter (0.9998395 g/ml at 4°C), but changes with temperature and dissolved substances. Ice is less dense.
  42. [42]
    Air Properties Definitions
    The sea level standard value of air density r is. r = 1.229 kilograms/cubic meters = .00237 slug/cubic feet. When working with a static or unmoving gas, it is ...
  43. [43]
    2.12: Water - Gas, Liquid, and Solid Water - Biology LibreTexts
    Nov 22, 2024 · Ice is less dense than water because the orientation of hydrogen bonds causes molecules to push farther apart, which lowers the density. For ...<|separator|>
  44. [44]
    Changing density ice and liquid water - Physics Forums
    Sep 2, 2015 · At 0°C, the density of ice is 0.9187 g/cm³, while liquid water has a density of 0.9998 g/cm³. An increase in ice density by more than 8.83% ...
  45. [45]
    [PDF] 5. Phase Transitions - DAMTP
    The liquid-gas transition releases latent heat, which means that S = @F/@T is discontinuous. Alternatively, we can say that V = @G/@p is discontinuous.
  46. [46]
    Diamond - HyperPhysics
    Diamond is a crystalline form of carbon which has a density of has a density of 3.51 gm/cm 3 compared to 2.26 gm/cm 3 in the graphite form.
  47. [47]
    14.4A: Graphite and Diamond - Structure and Properties
    Jan 15, 2023 · Graphite has a lower density than diamond. This is because of the relatively large amount of space that is "wasted" between the sheets. Graphite ...Diamond · Graphite · The Bonding in Graphite · Silicon dioxide: SiO2
  48. [48]
    [PDF] Thermophysical Properties of Fluid D2O - Standard Reference Data
    Oct 15, 2009 · At 20°C, its density is 1.1050 kg/dm3 compared to 0.9982 kg/dm3 for H20. In view of the applications of heavy water, many re- search workers ...
  49. [49]
    [PDF] EN380 Naval Materials Science and Engineering Course Notes, US ...
    Example: Calculate the change in density of an iron specimen as it changes from BCC to FCC at. 910o C. The lattice parameter of iron at 20o C is 2.87 Å and the ...
  50. [50]
    BCC to FCC | Harvard Natural Sciences Lecture Demonstrations
    The transition from BCC to FCC results in an 8 to 9% increase in density, causing the iron sample to shrink in size as it is heated above the transition ...
  51. [51]
    Fluid Density in Miscible Displacement - PERM Inc.
    To find the density of a mixture that is assumed as an ideal solution it is enough to calculate the mass and volume of each of the components of the mixture ...
  52. [52]
    1.22.3: Volume: Partial and Apparent Molar
    ### Definition and Formula for Apparent Molar Volume
  53. [53]
    Model for Calculating the Density of Aqueous Electrolyte Solutions
    ### Summary of Model for Calculating Density of Aqueous Electrolyte Solutions
  54. [54]
    6.3 Density – Introduction to Oceanography
    ... salts and other dissolved substances increases surface seawater density to between 1.02 and 1.03 g/cm3. The density of seawater can be increased by reducing ...
  55. [55]
    Precision density and volume contraction measurements of ethanol ...
    May 1, 2013 · Using this capability, volume contraction of ethanol–water binary mixtures were measured with a ∼0.5 pL sample volume at different temperatures.Missing: alcohol- | Show results with:alcohol-
  56. [56]
    [PDF] Thermophysical Properties of Stainless Steels - OSTI.GOV
    Hu119 derived a formula by which the densities of alloys can be cal- culated from their compositions and the density values of their constituent elements: p ...
  57. [57]
    Metals & Alloys - The Chemistry of Art
    419.5. Densities, Selected Alloys. Alloy, Density, g·cm-3. Admiralty Brass, 8.5. Cupronickel, 8.9. Monel, 8.37-8.82. Cast Iron, 6.85-7.75. Carbon Steel, 7.85.
  58. [58]
    Rule-of-Mixture Equation - an overview | ScienceDirect Topics
    ... density of composites, ρc, can be calculated by the following equation: [10.1] ρ c = ρ f V f + ρ m V m. if the densities of fiber, ρf, and matrix, ρm, are ...
  59. [59]
    Fiberglass Composite Material Design Guide
    Density (lb/in3) .055 .065 .02 .10 .30. Fiberglass & polyester. Material Cost $/lb, $2.00-3.00. Strength, yield (psi), 30,000. Stiffness (psi), 1.2 x 10 6.
  60. [60]
    Apparent Porosity - an overview | ScienceDirect Topics
    The % water absorption increased with increasing % apparent porosity and the bulk density increased somewhat for the membrane MB3 in comparison to membrane MB1 ...
  61. [61]
    [PDF] characterizing the influence of print parameters on porosity - OSTI
    To obtain a porosity measurement, the filament's apparent density returned by the helium pycnometry was treated as a maximum. The volumetric density of each ...<|control11|><|separator|>
  62. [62]
    The Influence of Porosity on Mechanical Properties of PUR-Based ...
    Apr 20, 2023 · The work is focused on the mechanical behavior description of porous filled composites that is not based on simulations or exact physical models.
  63. [63]
    Density of Elements Chart - Angstrom Sciences
    Density, Name, Symbol, #. 0.0899 g/L, Hydrogen, H, 1. 0.1785 g/L, Helium, He, 2. 0.9 g/L, Neon, Ne, 10. 1.2506 g/L, Nitrogen, N, 7. 1.429 g/L, Oxygen, O, 8.Missing: CRC | Show results with:CRC
  64. [64]
    Mercury - Element information, properties and uses | Periodic Table
    Density (g cm−3), 13.5336. Atomic number, 80, Relative atomic mass, 200.592. State at 20°C, Liquid, Key isotopes, 202Hg. Electron configuration, [Xe] 4f145d106s ...
  65. [65]
    Density of Wood Species: Data & Material Guide
    Wood densities vary; for example, Alder is 420-680 kg/m³, Apple is 650-850 kg/m³, and Beech is 700-900 kg/m³.
  66. [66]
    Density (Unit weight) of Concrete - Civil Engineering
    The bulk density or unit weight of concrete is the mass or weight of the ... Typical density of concrete 2.3 g/cm3 (143.6 pcf or lb/ft3 / 2300 kg/m3).
  67. [67]
    [PDF] CRC Handbook of Chemistry and Physics, 88th Edition
    Density: Density values for solids and liquids are always in units of grams per cubic centimeter and can be assumed to refer to temperatures near room ...
  68. [68]
    21.1: Periodic Trends and Charge Density - Chemistry LibreTexts
    Jul 12, 2023 · Ionization energies, the magnitude of electron affinities, and electronegativities generally increase from left to right and from bottom to top.<|control11|><|separator|>
  69. [69]
    U.S. Standard Atmosphere: Temperature, Pressure, and Air ...
    A standard atmosphere can be regarded as an average pressure, temperature and air density for various altitudes.Missing: formula | Show results with:formula
  70. [70]
    [PDF] On the barometric formula - Instituto Superior Técnico
    ~8! For our atmosphere and taking T5290 K, this height ~the so-called scale height! is 8.5 km. Such a value was ...
  71. [71]
    Scale height – Knowledge and References - Taylor & Francis
    Notes on the Barometric Formula​​ For our atmosphere with T = 290 K, H = 8.5 km (the so-called scale height). Such a value overlaps with the value estimated by E ...
  72. [72]
    A Table of the Standard Atmosphere to 86 km in SI units
    Apr 18, 2021 · In this table from -2 to 86 km in 2 km intervals. alt is altitude in kilometers. sigma is density divided by sea-level density.
  73. [73]
    Water 4°C density - kg-m3.com
    Mass density of Water 4°C is 999.9749 kg/m3. Water 4°C specific gravity, mass- and volume calculator.
  74. [74]
    Seawater - Properties - The Engineering ToolBox
    Seawater properties like density, saturation pressure, specific heat, electrical conductivity and absolute viscosity. ; 20, 0.02291, 1025 ; 21, 0.02437, 1025 ; 22 ...<|separator|>
  75. [75]
    [PDF] Some Physical Properties of Sea Water or, More than you ever
    For a column of saltwater with a density of 1025 kg/m3 that is 10 m high with a gravity of 9.81 m/s2, using (p1) we get: δP = (1025)(9.81)(10) ~ 105 Pa,.Missing: m³ | Show results with:m³
  76. [76]
    Seawater density - Coastal Wiki
    Feb 22, 2021 · The seawater density \rho in the ocean mainly depends on salinity S, temperature T and pressure p. Pressure only plays a role at depths much greater than those ...Missing: deep | Show results with:deep
  77. [77]
    Water Density, Specific Weight and Thermal Expansion Coefficients
    Data on the density and specific weight of water across various temperatures and pressures. Useful for engineering, fluid dynamics, and HVAC calculations.Missing: relative | Show results with:relative
  78. [78]
    [PDF] 2017 MIT Seawater Properties Table r2
    Properties are listed as a function of temperature and salinity for the pressures, P = P0 , 7 MPa and 12 MPa. Here, P0 = 0.101325 MPa for t <= 100°C and for t > ...
  79. [79]
    Molar Volume | The Elements Handbook at KnowledgeDoor
    Our table of molar volumes has over 130 values covering 97 elements. Each value has a full citation identifying its source. The integrated unit conversion ...
  80. [80]
    Molar Volume of all the elements in the Periodic Table - SchoolMyKids
    This Molar Volume table gives the Molar Volume of all the elements of periodic table in cm3/mol. Click on 'Element Atomic Number', 'Element Symbol', 'Element ...
  81. [81]
    Density | The Elements Handbook at KnowledgeDoor
    Our table of densities has over 750 values covering 96 elements. Each value has a full citation identifying its source. The integrated unit conversion ...Missing: STP | Show results with:STP