Fact-checked by Grok 2 weeks ago

Molar volume

Molar volume is a fundamental in and that describes the volume occupied by one of a substance under defined conditions of and , serving as a key measure for comparing the space taken up by different materials on a per-mole basis. It is formally defined as the ratio of the total volume V of a substance to its n, expressed as V_m = V/n, where the SI unit is cubic meter per mole (m³/mol), though liters per mole (L/mol) or cubic centimeters per mole (cm³/mol) are commonly used in practice. For pure substances, this quantity provides insight into molecular packing and density, while for mixtures, the mean molar volume \bar{V}_m = V / n (where n is the total moles) accounts for the average behavior across components, which is essential in fields like solution chemistry and phase equilibria. In the context of gases, molar volume is particularly significant due to its relation to the ideal gas law PV = nRT, where it equals V_m = RT/P; for an ideal gas at standard temperature and pressure (STP, defined as 0 °C or 273.15 K and 1 atm or 101.325 kPa), the standard molar volume is exactly 22.413 969 54 dm³/mol (or approximately 22.4 L/mol). Under the IUPAC-recommended standard conditions of 0 °C and 1 bar (100 kPa), this value is exactly 22.710 954 64 dm³/mol, reflecting slight variations in pressure standards that impact gas density calculations. Molar volume plays a critical role in applications such as determining substance densities (\rho = M / V_m, where M is ), predicting reaction volumes in , and analyzing real gas deviations via compressibility factors, with SI units preferred for modern conventions. Its measurement and calculation enable precise stoichiometric computations and thermodynamic modeling across solids, liquids, and gases.

Fundamentals

Definition

Molar volume, denoted as V_m, is defined as the volume occupied by one of a substance under specified conditions of and . It is mathematically expressed as V_m = \frac{V}{n}, where V is the total of the substance and n is the in moles. This thermodynamic property provides essential insights into the effective size and packing of molecules or atoms in a , reflecting how closely particles are arranged under given conditions. Molar volume serves as a fundamental in equations of state, which describe the relationships among , , , and composition for substances in various phases, and it contributes to the construction of phase diagrams that map equilibrium boundaries between phases. Molar volumes are commonly reported under standard conditions, such as (STP) defined as 0 °C (273.15 K) and 1 atm (101.325 kPa), to enable consistent comparisons across substances and experiments. The concept originated from Amedeo Avogadro's 1811 hypothesis that equal volumes of different gases at the same temperature and pressure contain equal numbers of molecules, which provided the theoretical foundation for linking volume to the and ultimately led to the formalization of the concept in 1900.

Relation to Specific Volume and Density

The molar volume V_m relates to the v, defined as the volume occupied per unit mass (v = V / m), through the M of the substance, given by the equation V_m = v \cdot M. This connection follows from the definitions: the total volume V = n V_m for n moles, and the total mass m = n M, so v = V / m = (n V_m) / (n M) = V_m / M, rearranging to yield V_m = v M. Density \rho, the mass per unit volume (\rho = m / V), is inversely related to molar volume by \rho = \frac{M}{V_m}. Deriving this, substitute the expressions for mass and volume: \rho = (n M) / (n V_m) = M / V_m. Conversely, molar volume can be computed as V_m = M / \rho. For example, liquid water at 25°C has a molar mass M \approx 18 g/mol and density \rho \approx 1 g/cm³, yielding V_m \approx 18 cm³/mol. The unit for molar volume is cubic meters per (m³/). In practice, chemists often use cm³/ or /, with interconversions such as 1 / = 1000 cm³/ = 10^{-3} m³/, facilitating calculations across scales. These relationships enable engineers to convert between mass-specific properties like and mole-specific ones like molar volume, which is crucial for scaling chemical processes from lab experiments (handling small molar quantities) to industrial production (requiring large volumes based on ).

Gases

Ideal Gases

The molar volume of an ideal gas, denoted as V_m, represents the volume occupied by one mole of the gas under specified conditions and is directly derived from the ideal gas law. The ideal gas law is expressed as PV = nRT, where P is the pressure, V is the volume, n is the number of moles, T is the absolute temperature in Kelvin, and R is the universal gas constant. This empirical equation combines earlier observations, such as Boyle's law (PV constant at fixed T and n), Charles's law (V/T constant at fixed P and n), and Avogadro's law (V/n constant at fixed P and T). Dividing both sides by n yields the molar form: V_m = \frac{RT}{P}. The value of R depends on the units used; common forms include R = 8.314 \, \mathrm{J \cdot mol^{-1} \cdot K^{-1}} for SI units or R = 0.0821 \, \mathrm{L \cdot atm \cdot mol^{-1} \cdot K^{-1}} for pressure in atmospheres and volume in liters. The model relies on key assumptions about the behavior of gas molecules: they are treated as point particles with negligible own volume, exhibit no attractive or repulsive forces except during instantaneous elastic collisions, and undergo constant random motion. These simplifications hold reasonably well for real gases under conditions of low pressure and high temperature, where the average distance between molecules is large compared to their size, minimizing intermolecular interactions. Standard molar volumes are calculated using the equation V_m = \frac{RT}{P} at defined reference conditions. At (STP), defined as 0 °C (273.15 ) and 1 , the molar volume is approximately 22.4 L/mol. To compute this, substitute T = 273.15 \, \mathrm{K}, P = 1 \, \mathrm{atm}, and R = 0.0821 \, \mathrm{L \cdot atm \cdot mol^{-1} \cdot K^{-1}}: V_m = \frac{(0.0821)(273.15)}{1} \approx 22.4 \, \mathrm{L/mol}. At 25 °C (298.15 ) and 1 , the molar volume is approximately 24.5 L/mol. Using R = 0.0821 \, \mathrm{L \cdot atm \cdot mol^{-1} \cdot K^{-1}} and P = 1 \, \mathrm{atm}: V_m = \frac{(0.0821)(298.15)}{1} \approx 24.5 \, \mathrm{L/mol}. As an example, the molar volume of helium—a gas that closely approximates ideal behavior at ambient conditions—is 24.5 L/mol at room temperature (25 °C) and 1 atm pressure, calculated via V_m = \frac{RT}{P} with the values above. This uniformity in molar volume for all ideal gases at the same T and P stems from Avogadro's principle, enabling stoichiometric predictions without gas-specific adjustments.

Real Gases

Real gases exhibit deviations from the primarily due to intermolecular forces and the finite of gas molecules, which become pronounced at high s and low s. Attractive van der Waals forces between molecules reduce the force of collisions with container walls, effectively lowering the observed compared to ideal predictions, while repulsive forces and the physical of molecules exclude a portion of the available , making the gas less compressible than assumed. These effects are negligible under conditions of low and high , where molecular interactions are minimal, but they lead to significant errors in calculations for dense gases. The van der Waals equation addresses these deviations by modifying the ideal gas law to incorporate corrections for molecular attractions and excluded volume: \left(P + \frac{a}{V_m^2}\right)(V_m - b) = RT Here, V_m is the molar volume, a is an empirical constant representing the strength of intermolecular attractions (in units of pressure times volume squared per mole squared), and b is the excluded volume per mole due to the finite size of molecules. This equation, proposed by Johannes Diderik van der Waals in his 1873 doctoral thesis, provides a more accurate estimate of molar volume for real gases by solving for V_m from the rearranged cubic form: V_m^3 - \left(b + \frac{RT}{P}\right)V_m^2 + \frac{a}{P}V_m - \frac{ab}{P} = 0 The roots of this cubic equation yield possible molar volumes, with the largest real root typically corresponding to the gas phase under most conditions; numerical methods are often used for precise solutions. Another key metric for assessing non-ideal behavior is the compressibility factor Z = \frac{P V_m}{RT}, which equals 1 for ideal gases but deviates for real gases—typically Z < 1 when attractions dominate (e.g., near condensation) and Z > 1 when volume exclusion prevails (e.g., at high pressures). These deviations are visualized in generalized compressibility charts, such as the Standing-Katz chart, plotting Z against reduced pressure P_r = P / P_c and reduced temperature T_r = T / T_c, where P_c and T_c are the critical pressure and temperature; this principle of corresponding states allows estimation of Z for many gases using universal curves. For instance, at 300 and 100 , the predicts a molar volume for CO₂ of approximately 0.246 / (V_m = RT/P, with R = 0.0821 L ⁻¹ ⁻¹). Applying the with gas-specific constants a = 3.592 L² ⁻² and b = 0.0427 L ⁻¹ yields a corrected V_m \approx 0.080 / after solving the cubic, illustrating how, near the critical point (T_c = 304 , P_c ≈ 73 ), intermolecular attractions dominate, reducing the molar volume by about 67%.

Liquids

Pure Liquids

Pure liquids exhibit relatively constant molar volumes compared to gases, owing to their low isothermal compressibility, typically on the order of $10^{-4} bar^{-1}, which results in minimal volume changes under moderate pressure variations. This stability arises from the strong intermolecular forces in the liquid state, leading to typical molar volumes ranging from 10 to 100 cm³/mol for common substances. The molar volume of pure liquids shows a pronounced temperature dependence, quantified by the isobaric thermal expansion coefficient \alpha, defined as \alpha = \frac{1}{V_m} \left( \frac{\partial V_m}{\partial T} \right)_P, where V_m is the molar volume and the derivative is taken at constant pressure. For most liquids, \alpha is positive, indicating volume expansion with increasing due to enhanced molecular motion. However, water displays an anomalous behavior, with its molar volume reaching a minimum of approximately 18.0 cm³/ at , below which it contracts upon cooling; this density maximum stems from the restructuring of bonds that temporarily enhances molecular packing efficiency. Molar volumes of pure liquids are commonly measured using pycnometry and dilatometry, techniques that leverage precise volume and mass determinations. Pycnometry involves a density bottle (pycnometer) of known volume to assess liquid density, from which V_m is derived via V_m = M / \rho, where M is the molar mass and \rho is the density. The procedure entails: (1) cleaning and drying the pycnometer, then weighing it empty; (2) filling it completely with water at a known temperature, inserting the stopper, removing air bubbles by tapping, drying the exterior, and weighing the filled pycnometer; (3) calculating the pycnometer volume V using the water's known density \rho_w at that temperature via V = (m_w - m_e) / \rho_w, where m_w and m_e are the masses of the filled and empty pycnometer; (4) emptying, cleaning, and refilling the pycnometer with the sample liquid at the same temperature, weighing it, and computing the sample density \rho = (m_s - m_e) / V, where m_s is the mass of the sample-filled pycnometer. Dilatometry complements this by measuring volumetric changes with temperature, typically using a dilatometer—a bulb connected to a calibrated capillary—filled with the liquid; the liquid level in the capillary is observed at various temperatures to determine expansion, yielding \alpha and thus temperature-dependent V_m. Representative examples illustrate how molecular structure influences molar volumes in pure liquids. Mercury, a non-polar metal with weak van der Waals interactions, has a compact atomic arrangement yielding a molar volume of 14.8 cm³/ at 25°C. In contrast, exhibits a larger molar volume of 58.7 cm³/ at 25°C, reflecting its extended molecular chain and moderate bonding between hydroxyl groups. Water's smaller value of about 18.0 cm³/ at 25°C, compared to similar-sized molecules without bonding, underscores the role of extensive tetrahedral hydrogen-bonded networks in promoting dense packing, though this network partially disrupts above 4°C, leading to expansion.

Liquid Mixtures

In liquid mixtures, the molar volume V_{m,\text{mix}} describes the total volume per mole of the mixture and is influenced by the interactions between components. For ideal mixing, where intermolecular forces between unlike molecules are similar to those between like molecules, the molar volume of the mixture is additive: V_{m,\text{mix}} = \sum x_i V_{m,i}, with x_i denoting the mole fraction of component i and V_{m,i} its pure-component molar volume. This additivity implies no volume change upon mixing, \Delta V_{\text{mix}} = 0, as the partial molar volumes equal the pure molar volumes. Non-ideal mixing arises when interactions differ, leading to an excess molar volume V_m^E = V_{m,\text{mix}} - \sum x_i V_{m,i}, which quantifies deviations from ity. Positive V_m^E often results from that hinder close packing, increasing the overall volume, while negative V_m^E stems from attractive interactions like hydrogen bonding that promote denser packing. For instance, the ethanol- system exhibits negative excess molar volume due to enhanced hydrogen bonding and efficient molecular packing between the and molecules, reducing the volume below the ideal sum. To analyze these behaviors, partial molar volumes are calculated as V_{m,i} = \left( \frac{\partial V}{\partial n_i} \right)_{T,P,n_j}, representing the volume change upon adding one of component i at constant T, P, and moles n_j of other components. Consistency across a is ensured by the Gibbs-Duhem relation, \sum x_i d\mu_i = 0 at constant T and P, where \mu_i is the ; for volumes, it implies \sum x_i dV_{m,i} = 0, allowing determination of one partial molar volume from experimental data on others. These properties find applications in , such as predicting phase in where excess volumes affect vapor-liquid equilibria, and in solubility assessments where partial molar volumes influence how solutes partition in solvents under . The benzene-toluene serves as a classic example of nearly ideal , with V_m \approx 89 cm³/ for benzene-rich compositions and minimal excess volume to similar aromatic structures.

Solids

Crystalline Solids

In crystalline solids, the molar volume is precisely determined from the atomic-scale arrangement in the periodic lattice, typically via to obtain parameters. The molar volume V_m is calculated using the relation V_m = \frac{N_A V_\text{cell}}{Z}, where N_A is Avogadro's constant ($6.022 \times 10^{23} mol^{-1} ), V_\text{cell} is the volume of the , and Z is the number of units per . This formula directly links the macroscopic property to the microscopic structure, enabling accurate predictions of volume based on lattice geometry. The molar volume in crystalline solids shows only minor variation with temperature, primarily due to of the . Linear thermal expansion coefficients are typically on the order of $10^{-5} K^{-1} , leading to volume expansion coefficients around $3 \times 10^{-5} K^{-1} . For instance, in , the linear coefficient is approximately $4.4 \times 10^{-5} K^{-1} . In non-cubic crystals, this expansion is anisotropic, with different coefficients along principal axes, reflecting the directional bonding in the . Representative examples illustrate how structure dictates molar volume. adopts the rock salt structure, a face-centered cubic lattice with Z = 4 and edge length of 5.64 , yielding V_m \approx 27.0 cm³/ and reflecting the ionic packing of Na^+ and Cl^- ions. In contrast, exhibits the structure, with Z = 8 carbon atoms per of volume 0.0454 nm³, resulting in a highly compact V_m \approx 3.4 cm³/ due to strong covalent tetrahedral bonding. Several factors govern the molar volume in crystalline solids. Ionic radii directly scale the interatomic distances and thus the unit cell volume; larger radii, as in crystals with bigger cations or anions, increase V_m while maintaining similar coordination. The , determined by the radius ratio of cations to anions, influences packing density—higher coordination often accommodates larger volumes per ion to satisfy electrostatic balance, as per . Polymorphism further modulates volume through distinct lattice arrangements; for , the rhombohedral polymorph has V_m \approx 36.9 cm³/mol, while orthorhombic is denser at \approx 34.1 cm³/mol, affecting stability under varying pressure conditions.

Amorphous Solids

Amorphous solids exhibit higher molar volumes compared to their crystalline counterparts primarily due to structural and the presence of excess volume resulting from inefficient atomic or molecular packing during the process. This volume arises from the rapid quenching of the melt, which prevents the achievement of the more efficient ordered arrangement found in crystals. For instance, fused silica, an amorphous form of SiO₂, has a of 2.20 g/cm³, yielding a molar volume of approximately 27.3 cm³/ (calculated from the of 60.08 g/), while crystalline α-quartz has a of 2.65 g/cm³ and a molar volume of about 22.7 cm³/. This difference, typically on the order of 10-20% higher volume for amorphous phases, underscores the impact of on packing across various amorphous materials. The glass transition temperature (T_g) significantly influences the molar volume behavior in amorphous solids, marking the point where molecular mobility increases, allowing structural relaxation and a shift in characteristics. Below T_g, the material behaves as a rigid with a low coefficient of , while above T_g, it enters a rubbery or supercooled state with higher expansion, leading to greater volume changes upon heating. In silica , T_g occurs at approximately 1200°C, where the expansion transitions from the glassy regime to that resembling the supercooled , affecting applications requiring thermal stability. This transition highlights how molar volume in amorphous solids is not fixed but evolves with temperature, influenced by the of structural rearrangement. Molar volumes in amorphous solids are measured using techniques adapted for their non-periodic structure, such as dilatometry, which monitors dimensional changes to determine volume expansion while accounting for long relaxation times near T_g that can cause time-dependent volume recovery. Complementary density measurements often employ scattering to analyze the , providing insights into local packing and free volume without relying on long-range order. These methods reveal subtle differences in molar volume attributable to processing conditions or composition, ensuring accurate characterization distinct from crystalline solids. Representative examples illustrate these principles in diverse amorphous systems. In polymers like , the molar volume of the C₈H₈ is approximately 95-99 cm³/mol, depending on ; atactic polystyrene, with random , exhibits a lower (∼1.05 g/cm³) and thus higher molar volume than isotactic polystyrene (∼1.11 g/cm³), due to less efficient chain packing in the disordered configuration. Similarly, metallic glasses, such as Zr-based alloys, display molar volumes 1-2% higher than their crystalline equivalents owing to disorder, though their short-range enables packing densities approaching 85-90% of close-packed structures, contributing to unique mechanical properties.

References

  1. [1]
    [PDF] Quantities, Units and Symbols in Physical Chemistry - IUPAC
    ... defined for a mixture as m/n and V/n, where n = Pi ni. These quantities are called the mean molar mass and the mean molar volume respectively. (7) These ...
  2. [2]
    NIST Guide to the SI, Chapter 8
    Jan 28, 2016 · Definition: volume of a substance divided by its amount of substance: Vm = V/n. Notes: 1. The word "molar" means "divided by amount of substance ...
  3. [3]
    [PDF] The gas laws pV = nRT
    How to use the perfect gas equation to calculate the molar volume of a perfect gas at any temperature and pressure? V = nRT p. V m = V n. =.
  4. [4]
    Determination of partial molar volumes from free energy perturbation ...
    Jan 15, 2015 · Partial molar volume is an important thermodynamic property that gives insights into molecular size and intermolecular interactions in solution.
  5. [5]
    Equation of state to predict the solid-liquid-vapor phase ...
    In this case, the pressure P of the fluid is a function of the molar volume v, and the temperature T, while R is the gas constant (= 83.14472 cm3 bar/Kelvinmol) ...
  6. [6]
    How Was Avogadro's Number Determined? - Scientific American
    Feb 16, 2004 · Avogadro is most famous for his hypothesis that equal volumes of different gases at the same temperature and pressure contain the same ...
  7. [7]
    Amedeo Avogadro - Science History Institute
    Avogadro's Hypothesis​​ In 1811 Avogadro hypothesized that equal volumes of gases at the same temperature and pressure contain equal numbers of molecules.
  8. [8]
    Molar volume – Knowledge and References - Taylor & Francis
    Molar volume is the volume occupied by 1 mole of a substance under standard state conditions (25°C, 1 atm). Typical units are L/mol, cm3/mol, m3/mol, m3/kmol.
  9. [9]
    Molar Volume - an overview | ScienceDirect Topics
    Molar volume is defined as the volume occupied by one mole of a compound or element at a specified temperature and pressure, calculated using the formula \( V_ ...
  10. [10]
    Water | H2O | CID 962 - PubChem - NIH
    Pure Water Density Standard, UKAS ISO/IEC17025 and ISO Guide 34 certified, density: 0.9982 g/mL at 20 C, density: 0.9970 g/mL at 25 C · Water,bacteriostatic ...
  11. [11]
  12. [12]
    Fundamental Assumptions
    First, the volume of the molecules is assumed to be negligible. Second, the molecules are assumed not to exert forces on one another, except when they collide.
  13. [13]
    Deviations from Ideal Gas Law Behavior:
    The assumption that there is no force of attraction between gas particles cannot be true. If it was, gases would never condense to form liquids. In reality, ...
  14. [14]
    Ideal Gas Law - Chemistry 301
    As all gases that are behaving ideally have the same number density, they will all have the same molar volume. At STP this will be 22.4 L. This is useful if you ...Missing: definition | Show results with:definition
  15. [15]
  16. [16]
    7.1: Volume Properties - Chemistry LibreTexts
    Apr 12, 2022 · Two volume properties of a closed system are defined as follows: (7.1.1) cubic expansion coefficient α = def 1 V ⁢ ( ∂ V ∂ T ) p ...
  17. [17]
    Water Density, Specific Weight and Thermal Expansion Coefficients
    Water density is mass/volume, highest at 4°C (1000 kg/m³). Specific weight is weight/volume. Density depends on temperature and pressure.Missing: molar | Show results with:molar
  18. [18]
    [PDF] Hydrogen Bonding in Water - Esalq
    Water's hydrogen bonding holds water molecules up to about 15% closer than if than if water was a simple liquid with just van der Waals dispersion interactions.
  19. [19]
    Pycnometer - Chemical Engineering | University of Utah
    Feb 24, 2008 · Determine the exact volume of your pycnometer by filling it fully with water, inserting the stopper, and tapping the sides gently to remove the ...
  20. [20]
    [PDF] Volume dilatometry - NIST Technical Series Publications
    The level of the liquid in the capillary is measured when the dilatometer is at atmospheric pressure. Then, at the same temperature, a vacuum is ap- plied to ...
  21. [21]
    Technical data for the element Mercury in the Periodic Table
    Technical data for Mercury ; Bulk physical properties ; Density, 13.534 g/cm ; Density (Liquid), 13.534 g/cm ; Molar Volume, 0.0000148213 ; Brinell Hardness, N/A.
  22. [22]
    Ethanol - Thermophysical properties - The Engineering ToolBox
    Values are given for liquid at 25 oC (77 oF, 298 K) and 1 bara, if not other phase, temperature or pressure given. For full table with Imperial Units - rotate ...
  23. [23]
    A Flexible Mixing Rule Satisfying the Ideal-Solution Limit for ...
    From eq 1 it is clear that for an ideal solution the partial molar volume of component q, υqid.sol, is equal to the pure compound volume υq at any pressure ...
  24. [24]
    Excess molar properties and hydrogen bonding in ethanolamine ...
    By analyzing molecular properties and interactions—such as hydrogen bonding, excess molar volume, and dynamic viscosity—valuable insights can be gained into how ...
  25. [25]
    Molar excess volumes of binary liquid mixtures: 2-pyrrolidinone with ...
    The observed changeover VE < 0 to VE > 0 suggests that the steric effect is primarily responsible for the positive contributions to VE. ... Key words: excess ...
  26. [26]
    Structural transformations, composition anomalies and a dramatic ...
    Mar 29, 2012 · Water-ethanol mixtures exhibit many interesting anomalies, such as negative excess partial molar volume of ethanol, excess sound absorption ...
  27. [27]
    [PDF] Simple mixtures
    to derive an expression for the partial molar volume of water in solution. The molar volume of pure water is 18.079 cm3 mol-1. A: K2SO4. B: H2O nAdVA + nBdVB ...Missing: typical | Show results with:typical
  28. [28]
    Excess molar volumes, densities and P–T–x phase equilibria for ...
    The excess molar volumes of the liquid phase were determined and the excess molar volumes behavior indicates that three mixtures of benzene + toluene, heptane ...
  29. [29]
    [PDF] partial molar volume lab
    Thus, if the partial molar volume of solute in aqueous solution is greater than the molar volume of solid solute, an increase in pressure will increase the ...
  30. [30]
    [PDF] Chemistry 440 Problem set 5 - Oregon State University
    The partial molar volumes of two liquids A and B in a mixture in which the mole ... Benzene and toluene form nearly ideal solutions. Consider an equi-molar ...
  31. [31]
    Sodium Chloride NaCl - Optical Materials - Crystran
    Density, 2.17 g/cc ; Melting Point, 801°C ; Thermal Conductivity, 6.5 W m-1 K-1 @ 273 K ; Thermal Expansion, 44 x 10-6 K-1 ; Hardness, Knoop 18.2 in with 200g ...
  32. [32]
    Behavior of the Thermal Expansion of NaCl at Elevated Temperatures
    The coefficient of linear thermal expansion (α) for NaCl has been determined from 50°C to near the melting point (800°C) with an automatic recording dilatometer ...
  33. [33]
    mp-22862: NaCl (Cubic, Fm-3m, 225) - Materials Project
    NaCl is Halite, Rock Salt structured and crystallizes in the cubic Fm̅3m space group. Na¹⁺ is bonded to six equivalent Cl¹⁻ atoms to form a mixture of corner ...Missing: molar | Show results with:molar
  34. [34]
    [PDF] Molar Volume, Ionic Radii in Stoichiometric and Nonstoichiometric ...
    Apr 11, 2012 · One of the important issues is a change in spheres occupied by ions (ionic radii) dependent on the interaction character, coordination number, ...Missing: factors | Show results with:factors
  35. [35]
    Ionic Coordination and Pauling's Rules - Tulane University
    Sep 26, 2014 · The radii increase with increasing total number of electrons downward in table. Ionic radius also increases with increasing coordination number ...
  36. [36]
    [PDF] Molar volumes and densities of minerals
    These tables present critically chosen "best values" for the density and molar volume of selected mineral compounds. No attempt was made to be all-inclusive; ...
  37. [37]
    Fused Silica Material Properties - Translume
    Fused Silica is a light material. Its density is 2.202g/cm3 (To a very small degree, the density is influenced by the annealing conditions of the glass near the ...
  38. [38]
    Quartz: Mineral information, data and localities.
    Density: 2.65 - 2.66 g/cm3 (Measured) 2.66 g/cm3 (Calculated). Optical Data of QuartzHide. This section is currently hidden. Click the show button to view ...
  39. [39]
    What should the density of amorphous solids be? - AIP Publishing
    Nov 20, 2019 · A survey of published literature reveals a difference in the density of amorphous and crystalline solids (organic and inorganic) on the order of 10%–15%.
  40. [40]
    Properties of fused silica - Heraeus Covantics
    For fused quartz the amount is approximately 20 ppm and consists primarily of Al2O3 with much smaller amounts of alkalis, Fe2O3, TiO2, MgO and ZrO2.
  41. [41]
    Dilatometry - an overview | ScienceDirect Topics
    Dilatometry measures volume changes of materials with temperature, used for fat melting, solid fat content, and polymer glass transitions.
  42. [42]
    On Structure and Properties of Amorphous Materials - PMC
    Free volume can be measured by X-ray diffraction methods, as well as by dilatometric measurements. When applied to glassy materials, X-ray and dilatation ...
  43. [43]
    [PDF] V. Polystyrene - Standard Reference Data
    This corresponds to a packing fraction of 0.63. For atactic polystyrene values of 1.04 to 1.065 Mg m -3 have been reported for the density [7].
  44. [44]
    Density change upon crystallization of amorphous Zr–Cu–Al thin films
    It is found that, in general, the density change decreases from Cu-/Al-rich compositions to Zr-rich compositions, and the previously reported good glass-forming ...
  45. [45]
    Why is the density of glassy metals higher than that of crystalline ...
    Nov 30, 2014 · Metallic glasses present lower densities than their cristalline counterparts. I can even be characterized by a parameter called free volume.