Fact-checked by Grok 2 weeks ago

Number density

Number density, denoted as n, is a fundamental in physics and chemistry that quantifies the concentration of particles—such as atoms, molecules, ions, or subatomic particles—per unit volume in a substance, gas, , or space. It is distinct from mass density, focusing instead on counting entities rather than their , and is essential for describing the microscopic structure of in fields ranging from kinetic theory to . In gases and fluids, number density is closely related to the , where the total number density of air molecules n_a at (e.g., 0°C and 1013 ) is approximately $2.69 \times 10^{19} molecules per cubic centimeter, enabling calculations of partial pressures and mixing ratios for specific gases like CO₂. For a specific X, it is given by n_X = C_X \cdot n_a, where C_X is the volume mixing ratio (e.g., in parts per million by volume), highlighting its role in and kinetics. In , number density N for atoms in materials is calculated as N = \frac{\rho N_A}{M}, with \rho as mass density, N_A as Avogadro's number, and M as , supporting applications in fuel depletion and interaction rates. The unit for number density is cubic meters inverse (\mathrm{m}^{-3}), though practical units like molecules per cubic centimeter (\mathrm{cm}^{-3}) are common in and atmospheric contexts due to their convenience for high densities near Earth's surface, where values reach about $2.7 \times 10^{19} \, \mathrm{cm}^{-3}. In kinetic theory, it underpins derivations of as P = \frac{1}{3} n m \langle v^2 \rangle, linking macroscopic properties like to microscopic particle motions and velocities. Across disciplines, variations in number density influence phenomena such as behavior, stellar interiors, and doping, where precise control of particle counts per volume determines electrical and .

Fundamentals

Definition

Number density, often denoted by the symbol n, is a fundamental quantity in physics that quantifies the concentration of particles within a given volume. It is mathematically defined as the ratio of the total number of particles N to the volume V occupied by them, expressed as n = \frac{N}{V}. This measure provides an indication of how densely particles are packed in a system, serving as a key parameter in various physical analyses. The term "particles" in this context broadly encompasses atoms, molecules, ions, and, in specialized fields such as plasma physics, subatomic entities like electrons and ions. For instance, in plasmas, the number density accounts for both charged species, enabling the description of collective behaviors in ionized gases. This inclusive definition allows number density to apply across different states of matter, including gases, liquids, solids, and plasmas, where it characterizes the volumetric distribution of constituents. In multicomponent systems or mixtures, the total number density represents the sum of the partial number densities of each individual . The partial number density for a specific component is simply the number of that type of particle per unit volume, facilitating the of compositional variations within the . This distinction is particularly useful in heterogeneous environments, such as atmospheric or chemical systems.

Units and dimensions

The International System of Units (SI) defines number density as having the unit of reciprocal cubic metre (m⁻³), which quantifies the number of particles per unit volume. This unit arises from the base SI unit of length, the metre, raised to the power of negative three, ensuring coherence with other derived quantities in the system. Common alternative units for number density include cubic centimetre inverse (cm⁻³), widely used in plasma physics and microscopy due to the scale of samples involved, and litre inverse (L⁻¹), prevalent in chemical and biological applications for solution concentrations. Conversion between these units follows volume scaling factors; for instance, 1 cm⁻³ equals 10⁶ m⁻³, as one cubic metre contains 10⁶ cubic centimetres. In dimensional analysis, the formula for number density is = L⁻³, where L denotes , emphasizing its role as an inverse measure independent of or other properties. Particle counting conventions typically express the total particle count N in a V as n = N/V, with implications for linking to macroscopic scales via Avogadro's constant (N_A = 6.02214076 \times 10^{23} \, \mathrm{mol}^{-1}), which facilitates conversions to molar-based quantities without altering the fundamental particle-per-volume definition.

Physical and Chemical Relations

Relation to mass density

The mass density \rho of a system is defined as the total mass M divided by the volume V, yielding \rho = M / V. This total mass arises from the sum of the masses of all particles within the volume, expressed as M = N \langle m \rangle, where N is the total number of particles and \langle m \rangle is the average mass per particle. Substituting this into the density definition gives \rho = (N / V) \langle m \rangle = n \langle m \rangle, where n is the number density. This equation demonstrates the direct proportionality between mass density and number density, scaled by the average particle mass. For pure substances composed of identical particles, \langle m \rangle reduces to a constant mass m per particle, simplifying the relation to \rho = n m. In mixtures, however, \langle m \rangle serves as a weighted , calculated as \langle m \rangle = \sum_i (n_i m_i) / n, where n_i and m_i are the number density and mass of the i-th component, respectively, and n = \sum_i n_i. This average reflects the compositional makeup, allowing the relation to account for varying particle types without altering the fundamental proportionality. In non-uniform systems, such as fluids with spatial gradients in composition or particle distribution, variations in \rho can stem from either changes in n or \langle m \rangle, or both, influencing local properties like pressure and flow behavior. This interplay underscores the utility of the relation in modeling heterogeneous media. The conceptual linkage between number density and mass density emerged in the mid-19th century through the foundational work on kinetic theory of gases, where James Clerk Maxwell employed the number of molecules per unit volume to connect microscopic particle motions to macroscopic density.

Relation to molar concentration

The molar concentration c, a measure of the amount of substance per unit volume, is directly related to the number density n through Avogadro's constant N_A. Specifically, c = \frac{n}{N_A}, where N_A = 6.02214076 \times 10^{23} mol^{-1} is the number of particles in one mole of substance. This relation scales the microscopic count of individual particles to the of moles, facilitating connections between physical and chemical descriptions of . The inverse relationship follows as n = c N_A, allowing conversion from chemical concentrations to particle densities. This derivation arises from the definition of molar concentration as moles per volume: if N is the total number of particles in volume V, the number of moles is N / N_A, so c = \frac{N / N_A}{V} = \frac{n}{N_A}, since n = N / V. In the SI system, molar concentration has units of mol m^{-3}, though in chemical practice, it is often expressed as mol L^{-1} (denoted as molarity or ). While number density is typically employed in physics to quantify particle distributions for applications like kinetic theory and processes, molar concentration is the standard in chemistry for expressing solution compositions in reactions and equilibria. This distinction underscores how the two quantities provide equivalent yet contextually tailored insights into molecular-scale arrangements.

Relation to other composition measures

In mixtures, the partial number density n_i of component i is defined as the product of its x_i and the total number density n, expressed as n_i = x_i n. This relation holds under the assumption that the is composed of distinguishable , with x_i representing the of total moles attributed to component i. For gaseous mixtures, this is equivalent to the partial number density being the mixing ratio times the total air number density. The number density also connects to volume fractions in mixtures through partial molar volumes. The volume fraction \phi_i of component i approximates \phi_i \approx (n_i v_i) / \sum_j (n_j v_j), where v_i is the partial molar volume per particle (obtained by dividing the partial molar volume \bar{V}_i by Avogadro's number). This approximation applies particularly to dilute solutions, where interactions between solute molecules are negligible, and the total volume is dominated by the . In such cases, deviations from ideality can lead to excess volumes, but the formula provides a estimate for analysis. In multiphase systems, an effective number density is obtained via volume averaging over the phases present. The effective n is given by n = \sum_k \alpha_k n_k, where \alpha_k is the volume fraction of phase k and n_k its intrinsic number density. This averaging technique derives from ensemble or spatial averaging methods used in modeling, ensuring conservation of particle number across heterogeneous media. Within , the total number density n serves as a key parameter in , characterizing the average particle occupancy in . The phase-space N(\mathbf{x}, \mathbf{p}, t), which gives the number density of particles per unit volume in position and , integrates over to yield the spatial number density n(\mathbf{x}, t) = \int N d^3p. This parameterization facilitates the derivation of equilibrium distributions, such as Maxwell-Boltzmann, by normalizing probabilities based on n. These relations rely on assumptions of ideality, where molecular interactions do not alter partial volumes beyond simple additivity, and additivity, positing that the total volume equals the sum of component volumes without contraction or expansion upon mixing. In non-ideal mixtures, such as those with strong intermolecular forces, these assumptions fail, leading to excess molar volumes and requiring corrections via activity coefficients or virial expansions.

Applications Across Disciplines

In gases and kinetic theory

In the context of gases, number density plays a central role in the , which describes the behavior of dilute gases under equilibrium conditions. The law is expressed as PV = NkT, where P is , V is , N is the total number of molecules, T is , and k is the (k = 1.38 \times 10^{-23} J/K). Dividing by volume yields P = n k T, where n = N/V is the number density, directly linking to the molecular concentration at fixed . This relation holds for low-density gases where intermolecular interactions are negligible, providing a foundational equation in kinetic theory. Number density is essential in deriving transport properties, such as the , which represents the average distance a travels between collisions. In kinetic theory for hard-sphere molecules, the mean free path \lambda is given by \lambda = \frac{1}{\sqrt{2} \pi d^2 n}, where d is the molecular . This formula arises from considering the relative motion of molecules, with the \sqrt{2} factor accounting for collisions between identical particles and \pi d^2 as the effective collision cross-section. As number density increases, \lambda decreases inversely, leading to more frequent collisions and shorter paths, which governs the transition from ballistic to diffusive motion in gases. The further informs rates of and . For through a small , the number of molecules effusing per unit time per unit area is J = \frac{1}{4} n \bar{v}, where \bar{v} = \sqrt{\frac{8 k T}{\pi m}} is the average molecular speed and m is the ; this rate is thus proportional to number density, explaining where effusion speed scales inversely with \sqrt{m} at constant (and thus constant n). Similarly, the self-diffusion coefficient D in gases is approximated as D \approx \frac{1}{3} \lambda \bar{v}, substituting the mean free path expression yields D \propto 1/n, indicating that diffusion slows with increasing molecular concentration due to heightened . For non-ideal gases at higher densities, number density appears in the of the equation of state, which corrects the law for intermolecular forces. The is expanded as \frac{P}{k T} = n + B_2(T) n^2 + B_3(T) n^3 + \cdots, where B_i(T) are the virial coefficients depending on , with B_2 capturing pairwise interactions (positive for repulsive forces, negative for attractive). This series converges at low densities (n a^3 \ll 1, where a is a molecular scale) and provides insights into deviations from ideality, such as those near critical points. Experimentally, number density in gases is determined using the by measuring and , as n = P / ([k](/page/K) T), often with gauges and thermometers for low-density systems. This assumes ideality but can be refined with virial corrections for denser conditions, enabling precise characterization in applications like .

In liquids and solids

In liquids, the number density n typically ranges around $10^{28} m^{-3} for molecules, as exemplified by liquid at standard conditions where n \approx 3.34 \times 10^{28} m^{-3}. This value reflects the close packing of molecules in the condensed phase, with packing efficiencies approaching 0.64 for random close-packed structures similar to those in simple liquids modeled as . Coordination numbers in liquids average around 10 to 12 nearest neighbors, influencing local density and structural correlations that maintain overall uniformity despite the lack of long-range order. In solids, number density is determined from the crystal lattice parameters, providing a precise measure of atomic or molecular arrangement. For a face-centered cubic (FCC) structure, common in metals like aluminum and , n = 4 / a^3, where a is the and the factor of 4 accounts for the atoms contributed per . This yields densities on the order of $10^{28} to $10^{29} m^{-3}, depending on the material; for example, in with a \approx 3.61 \times 10^{-10} m, n \approx 8.47 \times 10^{28} m^{-3}. Such calculations highlight the high packing of 74% in FCC lattices, maximizing atomic occupancy within the unit cell volume./10:_Liquids_and_Solids/10.07:_Lattice_Structures_in_Crystalline_Solids) Number density in both liquids and varies with and due to and effects. As increases, the volume expands, reducing n by approximately 0.1% to 1% per for most materials, with liquids showing greater sensitivity than . Elevated conversely compresses the material, increasing n linearly with applied stress, as seen in where rises from 1000 m^{-3} at 1 to about 1045 m^{-3} at 1000 near . These changes are critical for understanding phase stability and material behavior under environmental stresses. Amorphous solids exhibit number densities 10% to 15% lower than their crystalline counterparts due to disordered atomic packing, which introduces voids and reduces overall efficiency compared to the periodic lattices in crystals. For instance, amorphous silica has a density about 12% below that of quartz, reflecting less efficient space filling despite similar short-range coordination. This density deficit influences mechanical properties, such as lower rigidity, and is a key distinction in applications like glasses versus ceramics. Crystalline forms, by contrast, achieve higher densities through optimized long-range order. Atomic number density in solids is commonly measured using X-ray diffraction (), which probes lattice spacings to determine volumes and thus n. By analyzing diffraction peaks via , researchers extract the a, enabling direct computation of from the known number of atoms per ; this technique achieves atomic-scale precision, often resolving variations to within 0.1%. is particularly valuable for polycrystalline samples, providing averaged structural data without requiring single crystals.

In astrophysics and plasmas

In astrophysics, number density plays a crucial role in characterizing the extreme conditions within stellar interiors and the diffuse . In the core of , the mass density reaches approximately 150 g/cm³, leading to a particle number density on the order of $10^{32} m^{-3} when divided by the proton mass for fully ionized , reflecting the high compression under gravitational forces. In contrast, the interstellar medium (ISM) exhibits much lower densities, with typical values for neutral atomic around 1 cm^{-3} (or $10^6 m^{-3}) in the diffuse phase, varying by region due to temperature and ionization states. In plasmas, which dominate many astrophysical environments such as stellar atmospheres and the ISM's ionized components, quasineutrality ensures that the electron number density n_e approximately equals the ion number density n_i (or more precisely n_e = \langle Z_i \rangle n_i accounting for levels), maintaining overall charge balance over scales larger than the . This condition arises from the of charged particles, where rapid adjustments via prevent significant charge separation in quasineutral regimes. The further incorporates number density to describe the ionization balance in plasmas, relating the ratios of successive ionization states—such as n_{r+1} n_e / n_r—to temperature, n_e, and partition functions, enabling predictions of composition in astrophysical contexts like H II regions. This accounts for the statistical distribution of particles, with n_e directly influencing the ionization fraction under local . Observational estimates of number density in astrophysical plasmas often rely on , where line intensity ratios from forbidden transitions—such as those in planetary nebulae—provide diagnostics for electron densities along the , assuming geometric models to convert column densities to local values. For instance, the ratio of emission lines sensitive to collisional de-excitation yields n_e estimates in the range of $10^3 to $10^6 cm^{-3} for nebular plasmas. In cosmological models, the density n_b parameterizes the abundance of ordinary matter, with current values derived from data yielding \Omega_b h^2 \approx 0.0224, corresponding to n_b \approx 0.25 m^{-3} today, conserved since and influencing . This density is probed through light element abundances and informs the baryon-photon ratio \eta \approx 6 \times 10^{-10}, linking early universe physics to large-scale cosmic evolution.

Examples and Calculations

Illustrative examples

To illustrate the concept of number density, consider everyday and specialized examples across different states of and environments. These values highlight the vast of molecular or atomic concentrations, from dense solids to sparse cosmic gases. In dry air at (, defined as 0°C and 1 ), the number density of molecules is approximately 2.7 × 10^{25} m^{-3}. This value arises from the , reflecting the dilute nature of gases compared to condensed . Liquid at has a number density of about 3.3 × 10^{28} molecules per cubic meter, calculated from its mass of 1000 kg m^{-3} and of 18 g ^{-1}. This high underscores the close packing in liquids. In a , the atomic number density is roughly 5 × 10^{28} atoms per cubic meter (or 5 × 10^{22} cm^{-3}), determined by the structure with a of 0.543 nm. In the diffuse interstellar medium, neutral atoms typically have a number density of about 1 atom cm^{-3} (equivalent to 10^6 m^{-3}), representing the low-density gas filling much of the space between . The following table summarizes these illustrative values, drawing from standard :
Material/EnvironmentNumber DensityUnitsSource/Reference
Air at STP (dry)2.7 × 10^{25}m^{-3} Atmospheric Sciences
Liquid (25°C)3.3 × 10^{28}m^{-3} Physics
crystal5 × 10^{28}m^{-3} EE Course Notes
Interstellar (diffuse )1cm^{-3}UCSD Center for Astrophysics & Space Sciences

Computational methods

One common computational method for determining number density in gases is through the equation of state for ideal gases, where the number density n is calculated as n = \frac{P}{kT}, with P denoting , k Boltzmann's constant, and T . This approach assumes non-interacting particles and is widely applied in thermodynamic modeling of dilute gases. In , () simulations provide detailed number profiles by tracking particle positions over time in atomistic models. These simulations solve Newton's for interacting atoms, allowing computation of local densities \rho(z) along spatial coordinates, such as in lipid bilayers or solid-liquid interfaces. For instance, can reveal density variations in complex systems like aluminum-gallium interfaces, where profiles are averaged over trajectories to obtain distributions. Experimentally, number density can be derived from scattering techniques, particularly scattering, which relates the measured differential cross-sections to atomic scattering lengths and sample composition. In (SANS), the scattering length density \rho_b, proportional to number density times the average scattering length, is fitted from data to infer n. This method is effective for condensed matter, as cross-sections depend on interactions, enabling precise determination in materials with varying isotopic compositions. Uncertainty in number density arises primarily from errors in measuring particle count N or V, as n = N / V, and propagates via relative uncertainties \frac{\Delta n}{n} = \sqrt{ \left( \frac{\Delta N}{N} \right)^2 + \left( \frac{\Delta V}{V} \right)^2 }. In particle counting experiments, statistics dominate for low counts, while volume measurements introduce systematic errors from calibration or geometry. Forward error propagation techniques quantify these, ensuring reliable estimates in applications like aerosol analysis. Software tools facilitate these computations; for example, LAMMPS (Large-scale Atomic/Molecular Massively Parallel Simulator) outputs number density through commands like compute chunk/atom for binning atoms and calculating local n from counts divided by bin volumes. This open-source package supports MD simulations across scales, integrating density profiles into thermodynamic analyses.

References

  1. [1]
    Number Density | BoxSand – Flip the Classroom
    Number Density ... The number of atoms or molecules per unit volume. Kinetic Theory of Gases · Physics Dictionary.<|control11|><|separator|>
  2. [2]
    [PDF] CHAPTER 1. MEASURES OF ATMOSPHERIC COMPOSITION
    mol/mol). 1.2 NUMBER DENSITY. The number density nX of a gas X is defined as the number of molecules of X per unit volume of air. It is expressed commonly in.
  3. [3]
    [PDF] 22.05 Reactor Physics – Part Three Useful Tools - DSpace@MIT
    Number Density: Many nuclear engineering calculations require knowledge of the number of atoms per unit volume. Examples are fuel depletion and reaction rate ...
  4. [4]
    ATMO336 - Fall 2010 - atmo.arizona.edu
    The number density of a gas is defined as the number of gas molecules per unit volume. In Earth's atmosphere, near sea level there are about 2.7x1019 molecules ...
  5. [5]
    [PDF] Kinetic Theory - Macmillan Learning
    ... kinetic energy: P = 1. 3. N. V m1v22av = 2. 3. N. V a. 1. 2 mv2 bav. KT-3 where N>V is called the number density. This result shows that kinetic theory pre-.
  6. [6]
    Solar Physics Glossary - NASA
    The amount of mass or number of particles per unit volume. In cgs units mass density has units of gm cm-3. Number density has units cm-3 (particles per ...
  7. [7]
    [PDF] Basic Plasma Physics - DESCANSO
    This means that the ion and electron densities are nearly equal, ni ne , a condition commonly termed “quasi-neutrality.” This condition exists throughout the ...
  8. [8]
    [PDF] Properties of Gases - The University of Texas at Dallas
    There are four major phases of matter: solids, liquids, gases and plasmas. ... The number density of the molecules (i.e the number of molecules per unit.
  9. [9]
    [PDF] Approximate Thermodynamic State Relations in Partially Ionized ...
    Jun 10, 2004 · is ρe⫽mene , where ne is the partial number density of free electrons in the mixture. Since the particle masses are known, the partial mass ...
  10. [10]
    Number Density - an overview | ScienceDirect Topics
    Number density is defined as the distribution of particles, such as nuclei or electrons, normalized to the average values over a given volume, measured as a ...
  11. [11]
    Density - The Physics Hypertextbook
    Mass density equations. space, algebraic, calculus, SI unit. volume ... number density, especially in this book… particle density, see current and ...
  12. [12]
    Mass, Weight, Density
    Density is defined as mass per unit volume. Data can be entered into any of the boxes below. Specifying any two of the quantities determines the third.
  13. [13]
    [PDF] EQUATION OF STATE
    The number density of particles with all momenta, contained within a unit volume, is n = ∞. Z0 n (p)dp, ρ = nm. (3) where ρ is the mass density of gas. Note ...
  14. [14]
    Nuclear Notation
    If the density ρ of the material is known, then the number of nuclei per unit volume n can be calculated from n =ρNA/A.
  15. [15]
    [PDF] Mean Molecular Mass
    Free Electron Number Density. For a fully ionized gas the number density of free electrons, ne, can be written as ne = 100+. ∑ j=1 jnj,. (16) since each atom ...Missing: relation | Show results with:relation
  16. [16]
    Atom - Kinetic Theory, Gases, Particles - Britannica
    Oct 27, 2025 · He derived the basic equation of kinetic theory, which reads P = NMV2. Here P is the pressure of a volume of gas, N is the number of molecules ...
  17. [17]
    NIST Guide to the SI, Chapter 8
    Jan 28, 2016 · ... definition. This implies that the numerical value of a given ... number density of particles, n = N / V; and charge density, ρ = Q ...
  18. [18]
    Molarity and number density - Nexus Wiki - ComPADRE
    May 18, 2021 · The "number density" in chemistry is molarity: the number of moles of a particular chemical in a given volume.
  19. [19]
    Virial coefficients, thermodynamic properties, and fluid-fluid ...
    May 25, 2010 · ... unit diameter, x i = ρ i / ρ is the mole fraction of species i ⁠, ρ i being the partial number density of particles of species i ⁠, and ...
  20. [20]
    [PDF] Density and partial molar volumes of the liquid mixture water + ...
    In case of the mixture water + methanol + propan-2-ol, the location of the excess molar volume at approximately xH2O = 0.55 does not show any significant.
  21. [21]
    [PDF] Ensemble Averaged Conservation Equations For Multiphase, Multi ...
    The bulk average density corresponds intuitively to the idea of the mass of phase k per unit volume of mixture, or the observed material density.
  22. [22]
    [PDF] Chapter 3: Kinetic Theory [version 1203.1.K] - Caltech PMA
    In kinetic theory, the key concept is the “distribution function” or “number density of particles in phase space”, N; i.e., the number of particles of some ...
  23. [23]
    Density and Refractive Index of Binary Ionic Liquid Mixtures ... - MDPI
    Dec 5, 2022 · Therefore, it is a very accurate assumption to assume an ideal mixture and apply a volume additivity rule.Missing: limitations | Show results with:limitations
  24. [24]
    [PDF] Ideal gas law - atmo.arizona.edu
    Aug 24, 2010 · ... number density of the gas in molecules per cubic meter and T is the temperature in Kelvin and kB is Boltzmann's constant = 1.38x10-23 J/K.
  25. [25]
    13.3 The Ideal Gas Law – College Physics - UCF Pressbooks
    A mole is the number of atoms in a 12-g sample of carbon-12. The number of molecules in a mole is called Avogadro's number \boldsymbol{N_{\textbf{A}}},. \ ...Missing: concentration | Show results with:concentration
  26. [26]
    18. The Kinetic Theory of Gases - Home Page of Frank LH Wolfs
    The mean free path of a molecule is related to its size; the larger its size the shorter its mean free path. The mean square speed of the molecules can be ...
  27. [27]
    The Feynman Lectures on Physics Vol. I Ch. 43: Diffusion - Caltech
    This distance between collisions is usually called the mean free path: Mean free path l=τv. ... By na(x,y,z) we mean the number density of special ...
  28. [28]
    [PDF] Chapter 3 3.8 Mean Free Path and Diffusion
    The number density of particles in this volume is n: n = N. V. = N. Nλσ. = 1 ... same as the number density in the sample as a whole. We can rewrite this ...
  29. [29]
    [PDF] Interacting Systems - Physics Courses
    The virial expansion is typically applied to low-density systems. When the density is high, i.e. when na3 ∼ 1, where a is a typical molecular or atomic ...Missing: non- | Show results with:non-
  30. [30]
    [PDF] V.C The Second Virial Coefficient & van der Waals Equation - MIT
    This ratio is roughly 10−3 for air at room temperature and pressure. The corrections to ideal gas behavior are thus small at low densities. On dimensional ...
  31. [31]
    Measuring Gas Density | Physics Van | Illinois
    Oct 22, 2007 · ... Ideal Gas Law' or PV=nRT. In ... To use this equation, all you have to do is measure the pressure of the gas and measure the temperature.
  32. [32]
    Number Density - Chemistry 301
    Number density can be thought of as the number of particles that are present in a particular volume.Missing: definition | Show results with:definition
  33. [33]
    [PDF] Basic understanding of condensed phases of matter via packing ...
    Jul 10, 2018 · The packing fraction of a periodic packing is given by φ = Nv1. Vol(F). = ρv1,. (2) where ρ = N/Vol(F) is the number density, i.e., the number ...
  34. [34]
    Thermal Expansion of Solids and Liquids | Physics - Lumen Learning
    Thermal expansion is clearly related to temperature change. The greater the temperature change, the more a bimetallic strip will bend.
  35. [35]
    Liquids - Densities vs. Pressure and Temperature Change
    The density of a liquid will change with temperature and pressure. The density of water versus temperature and pressure is indicated below.
  36. [36]
    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%.
  37. [37]
    [PDF] Introduction to Solid State Chemistry Lecture Notes No. 5 X-RAYS ...
    X-ray diffraction in crystalline solids takes place because the atomic spacings are in the 10–10 m range, as are the wavelengths of X-rays. 5. DIFFRACTION AND ...
  38. [38]
    Our Sun: Facts - NASA Science
    The density of the Sun's core is about 150 grams per cubic centimeter (g/cm³). That is approximately 8 times the density of gold (19.3 g/cm³) or 13 times ...Missing: typical | Show results with:typical
  39. [39]
    The Interstellar Medium - Astronomy & Astrophysics
    Apr 26, 1999 · These regions have a typical temperature of about 100K and a density between 10--100 atoms/cm3. Surrounding the clouds is a warmer lower density ...
  40. [40]
    Quasineutral plasma models | Phys. Rev. E
    Feb 4, 2005 · Hence, while quasineutrality can apply throughout ion-ion plasmas ... (a) Density of positive ions (dashed curve) and electrons (solid curve).
  41. [41]
    [PDF] Introduction to Plasma Physics
    The first of these, 'quasi-neutrality', is actually just a mathematical way of saying that even though the particles making up a plasma consist of free ...
  42. [42]
    [PDF] IONIZATION, SAHA EQUATION
    In these equations the degree of ionization should be treated as a function of density and tem- perature, x(ρ, T). Differentiating equation (i.4) we obtain.
  43. [43]
    [PDF] Saha Equation Normalized to Total Atomic Number Density - arXiv
    The Saha equation describes the relative number density of consecutive ionization levels of a given atomic species under conditions of thermodynamic.
  44. [44]
    Exercise 3: Determining the Gas Density in Planetary Nebulae
    In this exercise you will learn how and why a particular pair of sulfur emission lines is useful for determining density in planetary nebulae.The Sulfur Lines · Collisional Excitation And... · Determining The Density<|separator|>
  45. [45]
    7 Spectral Lines‣ Essential Radio Astronomy
    The number density nn of atoms in the nth electronic energy level is given by the Saha equation, a generalization of the Boltzmann equation (for a ...7.2 Recombination Lines · 7.3 Line Radiative Transfer · 7.7 Molecular Line Spectra
  46. [46]
    Baryonic Density - Nasa Lambda
    Baryonic density is determined using CMB and deuterium abundance, with only 4-5% of the universe comprised of baryons. Recent CMB and D/H+BBN determinations ...
  47. [47]
    Big-Bang Nucleosynthesis and the Baryon Density of the Universe
    Jul 1, 1994 · This measurement of the density of ordinary matter is crucial to almost every aspect of cosmology and is pivotal to the establishment of two ...
  48. [48]
    Atmo336 - Fall 2012 - atmo.arizona.edu
    Air density can be defined as the number of air molecules per unit volume (number density). Near sea level there are about 2.7x1019 molecules per cm3(cubic ...Missing: STP | Show results with:STP
  49. [49]
    [PDF] The properties of water - Galileo
    ρ = 1 gm ⁄ cm3 = 103 kg ⁄ m3 . Since the molecular weight of water (ignoring isotopes like 18O and 2H) is 18, the number density of molecules in liquid water is.
  50. [50]
    [PDF] Silicon
    Feb 9, 2019 · Atomic Number. [Ne]3s23p2. Electron Configuration. Diamond Crystal ... • Silicon atom density. 5x1022 atoms/cm3. Page 11. 11. © tj. EE 4981.
  51. [51]
    The Ideal Gas Law | Physics - Lumen Learning
    The ideal gas law states that PV = NkT, where P is the absolute pressure of a gas, V is the volume it occupies, N is the number of atoms and molecules in the ...Missing: source | Show results with:source
  52. [52]
    12.5: Kinetic Theory - Physics LibreTexts
    Nov 5, 2020 · The fraction n=N/V is the number density of the gas. This is a first non-trivial result of the kinetic theory because it relates pressure (a ...
  53. [53]
    Determining Structural and Mechanical Properties from Molecular ...
    Aug 6, 2014 · Molecular dynamics (MD) simulations are commonly used to study the structure (number density ρ(z) and area per lipid AL) and bending ...Figure 2 · Determining Vesicle... · Molecular Dynamics...
  54. [54]
    Characterization of the Al-Ga solid-liquid interface using classical ...
    Apr 24, 2020 · We present a detailed characterization of the structure and transport properties of aluminum–gallium solid–liquid interfaces using classical molecular-dynamics ...<|separator|>
  55. [55]
    Neutron Scattering Lengths and Cross Sections - ScienceDirect.com
    In this article we examine the procedures to determine total cross sections from neutron transmission experiments performed at the time-of-flight ...
  56. [56]
    [PDF] Primary Determination of Particle Number Concentration with Light ...
    Jan 11, 2018 · Fundamentally, number concentration (or equivalently, number density) is obtained by dividing particle count by the sample volume. In its ...
  57. [57]
    Tutorial: Guide to error propagation for particle counting ...
    Forward error propagation is an established technique for uncertainty quantification (UQ). This article covers practical applications of forward error ...
  58. [58]
    10.3.1. Output from LAMMPS (thermo, dumps, computes, fixes ...
    This table summarizes the various commands that can be used for generating output from LAMMPS. Each command produces output data of some kind and/or writes ...Missing: software | Show results with:software
  59. [59]
    LAMMPS - a flexible simulation tool for particle-based materials ...
    For models with non-uniform density, the number of particles per processor can be load-imbalanced with the default partitioning. This reduces parallel ...<|control11|><|separator|>