Fact-checked by Grok 2 weeks ago

Free molecular flow

Free molecular flow is a regime of rarefied gas dynamics in which the mean free path of gas molecules greatly exceeds the characteristic dimensions of the container or flow path, resulting in negligible intermolecular collisions and dominant interactions between molecules and solid surfaces. This flow regime is characterized by a Knudsen number (Kn), defined as the ratio of the mean free path to the representative length scale, exceeding 10 (Kn > 10). Under these conditions, gas molecules travel in nearly straight-line trajectories between wall collisions, often assuming diffuse reflection where molecules re-emit from surfaces following a cosine distribution. The concept originated from the work of Danish physicist in the early , who investigated gas flow through narrow tubes at low pressures between 1909 and 1910, establishing foundational models for molecular and transmission probabilities. In free molecular flow, transport properties such as conductance in vacuum systems become independent of pressure, depending instead on geometry and temperature, with applications spanning vacuum technology, where it governs pumping efficiency in high-vacuum environments (pressures below approximately 10^{-3} mbar), to hypersonic during re-entry at altitudes above 90 km. Key characteristics include the absence of viscous effects, enabling analytical solutions via kinetic theory, such as the Knudsen formula for through orifices, and the emergence of phenomena like the Knudsen minimum in transitional flows near tube entrances. In practical contexts, free molecular flow influences in rarefied environments, where energy accommodation coefficients determine surface heating, and it underpins simulations in and nanoscale channels. Modern extensions, including models, refine predictions for advanced applications like propulsion and .

Fundamentals

Definition and Characteristics

Free molecular flow is a regime of rarefied gas dynamics in which the of gas molecules significantly exceeds the characteristic dimension of the enclosure or flow path, such that molecules predominantly collide with container walls rather than with one another. In this setup, the gas behaves as a collection of independent particles moving freely without substantial mutual interference, leading to a highly non-continuum mechanism. Key characteristics include straight-line molecular trajectories punctuated only by wall collisions, negligible inter-molecular interactions, and the absence of bulk viscous effects or due to the low gas density. Surface phenomena, such as adsorption, desorption, and , govern the overall behavior, rendering traditional hydrodynamic descriptions inapplicable. The flow lacks a coherent macroscopic , with particles diffusing randomly based on wall interactions, and effective pressures are sufficiently low that viscous transfer is effectively zero. This regime typically arises in high and environments, at pressures below $10^{-3} mbar for centimeter-scale systems, where the becomes comparable to or larger than the system's dimensions. It is commonly observed in enclosed volumes like chambers or narrow channels, such as those in particle accelerators or components, where gas ensures wall-dominated transport. Here, the throughput or depends on molecular impingement rates on surfaces, rather than pressure-driven .

Knudsen Number and Flow Regimes

The (Kn) is a defined as the ratio of the λ of gas molecules to a scale L of the system, such as the of a tube or the size of a chamber:
\text{Kn} = \frac{\lambda}{L}.
This number quantifies the degree of in a gas and determines the applicable physical model for analysis.
Gas flow regimes are classified based on the value of Kn as follows: continuum flow for Kn < 0.001, where the gas behaves as a continuous medium and the Navier-Stokes equations with no-slip boundary conditions are valid; slip flow for 0.001 < Kn < 0.1, characterized by velocity slip at walls but still largely continuum-like behavior; transitional (or Knudsen) flow for 0.1 < Kn < 10; and free molecular flow for Kn > 10, where molecules travel independently without significant intermolecular collisions. The transitional regime, also known as Knudsen flow, represents an intermediate state between viscous () and free molecular flows, where both intermolecular collisions and wall collisions play significant roles, leading to partial effects that invalidate pure assumptions. In this regime, modeling often requires statistical or hybrid approaches, as the is comparable to the system dimensions, resulting in a complex interplay of collision frequencies. The Knudsen number is influenced by factors including gas pressure (inversely proportional to λ via density), temperature (affecting molecular velocity and thus λ), gas type (via molecular collision cross-section), and system geometry (defining L). For example, in air at room temperature (20°C), the free molecular regime (Kn > 10) is typically reached at pressures below approximately 10^{-3} mbar for a characteristic length of 1 cm, such as a tube diameter.

Kinetic Theory Basis

Mean Free Path Calculation

The mean free path, denoted as \lambda, is defined as the average distance traveled by a gas molecule between successive collisions with other molecules in the gas. In kinetic theory, the derivation of \lambda begins by considering the collision frequency z, which is the number of collisions a molecule experiences per unit time. The average speed of the molecules is \bar{v}, and the time between collisions is $1/z, so \lambda = \bar{v} / z. The collision frequency arises from the relative motion of molecules: for hard spheres of diameter d, the effective collision cross-section is \pi d^2, and accounting for the relative velocity factor of \sqrt{2} (due to the random directions in a Maxwell-Boltzmann distribution), z = \sqrt{2} \, \pi d^2 \, n \, \bar{v}, where n is the number density of molecules. Substituting yields the standard expression \lambda = \frac{1}{\sqrt{2} \, \pi d^2 n}. This formula was first derived by James Clerk Maxwell in his foundational work on the dynamical theory of gases.%20-%20Illustrations%20of%20the%20dynamical%20theory%20of%20gases.pdf) The n relates to macroscopic variables via the : n = [P](/page/Pressure) / ([k](/page/K) [T](/page/Temperature)), where [P](/page/Pressure) is the , [k](/page/K) is Boltzmann's constant, and [T](/page/Temperature) is the . Thus, \lambda = \frac{[k](/page/K) [T](/page/Temperature)}{\sqrt{2} \, \pi d^2 [P](/page/Pressure)}. This shows that \lambda is inversely proportional to [P](/page/Pressure) and to the square of the molecular d, while it increases linearly with [T](/page/Temperature) (i.e., \lambda \propto [T](/page/Temperature) / [P](/page/Pressure)). Smaller molecules or lower pressures result in mean free paths, as collisions become less frequent. For practical estimation, consider air (modeled as molecules with d \approx 0.37 ) at 20°C (293 K) and 1 (101.3 kPa): \lambda \approx 68 . In conditions, such as P = 1 , \lambda scales to approximately 6.8 mm, and at levels (P < 10^{-3} Pa), it exceeds millimeters, highlighting its rapid increase with decreasing pressure. The derivation assumes a hard-sphere model for molecules, where collisions are elastic and instantaneous, and velocities follow the Maxwell-Boltzmann distribution for an ideal gas at thermal equilibrium. All molecules are identical in size and mass, with collisions occurring randomly without external forces.

Wall Collision Assumptions

In free molecular flow, the interaction of gas molecules with the walls is modeled under the assumption of diffuse reflection, where molecules are re-emitted from the surface with a velocity distribution that follows . This law posits that the flux of re-emitted molecules is proportional to \cos \theta, where \theta is the angle between the molecule's velocity vector and the surface normal, leading to a half-Maxwellian distribution at the wall temperature. The thermal accommodation coefficient, introduced by Maxwell, quantifies the fraction of incident molecular energy that is transferred to the wall upon collision, with the remainder reflected specularly or diffusely without full equilibration. For most engineering surfaces in vacuum systems, such as metals like silver or platinum, this coefficient ranges from 0.8 to 1.0, indicating near-complete thermalization during interactions. Free molecular flow models predominantly assume diffuse reflection for realistic, rough surfaces, where molecules lose memory of their incident trajectory and are re-emitted isotropically according to the cosine law; specular reflection, in which molecules bounce off with angle of incidence equaling angle of reflection, is considered only for idealized smooth surfaces like clean crystals. This diffuse assumption implies random re-emission directions, resulting in no net tangential momentum transfer to the wall and influencing overall flow resistance by randomizing molecular paths. Experimental validations in vacuum systems confirm these assumptions through measurements of sticking probability, which represents the fraction of incident molecules that adsorb rather than reflect, often approaching 1 for reactive gases on clean metals, and thermal accommodation, where heat flux data between parallel plates in the free molecular regime yields coefficients consistent with diffuse models for pressures below 0.1 Pa. In this regime, the mean free path exceeds system dimensions, making wall collisions the primary interaction mechanism.

Modeling and Equations

Conductance Formulas

In free molecular flow, conductance C is defined as the volume flow rate Q (throughput in pressure-volume units per unit time) divided by the pressure difference \Delta P across the component, with typical units of liters per second (L/s). This quantity characterizes the ease with which gas molecules pass through a vacuum system element under conditions where intermolecular collisions are negligible. For a simple aperture or orifice of area A, the conductance derives from the kinetic theory of gases, specifically the rate at which molecules impinge on a surface. The impingement rate on one side is \frac{1}{4} n \bar{v}, where n is the number density and \bar{v} is the average molecular speed. The net molecular flux through the orifice is thus \frac{1}{4} \bar{v} (n_1 - n_2) A, and since \Delta P = (n_1 - n_2) k T (assuming isothermal conditions), the throughput Q = \frac{1}{4} \bar{v} A \Delta P, yielding C = \frac{1}{4} \bar{v} A. Here, \bar{v} = \sqrt{\frac{8 k T}{\pi m}} = \sqrt{\frac{8 R T}{\pi M}}, with k the , T the temperature, m the molecular mass, R the , and M the molar mass. This assumes a transmission probability of unity, as all incident molecules pass through without wall interactions. For a long circular tube with diameter d and length L \gg d, the Knudsen formula provides the conductance as C = \frac{1}{3} \frac{d^3}{L} \sqrt{\frac{\pi R T}{2 M}}. This approximation accounts for multiple wall collisions, derived by integrating the cosine emission law over successive reflections, assuming diffuse scattering. The formula reduces the effective flow compared to an orifice due to back-scattering. In general, conductance for arbitrary geometries can be expressed using the transmission probability \alpha, the fraction of molecules entering the entrance that exit the other end: C = \frac{1}{4} \bar{v} A \alpha, where A is the entrance area. For the long tube, \alpha \approx \frac{4 d}{3 L}, consistent with the . This framework extends the orifice case by incorporating geometry-specific probabilities computed via or analytical approximations for simple shapes. These formulas assume no intermolecular collisions, relying on wall collision models such as cosine re-emission, and are valid for Knudsen numbers \mathrm{Kn} > 10. Below this threshold, transitional flow effects require modified equations like the full Knudsen interpolation.

Effusion and Transmission Probability

Knudsen effusion describes the process in free molecular flow where gas molecules escape through a small into a , under conditions where the λ greatly exceeds the aperture diameter. This regime ensures negligible intermolecular collisions, so molecules travel ballistically from the source to the aperture. The effusion rate J, representing the number of molecules effusing per unit time through an aperture of area A, is given by J = \frac{1}{4} n \bar{v} A, where n is the molecular number density in the source volume and \bar{v} is the average molecular speed. This expression arises from the kinetic theory flux of molecules incident on a surface, applied directly to the aperture as the effective "wall" area. For two isotopic species with molar masses M_1 and M_2, the ratio of effusion rates follows Graham's law, J_1 / J_2 = \sqrt{M_2 / M_1}, since \bar{v} \propto 1 / \sqrt{M} while n relates to partial pressure via the ideal gas law. The transmission probability α quantifies the efficiency of molecular transport through a channel or conduit in free molecular flow, defined as the fraction of molecules entering from one end that exit the other end without returning to the entrance. For an ideal with negligible thickness, α = 1, as all incident molecules pass through unimpeded. In contrast, for a cylindrical of d and length L >> d, Clausing derived α ≈ 4d / (3L), reflecting the increased likelihood of collisions and back-scattering with tube elongation. This approximation holds for long tubes where direct transmission dominates but is adjusted by higher-order terms for shorter geometries. Derivation of α typically assumes diffuse reflection at walls, where re-emitted molecules follow a cosine distribution relative to the surface , mimicking random thermal re-accommodation. Clausing's seminal approach formulated the problem as a over molecular trajectories, solving for the exit flux distribution numerically for various L/d ratios. Modern computations employ methods, simulating ensembles of molecular paths with random cosine-emitted directions after each wall collision, integrating probabilities to yield α; for example, in bent tubes or constrictions, α decreases due to shadowed regions and multiple reflections, often falling to 0.1–0.5 for 90° bends with L/d ≈ 10. These path-tracing techniques enable extensions to complex geometries while assuming behavior. In measurement applications, Knudsen effusion cells exploit these principles to determine vapor pressures of low-volatility materials. A sample is enclosed in a with a small , heated to , and the effusion rate is monitored via (Δm/Δt) over time, yielding P = \sqrt{\frac{2 \pi R T}{M}} \cdot \frac{\Delta m / \Delta t}{A} after correcting for the . Alternatively, gauges or mass spectrometers detect the effused beam intensity downstream, providing absolute from the flux equation. This method achieves accuracies of 5–10% for pressures down to 10^{-2} , with cells often using or for high-temperature stability up to 2000 . Extensions to multi-component gases treat each independently in effusion, as collisions are absent; the total flux sums partial J_i = (1/4) n_i \bar{v}_i A, enabling selective enrichment ratios per without cross-interference. For non-ideal surfaces, deviations from perfect —characterized by accommodation coefficients σ < 1—introduce partial specular components, increasing α by reducing trapping in channels; generalized models incorporate reflection kernels to compute effective probabilities, with σ ≈ 0.8–1.0 typical for metals in .

History

Early Work by Knudsen

(1871–1949), a Danish at the , laid the groundwork for free molecular flow theory through his experimental and theoretical studies on rarefied gas dynamics between 1909 and 1915. His work focused on gas behavior at low pressures where intermolecular collisions are negligible compared to wall interactions, addressing limitations in earlier viscous flow models that assumed continuum assumptions. Building on James Clerk Maxwell's from the 1860s, which described molecular motions statistically, Knudsen extended these ideas to practical flow scenarios in confined geometries. In his 1909 publication "Die Gesetze der Molekularströmung und der inneren Reibungsströmung der Gase durch Röhren," Knudsen analyzed gas flow through long, narrow tubes, establishing distinct regimes based on the ratio of the to the tube diameter—a criterion that defined the transition to free molecular flow and later formalized as the . He performed experiments using tubes with diameters on the order of 0.1 mm and mercury manometers to precisely measure pressure differences and flow rates at pressures below 0.1 , confirming that in this wall-dominated regime, flow rates were proportional to pressure differences and inversely to tube length, independent of gas . These setups allowed Knudsen to verify theoretical predictions by varying tube dimensions and gas types, such as air and , under isothermal conditions. Knudsen's 1909 paper "Die Molekularströmung der Gase durch Offnungen und die Effusion" introduced the effusion method for measuring vapor pressures by quantifying the steady-state escape of gas molecules through small orifices into , applying it initially to mercury vapor. In 1915, he further developed the cosine law for molecular desorption and from surfaces, assuming random, diffuse that follows a cosine relative to the surface normal, which became a cornerstone assumption for modeling wall interactions in free molecular flow. Early applications of his findings included interpreting forces in low-pressure bulbs, where thermal transpiration drives motion due to molecular bombardment differences, and calibrating gauges like the -type gauge he devised around 1910.

Developments in Vacuum Science

In the 1930s and 1940s, refinements to vacuum pumping technologies for molecular flow regimes were advanced by researchers building on earlier principles, with Wolfgang Gaede's molecular drag pump serving as a foundational design that enabled higher pumping speeds in low-pressure environments. These developments addressed the need for efficient gas removal in molecular flow conditions, where mean free paths exceed system dimensions, by optimizing rotor-stator interactions to impart to rarefied gases. By the 1950s, the introduction of turbomolecular pumps marked a significant milestone; patented in 1958 by W. Becker at Pfeiffer Vacuum, these multistage devices achieved compression ratios up to 10^4 for while maintaining high-vacuum levels below 10^{-6} mbar without oil backstreaming. This innovation facilitated sustained molecular flow operations in laboratory and industrial high-vacuum systems, revolutionizing applications requiring ultra-clean environments. The spurred extensive studies in rarefied gas dynamics, as hypersonic re-entry vehicles encountered free molecular flow regimes at altitudes above 80 km, where Knudsen numbers exceeded 10. NASA's contributions included pioneering models for hypersonic rarefied flows, such as axisymmetric simulations of gas past spheres and blunt bodies, which integrated free molecular assumptions with effects to predict drag and coefficients accurate within 5-10% of flight data. These efforts, documented in technical reports from the Ames and Research Centers, extended analytical frameworks to non-equilibrium conditions, influencing and during the Apollo era. Computational advances in the 1970s introduced simulations for free molecular flow, with the (DSMC) method, originated by in the late 1960s, gaining prominence for handling complex geometries where analytical solutions failed. DSMC probabilistically models particle trajectories and collisions, enabling predictions of conductance and in irregular ducts, and became essential for simulating transitional flows beyond pure molecular regimes. Post-2000 research has focused on nanoscale flows in and reservoirs, where channel dimensions of 10-100 nm yield Knudsen numbers greater than 1, amplifying free molecular effects like slip and . In , multi-scale models incorporating alongside viscous flow account for effects like tortuosity, aiding extraction efficiency estimates. Recent advancements include a 2023 generalized Knudsen theory by Qian, Wu, and Wang, which unifies specular and diffuse wall reflections via an accommodation coefficient parameter, refining predictions for in simulations across boundary conditions. Key milestones encompass the 1986 review by W. Steckelmacher, which synthesized 75 years of progress in Knudsen flow, highlighting empirical conductance data for tube networks and the shift toward computational validation. Additionally, integrations of the have advanced non-equilibrium modeling of free molecular flow, with direct numerical solutions resolving velocity distribution functions in structures and wall interactions, achieving rates of O(1/N) for N particles in one-dimensional cases.

Applications

High Vacuum Systems

In high vacuum systems, free molecular governs the behavior of gas molecules when the exceeds the system dimensions, typically at pressures below 10^{-3} mbar, enabling efficient evacuation without molecular collisions. Turbomolecular pumps exploit this regime by using high-speed rotating blades to impart directional momentum to gas molecules, achieving pumping speeds from 50 to 3000 l/s for and ultimate pressures as low as 10^{-10} mbar when combined with and metal seals. Cryopumps, operating in the same molecular conditions from 10^{-3} to 10^{-12} , capture gases through cryogenic and adsorption on surfaces cooled to 20 or below, routinely attaining pressures below 10^{-9} mbar in unbaked systems and even lower with bake-out procedures. However, conductance limitations in connecting pipes and apertures restrict overall system performance, necessitating careful component sizing based on molecular transmission probabilities. Leak detection in these systems relies on molecular flow models to predict gas ingress, with mass spectrometer detectors achieving sensitivities down to 10^{-12} mbar l/s by exploiting the unimpeded travel of tracer molecules through leaks. from chamber walls, dominated by adsorbed like , contributes significantly to residual gas loads; molecular flow simulations forecast pressure rises from these sources, guiding mitigation strategies such as bake-out at 120-200°C to desorb surface layers and reduce rates by orders of magnitude. System components like baffles and traps are designed to minimize backflow in the molecular regime, where transmission probabilities determine gas throughput—for instance, louvered cryosurfaces in cryopumps yield probabilities of 0.2-0.4 for and , optimizing capture while blocking . In particle accelerators such as the (LHC), liquid -cooled baffles and helium-cooled beam screens in beam pipes use these principles to shield pumps from and maintain , with molecular flow analysis ensuring pressure uniformity along kilometer-scale sections. Design considerations for large chambers emphasize laws in free molecular flow, where conductance for cylindrical ducts scales as C \propto d^3 / L (with d as and L as ), allowing prediction of gradients that can vary by factors of 10 or more across extended volumes without intermediate pumping. For example, in accelerator beam pipes, this scaling informs distributed pumping layouts to counteract gradients induced by distributed gas loads. Challenges in include virtual leaks, where trapped gas volumes in welds or blind holes release slowly through low-conductance paths, mimicking real leaks and limiting base pressures to 10^{-8} mbar or higher until resolved via venting holes or redesign. Surface effects, such as and , exacerbate at extreme low pressures, requiring non-evaporable getter coatings or repeated bake-outs to achieve stable conditions below 10^{-10} mbar.

Separation Processes

Free molecular flow plays a crucial role in processes, particularly through in porous barriers, where the regime approaches molecular flow conditions. During the , uranium enrichment for utilized large-scale plants that operated near the molecular flow regime, exploiting the slight difference in rates of and isotopes in gas. The separation factor \alpha is given by \alpha = \sqrt{M_2 / M_1}, where M_1 and M_2 are the molecular masses of the lighter and heavier isotopes, respectively, yielding \alpha \approx 1.0043 for UF_6. This effusion-based mechanism allows incremental enrichment across multiple stages, though the overall process requires thousands of cascades due to the small \alpha. In porous media, dominates under free molecular flow conditions, enabling selective gas transport in membrane separation technologies. The effective diffusivity D_K is expressed as D_K = \frac{d_p}{3} \sqrt{\frac{8 R T}{\pi M}}, where d_p is the pore diameter, R is the , T is the , and M is the . This mechanism separates gases based on inverse square root mass dependence, with lighter molecules diffusing faster through pores comparable to the . Applications include recovery and purification, where low-pressure operation maintains the Knudsen regime for efficient sieving without bulk flow interference. Molecular distillation leverages free molecular flow for purifying heat-sensitive compounds, such as pharmaceuticals and essential oils, using short-path evaporators. In these systems, the operating is reduced to ensure the exceeds the distance between the evaporator and condenser, typically 1–10 cm, allowing molecules to travel ballistically without collisions. This minimizes and enables high-purity at temperatures below 200°C. Practical examples highlight the utility of these processes. isotope separation via Knudsen or pumps exploits mass differences for applications like , where enriched ^3He/^4He ratios aid in dating and mantle-derived samples. Similarly, cryogenic at low pressures incorporates Knudsen in zeolite membranes to preferentially permeate oxygen or , supplementing traditional for small-scale or energy-efficient operations. Despite these advantages, free molecular flow separation processes suffer from low throughput compared to viscous flow methods, as transport is limited by wall collisions and rates rather than convective bulk motion. also declines in the transitional , where increasing leads to intermolecular collisions that reduce selectivity and require hybrid modeling for accurate prediction.

Other Engineering Applications

In , free molecular flow governs the interaction between and the rarefied upper atmosphere at altitudes exceeding 100 km, where the exceeds 10 and molecular collisions are negligible compared to wall interactions. This is critical for calculating aerodynamic and on reentry vehicles and , as the of gas molecules surpasses vehicle dimensions, leading to and of incident molecules on surfaces. For thrusters, such as free molecule micro-resistojets, the flow expands through microchannels under low- conditions, enabling low-thrust propulsion for attitude control and orbit adjustments in small without traditional nozzles, achieving specific impulses suitable for missions under 50 kg. In nanotechnology, free molecular flow facilitates gas transport within micro-electro-mechanical systems (MEMS) and nanoporous structures, where channel dimensions approach the molecular mean free path. Knudsen pumps exemplify this application, operating as no-moving-parts micropumps that exploit thermal transpiration—driven by temperature gradients along channel walls—to induce gas flow from cold to hot regions in the free molecular regime (Knudsen number ≥10). These pumps, fabricated using nanoporous membranes like silica aerogels or bismuth telluride, enable microfluidics for vacuum generation and precise fluid delivery in compact devices, such as micro-gas chromatographs achieving flow rates up to 200 sccm without mechanical components. In the energy sector, free molecular flow principles underpin models of gas transport in extraction from tight formations, where sizes (often <10 nm) result in dominating over viscous flow due to frequent molecule-wall collisions. Knudsen slip effects, incorporating velocity discontinuities at pore walls, enhance apparent permeability in these low-porosity reservoirs, allowing for more accurate predictions of gas production rates and supporting enhanced recovery techniques like hydraulic fracturing by accounting for rarefied diffusion in matrix pores. This modeling approach integrates coefficients with slip factors to simulate multi-scale transport, improving estimates of recoverable reserves in unconventional resources. Free molecular flow is essential in for ensuring collision-free paths in sources and flight tubes, particularly under high-vacuum conditions where the exceeds system dimensions. In sample inlets, this regime allows effusive molecular beams to reach the region without intermolecular , enabling precise of gas and by maintaining laminar, non-turbulent entry of like . Sensors leveraging this flow, such as residual gas analyzers, benefit from minimized fragmentation and accurate sensitivity calibration in the collisionless environment. Emerging applications include thrusters for space propulsion, where plume expansion occurs in the free molecular regime due to Knudsen numbers greater than 1, resulting in negligible intermolecular collisions and beam-like trajectories. This flow enables efficient propellant utilization, with specific impulses exceeding 1,000 s and efficiencies up to 80% at powers above 5 kW, while ground testing in vacuum chambers requires modeling of rarefied background flows to mitigate facility effects on plume propagation and contamination. Transmission probability considerations briefly inform design for optimal flux in these collision-dominant wall-interaction environments.

References

  1. [1]
    [PDF] free molecule heat transfer - NASA Technical Reports Server (NTRS)
    This basic assumption of free molecular flow permits one to describe separately the fluxes of individual species in the gas in terms of maxwellian ...
  2. [2]
    None
    ### Summary of Free Molecular Flow from Livesey_mod.pdf
  3. [3]
    [PDF] An Expression for the Gas Mass Flow Rate through a Tube from ...
    – Rigorous integral equation for free-molecular flow for arbitrary ... • b0 from Kennard infinite-length free-molecular flow. • b1 = 0.15 as before ...
  4. [4]
    [PDF] 19940009482.pdf - NASA Technical Reports Server (NTRS)
    Free molecular flow is the regime of rarefied gas dynamics where collisions between molecules in the flow can be ignored when compared to the frequency of ...
  5. [5]
    How to calculate flow rate and types of flow in vacuum physics Leybold
    Molecular flow prevails in the high and ultrahigh vacuum ranges. In these regimes the molecules can move freely, without any mutual interference. Molecular flow ...
  6. [6]
    Vacuum Technology Fundamentals - Pfeiffer Know-How
    What prevails is molecular flow. In this case, the mean free path is significantly greater than the diameter of the flow channel. In molecular flow, the product ...Missing: science | Show results with:science<|control11|><|separator|>
  7. [7]
    Knudsen Number - an overview | ScienceDirect Topics
    The Knudsen number is defined as the ratio of the mean free path of a gaseous medium to the characteristic length scale of a model, and it is related to the ...
  8. [8]
    Vacuum Technology Fundamentals - Pfeiffer Know-How
    Knudsen flow in medium vacuum. If the Knudsen number is between 0.01 and 0.5, this is termed Knudsen flow. Since many process pressures are in the medium vacuum ...
  9. [9]
    Mean Free Path - an overview | ScienceDirect Topics
    ... 10−10sec. The mean free path increases rapidly with decreasing pressure. For instance, at an air pressure of 1 mm Hg l ≈ 10−2 cm; in a high vacuum of the ...
  10. [10]
  11. [11]
    Mean Free Path, Molecular Collisions - HyperPhysics Concepts
    Mean Free Path Calculation ; For pressure P0 = mmHg = inHg = kPa ; and temperature T= K = C = F, ; and times the average molecular separation of x 10^ m.
  12. [12]
    Mean Free Path - an overview | ScienceDirect Topics
    At a temperature of 20°C and under a pressure of 1.013 bar, the mean free path for air is 66.4 nm.
  13. [13]
    Die Gesetze der Molekularströmung und der inneren ...
    Die Gesetze der Molekularströmung und der inneren Reibungsströmung der Gase durch Röhren. Martin Knudsen, ... First published: 1909. https://doi.org/10.1002/andp.
  14. [14]
    VII. On stresses in rarified gases arising from inequalities ... - Journals
    ... 1879https://doi.org/10.1098/rstl.1879.0067. Abstract. 1. In this paper I have followed the method given in my paper “On the Dynamical Theory of Gases” (Phil.
  15. [15]
    [PDF] Measurements of Thermal Accomodation Coefficients - OSTI.GOV
    The free molecular heat flux is directly proportional to the gas pressure, P; in the limit of vanishing pressure, the heat flux approaches zero, as it must in a ...Missing: seminal | Show results with:seminal
  16. [16]
    [PDF] Experimental Measurements of Thermal Accommodation ...
    An experimental apparatus is described that measures gas-surface thermal accommodation coefficients from the pressure dependence of the conductive heat flux.Missing: sticking probability
  17. [17]
    An experimental assembly for precise measurement of thermal ...
    Mar 31, 2011 · An experimental apparatus has been developed to determine thermal accommodation coefficients for a variety of gas-surface combinations.<|control11|><|separator|>
  18. [18]
    [PDF] Fundamentals of Vacuum Technology - Leybold
    Dec 1, 2016 · ... Annalen der Physik, 41, 1913, 337-380. W. Becker. Eine neue Molekularpumpe. Vakuum-Technik, 7, 1958, 149-152. W. Armbruster.
  19. [19]
    [PDF] Chapter 31 Problems: Kinetic Molecular Theory
    The rate of collisions per unit area of an ideal gas with a wall is, Eq. ... The rate is given in molecules per unit volume per unit time. To convert to ...
  20. [20]
    Gas - Effusion, Kinetic Theory, Particles - Britannica
    Oct 15, 2025 · The relationship between vz and the average speed v̄ is rather straightforward: vz = (1/2)v̄. If the rates for two different gases effusing ...Missing: (1/4) \bar
  21. [21]
    On the transmission probability for molecular gas flow through a tube
    4D/3L. 1+ 4D/3L. (4). Equation (4) had been used for many years until Clausing4 published his theory on the transmission probability in 1928. In his theory ...
  22. [22]
    Efficient numerical solution of the Clausing problem - AIP Publishing
    Jun 7, 2007 · The Clausing integral equation for free-molecular gas flow ... A comparison of Clausing's results for the transmission probability (Clausing ...
  23. [23]
    [PDF] Thermodynamic Measurements Using the Knudsen Cell Technique
    The Knudsen cell technique uses a small enclosure to measure vapor pressures and thermodynamic properties by establishing a condensed phase/vapor equilibrium.Missing: v | Show results with:v
  24. [24]
    [PDF] Multicomponent gas diffusion and an appropriate momentum ...
    Multicomponent gas diffusion is reviewed with particular emphasis on gas flows near solid boundaries--the so-called. Kramers-Kistemaker effect.
  25. [25]
    A generalized Knudsen theory for gas transport with specular and ...
    Nov 15, 2023 · The free molecular flow is conventionally described by Knudsen theory, under the diffuse reflection assumption. Recent experiments reported ...
  26. [26]
    Knudsen flow 75 years on: the current state of the art ... - IOP Science
    Clausing (1926) developed, as an alternative to conductances, the concept of transmission probability, still referred to as the Clausing factor, and ...
  27. [27]
    [PDF] 1301 the flow of gases in narrow channels
    molecule region of flow - where A» r, Knudsen's formula is applicable . To = the gas volume measured by the product of volume and pressure flowing through ...
  28. [28]
    Die Molekularströmung der Gase durch Offnungen und die Effusion
    Annalen der Physik. Article. Full Access. Die Molekularströmung der Gase durch Offnungen und die Effusion. Martin Knudsen, ... First published: 1909. https://doi ...Missing: flow | Show results with:flow
  29. [29]
    [PDF] HISTORY OF VACUUM DEVICES
    gauge [38] measured the change in temperature of the filament with a thermocouple. In 1910. Knudsen invented the radiometer gauge [39], which with care could ...
  30. [30]
    Historical development and future trends of vacuum electronics
    Sep 7, 2012 · The molecular pump of Gaede was improved in the form of a multistage turbomolecular pump by Becker of the company Pfeiffer Vakuum, with ...
  31. [31]
    [PDF] Direct simulation Monte Carlo (DSMC): A numerical method for ...
    Mar 18, 2006 · A brief summary of the advances, which have occurred on and due to DSMC, is given. A few limitations and constraints of the method are also ...
  32. [32]
    Micro/Nano-pore Network Analysis of Gas Flow in Shale Matrix
    Aug 27, 2015 · The gas flow in the nano-scale pore may fall in flow regimes such as viscous flow, slip flow and Knudsen diffusion. A 3-dimensional nano-scale ...Missing: MEMS post-
  33. [33]
    A generalized Knudsen theory for gas transport with specular and ...
    Nov 15, 2023 · Here we develop a generalized Knudsen theory, which is applicable to various boundary conditions covering from the extreme specular reflection to the complete ...
  34. [34]
    NUMERICAL SOLUTION OF THE NONLINEAR BOLTZMANN ...
    The kinetic-theory treatment, which consists of solving the nonlinear Boltzmann equation, is necessary for a gas flow under non equilibrium conditions, since ...
  35. [35]
    [PDF] AERODYNAMIC TORQUES - NASA Technical Reports Server
    At orbital altitudes, the interaction of the vehicle and the very rarefied atmosphere can best be characterized by the free-molecular flow regime of gas ...
  36. [36]
    (PDF) A Free Molecule Micro-Resistojet: An Interesting Alternative to ...
    In this paper, an electric thruster that is scalable to spacecraft in the kilogram to hundreds of kilogram range is discussed.
  37. [37]
    Knudsen pumps: a review | Microsystems & Nanoengineering - Nature
    Jun 15, 2020 · The Knudsen pump (KP) is a kind of micro-pump that can form thermally induced flows induced by temperature fields in rarefied gas environments.
  38. [38]
    Flow and Transport in Tight and Shale Formations: A Review - 2017
    Sep 18, 2017 · Knudsen Diffusion in Shale. When the mean free path of gas molecules is on the same order as the tube dimensions (as in shale), Knudsen ...
  39. [39]
    Free Molecular Flow in the Sample Inlet to the Mass Spectrometer
    Aug 8, 2025 · Effusion of hydrogen from sample reservoir. Peak heights indicate rate at which hydrogen passed through the mass spectrometer ion source; ...
  40. [40]
    [PDF] ADA506238.pdf - S3VI - Small Spacecraft Systems Virtual Institute
    Aug 18, 2009 · Due to the free molecular flow regime in which these thrusters operate, it is not uncommon for materials to redeposit within the Hall effect ...
  41. [41]
    Free molecular background flow in a vacuum chamber equipped ...
    Dec 9, 2005 · This study suggests three simple free molecular flow models by considering the fact that the background flow inside a vacuum chamber is highly ...Missing: definition | Show results with:definition