Fact-checked by Grok 2 weeks ago

Permeation

Permeation is the process of molecular of gases, vapours, or fluids through the material of a solid, occurring at a molecular level without visible of the material. This phenomenon is driven by a concentration gradient across the material, where molecules sorb onto one side, diffuse through the matrix, and desorb on the other side. The rate of permeation, often quantified as , depends on several key factors, including the material's intrinsic permeability, thickness, and , as well as the properties of the permeating substance. For instance, increasing the thickness inversely reduces the permeation rate, while higher temperatures generally accelerate it by enhancing molecular mobility. Permeation is distinct from , which is the broader random movement of molecules, and from , which involves surface-level entry rather than full traversal of the material. Permeation plays a critical role in various scientific and applications, such as technologies for , gas separation, and systems, where controlled permeability enables efficient separation processes. In materials , it is essential for designing protective barriers, like chemical-resistant gloves or hydrogen permeation coatings on metals, to prevent unintended leakage or embrittlement. Additionally, in packaging and microfluidic devices, managing permeation ensures product integrity and enables precise fluid handling.

Fundamentals

Definition and Principles

Permeation is the process by which a permeant, such as a gas, vapor, , or solute, penetrates and traverses a permeable barrier —typically a solid or semi-solid—through molecular-level mechanisms involving and . In non-porous materials, this occurs without macroscopic pores or visible defects. This movement occurs as the permeant transitions from a region of higher concentration to lower concentration across the barrier, driven by gradients. While pressure differences can influence (e.g., via ), permeation itself is distinct from pressure-driven bulk flow or convection. The fundamental principles of permeation in non-porous membranes are encapsulated in the solution-diffusion model, which posits that transport involves two sequential steps: the permeant first dissolves (or sorbs) into the upstream surface of the material according to its S, and then diffuses across the material matrix under a , governed by the D. The overall permeability P, defined as the steady-state flux of permeant per unit driving force (e.g., or concentration difference) normalized by the material thickness, is the product of these parameters: P = D \times S This model distinguishes permeation as a diffusive process at the molecular scale. In contrast, porous materials may involve additional convective flow through pores, described by Darcy's law. Several key factors influence the rate of permeation. The primary driving force is the concentration gradient across the material, which dictates the diffusive flux as per foundational transport principles. Temperature affects both solubility and diffusivity, typically increasing the permeation rate exponentially due to enhanced molecular mobility, with a rule-of-thumb that rates double for every 10°C rise in many polymer systems. Material properties, such as polymer density, crystallinity, and the presence of porosity, modulate D and S; for instance, higher crystallinity reduces free volume and thus lowers diffusivity, while porosity can enable alternative transport modes.

Mechanisms of Transport

Permeation occurs through a sequence of molecular-level steps that enable a permeant to cross a barrier material, such as a . The process begins with (or ), in which the permeant integrates into the bulk material, often governed by parameters that determine the equilibrium concentration within the matrix. Once dissolved, the permeant undergoes , a random molecular motion driven by concentration gradients that propels it across the barrier to the opposite side. Finally, desorption releases the permeant from the exit surface, allowing it to enter the receiving phase and complete the . In non-biological materials like , permeation typically involves simple , where small, non-polar molecules such as gases move passively through . The microstructure of the barrier significantly affects permeation pathways and rates by providing or restricting routes for molecular movement. Defects and voids, such as micro-pores formed during composite fabrication or in polymer blends, act as low-resistance channels that accelerate , often increasing permeability by orders of magnitude compared to defect-free materials. In , chain mobility enhances transport by allowing temporary openings in for permeant passage; increased mobility, induced by temperature or solvents, expands free volume and boosts coefficients, though excessive swelling can sometimes hinder net . These structural features collectively determine the permeability coefficient, which encapsulates the combined effects of and in quantifying overall transport.

Theoretical Models

Fick's Laws Application

Fick's first law describes the diffusive J of a permeating species as proportional to the negative gradient of its concentration c, expressed in one dimension as J = -D \frac{\partial c}{\partial x}, where D is the diffusion coefficient. This law, originally formulated by Adolf Fick in 1855 based on analogies to heat conduction, provides the basis for modeling steady-state permeation through by assuming a constant under equilibrium conditions. In the context of permeation, the law is adapted to represent the steady-state of solutes or gases across a membrane, where the concentration gradient drives transport from high to low concentration regions without net accumulation. For steady-state conditions in a of thickness L, integration of Fick's first law across the membrane, assuming a linear concentration profile (valid when D is constant), yields the permeation rate J = \frac{P \Delta c}{L}, where \Delta c = c_0 - c_L is the concentration difference across the boundaries at x = 0 and x = L, and P is the permeability, related to D through the or of the permeant in the membrane. This links diffusive directly to measurable permeation parameters, enabling predictions of flux in isotropic materials under non-reactive conditions. Fick's second law extends to time-dependent by incorporating mass conservation, resulting in the \frac{\partial c}{\partial t} = D \frac{\partial^2 c}{\partial x^2} for one-dimensional, non-steady-state in homogeneous media. This equation governs transient permeation profiles, such as the initial buildup of concentration in a exposed to a permeant, and its analytical solutions often involve the for cases like semi-infinite slabs or sudden exposure to constant surface concentrations. For example, in a semi-infinite with fixed surface concentration c_0 at x = 0 and initial uniform concentration c_i, the concentration profile is given by c(x, t) = c_0 + (c_i - c_0) \erf\left( \frac{x}{2\sqrt{D t}} \right), where \erf is the Gaussian error function, providing a mathematical description of how the permeation front advances over time. The application of Fick's laws to permeation relies on key assumptions, including material isotropy, constant D independent of concentration or position, and absence of chemical reactions or convection that could alter the gradient. These laws are most suitable for thin membranes where steady-state conditions are rapidly achieved (e.g., via quick integration across small L), but for thicker membranes, transient solutions from the second law are essential to capture lag times and non-uniform profiles before equilibrium flux is established. Limitations arise when D varies with temperature, permeant concentration, or external factors, requiring modified models beyond classical Fickian diffusion.

Permeability Coefficients

The permeability , denoted as P, quantifies a material's intrinsic to permit the permeation of a specific substance, such as a gas or , through its under a given driving force. It is defined as the product of the D (measuring molecular mobility) and the S (measuring uptake capacity), expressed as P = D \times S. This is independent of sample thickness, distinguishing it from , which represents the thickness-normalized (permeability divided by thickness) and describes the overall rate of permeation through a specific . Common units for gas permeability in polymers include the , where 1 = $10^{-10} cm³ (STP) · cm / (cm² · s · cmHg), while an used in packaging applications is cm³ · mm / (m² · day · atm). Several factors influence the . Temperature dependence typically follows an Arrhenius , P = P_0 \exp(-E_p / RT), where P_0 is a , E_p is the for permeation, R is the , and T is the absolute ; this reflects the thermally activated nature of and solubility processes, with permeability generally increasing exponentially with . Pressure effects vary by material: in rubbery polymers, permeability is often largely independent of at low levels, but in glassy polymers, it may decrease due to saturation of sorption sites. The size and shape of the permeant also play key roles, as larger or more rigid molecules experience greater steric hindrance, reducing and thus overall permeability. Permeability coefficients are classified by permeant type, with gas permeability being the most studied for applications like gas separation membranes; representative examples include oxygen (O₂) and carbon dioxide (CO₂), where values reflect material selectivity. Liquid permeability follows similar principles but is less common due to higher viscosities and interactions. In glassy polymers, the dual-mode sorption model accounts for non-linear behavior, combining Henry's law dissolution in equilibrium regions with Langmuir-type adsorption in non-equilibrium microvoids, leading to pressure-dependent permeability that often decreases initially before stabilizing. This model, developed for penetrants like CO₂, highlights how glassy structures enable higher initial solubility but constrain diffusion compared to rubbery counterparts. Relative permeability scales vary widely across materials, with elastomers exhibiting orders-of-magnitude higher values than metals, underscoring their use in permeable applications versus barriers. For instance, at 25°C, shows O₂ permeability around 15-20 , while is lower at about 0.15 , reflecting tighter chain packing. In contrast, metals like and aluminum display extremely low gas permeabilities (e.g., <10^{-10} ), making them nearly impermeable at ambient conditions due to dense crystalline lattices.
Material TypeExample MaterialO₂ Permeability (approx. barrer at 25°C)Relative Scale
Rubbery Polymer15-20High (permeable)
Rubbery Polymer0.1-0.2Moderate
Metal<10^{-10}Extremely Low (barrier)
MetalAluminum<10^{-15}Extremely Low (barrier)

Historical Development

Early Observations ()

In 1748, French physicist and clergyman Abbé Jean-Antoine Nollet (1700–1770) conducted pioneering experiments that marked the first documented scientific observations of permeation through semipermeable , using pig bladders as natural barriers. He sealed a container of with a pig bladder and submerged it in , observing that the bladder swelled as water permeated through the membrane into the alcohol, increasing its volume, while the alcohol did not escape. This demonstrated the membrane's selective permeability, allowing solvent passage but retaining the solute, distinguishing the process from mere leakage through imperfections. Nollet observed that the intense swelling of the indicated the development of significant due to the influx, providing an early empirical between permeation and hydrodynamic forces. These qualitative demonstrations highlighted a conceptual shift in understanding material transport: permeation was recognized as a directed, pressure-influenced movement across intact barriers, rather than random or . Although Nollet did not quantify rates or develop mathematical models, his work emphasized the role of selectivity in separating substances, laying foundational insights for later studies in . The era's limitations were evident in the reliance on visual and observations without precise tools, such as manometers or chronometers for rates, restricting analyses to descriptive accounts. Early terminology drew from hydrodynamics, with Nollet describing the in terms akin to "" or solvent "pushing" through pores, foreshadowing the later adoption of "" (from for ) to denote similar permeation-driven flows in the . These 18th-century efforts thus established permeation as a distinct physical observable in biological and simple artificial systems.

Key Advances (19th-20th Centuries)

In the mid-19th century, Thomas Graham advanced the understanding of permeation through his pioneering work on , published in 1861, where he demonstrated the separation of colloidal substances from crystalloids using semi-permeable membranes such as . Graham observed that crystalloids, like salts, readily through these membranes into surrounding water, while colloids remained behind, establishing a direct link between processes and selective permeation across barriers. This experimental approach not only introduced the term "" but also highlighted how structure influences , laying groundwork for later studies in both chemical analysis and biological systems. Concurrently, Adolf Fick formalized the mathematical principles of in 1855 through his publication "On Liquid Diffusion," deriving laws that describe the of solutes as proportional to the , analogous to conduction. As a physiologist, Fick applied these laws to biological contexts, modeling the permeation of substances like salts and water through tissues and membranes, including early considerations of transport across barriers. Fick's , in particular, provided a quantitative framework for predicting steady-state permeation rates in living systems, influencing subsequent research on nutrient and in physiological processes. Entering the , Richard Barrer significantly expanded the theoretical and experimental study of gas permeation in during the to 1960s, with his seminal 1939 paper detailing the permeation, diffusion, and solution of gases in organic polymers. Barrer's work formalized the concept of permeability as a measurable property in these materials, establishing the solution-diffusion model wherein gases first dissolve into the before diffusing across it under a concentration gradient. This model, which integrates solubility and diffusivity coefficients to predict overall permeation, became a cornerstone for understanding non-porous and inspired the naming of the permeability unit "" in his honor. Mid-20th-century advancements were propelled by demands for protective equipment, particularly research on rubber materials to minimize gas permeation rates in gas masks and suits. , developed in the late , emerged as a key material due to its exceptionally low gas permeability—up to 10 times lower than for gases like oxygen and —enabling effective barriers against chemical agents. This wartime focus accelerated quantitative studies of diffusion coefficients in elastomers, bridging theoretical models with practical applications in .

Measurement and Analysis

Experimental Techniques

Experimental techniques for measuring permeation rates and coefficients typically employ controlled setups to expose material samples to a permeant under defined conditions, monitoring the through the material over time. These methods are essential for characterizing transport properties in applications ranging from to gas separation membranes. Common approaches include pressure-based, time-dependent, and mass-based measurements, often standardized for reproducibility. The decay method involves placing a sample in a where a known volume of gas is introduced on the upstream side at elevated , while the downstream side remains at or low . The rate of upstream decrease is monitored using a , which directly correlates to the permeation flux through the material. This technique is particularly suitable for high-permeability materials, such as thin films, as it allows rapid assessment without needing sensitive downstream detectors. The time-lag method determines both permeability and diffusion coefficients from a single permeation experiment by analyzing the transient buildup of permeant on the downstream side. In this setup, the upstream side of the is exposed to a constant gas , and the downstream or is recorded until steady-state conditions are reached. The lag time \tau, defined as the intercept of the steady-state linear versus time plot with the time axis, relates to the diffusion coefficient D via the equation \tau = L^2 / (6D), where L is the thickness; permeability is then obtained from the steady-state slope. This method, originally developed by Daynes, is widely used for dense polymeric membranes due to its simplicity and ability to decouple diffusion from effects. Isostatic and gravimetric techniques are employed for measuring and vapor permeation, particularly in scenarios involving condensable permeants like or organic solvents. In the isostatic method, the sample separates two chambers maintained at equal but with different permeant concentrations, such as liquid on one side and vapor or dry gas on the other, allowing permeation driven by difference to be quantified by downstream concentration buildup. Gravimetric approaches, often aligned with standards like ASTM E96 for , involve weighing the mass gain or loss in a sealed over time, where the sample acts as a barrier to vapor from a controlled source. The ASTM D1434 standard outlines manometric procedures for gas rates in plastics. These methods are ideal for low-permeability materials and provide permeability coefficients that account for in the context of Fickian transport. Advanced tools enhance precision for complex systems, such as multi-gas mixtures or nanoscale films. Quadrupole mass spectrometry (QMS) is integrated into permeation cells to analyze downstream gas composition in real-time, enabling separation of individual fluxes in mixed-gas experiments and detection of trace permeants as low as $10^{11} molecules/s. For thin films, nano-scale setups utilizing (AFM) image surface microstructure and identify defects that influence local permeation, providing insights into heterogeneity that bulk methods overlook. Recent optical methods, such as infrared laser spectroscopy and fluorescence quenching (as of 2023), enable simultaneous measurement of and oxygen permeability in barrier films. These techniques are crucial for high-resolution studies in materials like metal-organic frameworks or ultrathin coatings.

Data Interpretation and Approximations

In the steady-state approximation for permeation through thin membranes, the flux J_{ss} is derived from Fick's first law as J_{ss} = P \cdot \frac{\Delta p}{L}, where P is the permeability coefficient, \Delta p is the difference across the , and L is the membrane thickness; this assumes a linear concentration gradient and negligible transient effects for membranes where L is much smaller than the lateral dimensions. For non-ideal gases at high pressures, corrections replace the pressure driving force \Delta p with the difference \Delta f to account for deviations from ideal behavior, improving accuracy in systems like CO₂ permeation through polymers. Transient analysis involves fitting pressure-time data from permeation experiments to extract the diffusion coefficient D and solubility coefficient S, where permeability P = D \cdot S; the time-lag \theta at steady-state onset is given by \theta = \frac{L^2}{6D}, allowing D to be determined from the intercept of the asymptotic linear portion of the cumulative permeated amount versus time. Edge effects, arising from lateral diffusion at membrane boundaries, are minimized by ensuring the thickness-to-radius ratio L/r < 0.2 or by sealing edges with impermeable materials, while system leaks are handled by subtracting the baseline permeation rate measured under vacuum conditions prior to analysis. Common error sources in permeation data include temperature fluctuations, which exponentially affect D and S via Arrhenius dependence, leading to variability in measured P; membrane defects such as pinholes or microcracks that artificially inflate ; and permeant adsorption on surfaces, which delays and biases \theta. Statistical methods for quantifying uncertainty, such as calculating 95% confidence intervals from replicate measurements using propagation, help assess the reliability of derived parameters like D and P, particularly in noisy datasets from low-permeability materials. Modern computational aids, including finite element modeling (FEM), simulate permeation in complex geometries like non-planar or heterogeneous membranes by solving Fick's laws numerically, enabling prediction of non-uniform fluxes and concentration profiles. These models validate Fickian assumptions by comparing simulated time-lag and steady-state fluxes against experimental data, identifying deviations due to or dual-mode in polymers.

Applications

Industrial and Materials Science

In industrial applications, permeation plays a critical role in designing gas barrier materials for , where materials like (EVOH) are employed to control oxygen ingress and extend product . EVOH exhibits exceptionally low oxygen rates (OTR), often below 1 cm³·m⁻²·day⁻¹·atm⁻¹ at standard conditions, which minimizes oxidation in oxygen-sensitive foods such as meats and , thereby preserving nutritional value and flavor. These permeation rates are used in predictive models to estimate , where lower OTR values correlate with longer storage durations under ambient conditions. Membrane technology leverages controlled permeation for processes like (RO) in , where thin-film composite membranes facilitate high water flux while rejecting salts. Typical RO membranes for seawater achieve water permeabilities of 1–2 liters per square meter per hour per bar (LMH/bar), enabling efficient production of potable water from saline sources with salt rejection rates exceeding 99%. In fuel cells, proton exchange membranes () such as allow selective proton permeation while minimizing gas crossover; this low hydrogen crossover enhances cell efficiency and safety. Materials engineering advances permeation control through nanocomposites, where incorporating nanofillers like clay or into polymer matrices creates tortuous paths that reduce gas permeability by up to 90% compared to neat . For carbon capture, specialized membranes achieve CO₂/N₂ selectivities exceeding 100, allowing efficient separation from flue gases; mixed-matrix designs with metal-organic frameworks demonstrate selectivities up to 300 while maintaining CO₂ permeance above 1000 gas permeation units (GPU). These engineered barriers are tested using like ISO 15105, which employs differential-pressure methods to quantify gas transmission rates in and sheets. The economic significance of permeation-based technologies is evident in the global membrane filtration market, valued at approximately $21 billion in 2025, driven by demand in , , and sectors. This growth underscores the role of permeation optimization in reducing operational costs, such as energy savings in plants through higher-flux membranes.

Biological and Environmental Uses

In biological systems, permeation plays a fundamental role in passive diffusion across , allowing nonpolar molecules like oxygen to traverse lipid bilayers without energy input. This process relies on the of the molecule in the hydrophobic core of the bilayer, enabling oxygen to dissolve and move down its concentration gradient to support in tissues such as the lungs and muscles. in the can modulate this permeation by reducing oxygen flux, which helps maintain membrane integrity while controlling rates. Permeation is equally critical in , where drugs are absorbed through biological barriers like the skin and . For delivery, lipophilic drugs permeate the via passive , influenced by molecular weight, , and enhancers that temporarily disrupt the barrier. In the gut, passive permeation across the accounts for the absorption of many orally administered drugs, driven by the paracellular or transcellular routes depending on the compound's size and charge, with the fraction absorbed () estimated mechanistically from permeability coefficients. This process is a key determinant of , as poorly permeable drugs may require strategies to enhance epithelial crossing. Selective permeation is facilitated by specialized proteins that enable precise transport while excluding unwanted solutes. Aquaporins, integral membrane channels, provide highly selective pores for water molecules, regulating transcellular flow in renal collecting ducts to concentrate urine and in plant roots to facilitate water uptake under osmotic stress. In neurons, ion channels ensure selective permeation of ions like sodium and , generating action potentials essential for ; these channels fluctuate between open and closed states, maintaining ion gradients across the . Such selectivity underpins neural excitability and prevents uncontrolled ion flux that could disrupt cellular . In environmental contexts, permeation underpins remediation techniques for contaminated sites. Soil vapor extraction (SVE) leverages the volatilization and diffusive permeation of organic contaminants from soil pores into the vapor phase, where a induces airflow to capture and remove them from the unsaturated zone, effectively treating volatile organic compounds like . Permeable reactive barriers (PRBs) function by allowing plumes to permeate through a subsurface wall of reactive media, such as zero-valent iron, where contaminants undergo degradation or immobilization via chemical reactions as the water flows passively. These passive systems have been deployed at numerous sites to intercept and treat chlorinated solvents without ongoing energy inputs. Despite these benefits, challenges in biological and environmental permeation applications persist. Biofouling, the accumulation of microorganisms and extracellular polymeric substances on membranes, significantly reduces permeability in and biomedical devices by increasing hydraulic resistance and clogging pores, often requiring pretreatment or coatings to mitigate performance decline. In landfills, permeation through cover soils via contributes to , accounting for a substantial portion of releases that exacerbate , with global estimates indicating approximately 65 of (CH₄), equivalent to about 1,820 CO₂-equivalents annually (using a 100-year of 28 for ). Engineered covers aim to limit this , but variability in and temperature can enhance escape.

Special Cases

Gas Solubility in Metals

Gas solubility in metals, particularly for diatomic gases like , follows Sieverts' , which describes the concentration of dissolved gas atoms as proportional to the of the gas : S = K \sqrt{p}, where S is the (atomic concentration), K is the temperature-dependent constant, and p is the . This arises from the of diatomic molecules at the metal surface, limiting by the between molecular gas and atomic species in the . exhibits exceptionally high , with up to 900 times its volume at and , enabling applications in and purification due to favorable . In steady-state permeation through a metal of thickness L, the J of diatomic gas is given by J = \frac{D K \sqrt{p}}{L}, where D is the coefficient of the atomic species; this combines with Fickian under a concentration driven by the difference. Permeation rates are thermally activated, with typical activation energies for in steels ranging from 20 to 50 kJ/mol, reflecting barriers to both and that increase with alloying elements like . These processes pose significant challenges in industrial settings, such as permeation in nuclear reactors, where isotopic diffuses through structural metals like , risking contamination of systems and environmental release unless mitigated by barriers. In oil and gas pipelines, (H₂S) environments promote atomic ingress, leading to sulfide stress cracking and reduced integrity, with permeation rates correlating directly to cracking susceptibility. Experimental measurements of gas solubility in metals often require high-temperature setups exceeding 300°C to achieve measurable dissolution rates, particularly for low-solubility systems like in ferritic steels, using controlled atmospheres to maintain partial pressures. Isotopic tracers, such as or , enable precise quantification by distinguishing permeated species via , improving accuracy in and determinations over protium-based methods.

Permeation in Polymers and Membranes

Permeation in polymers and membranes is governed by specific models that account for the amorphous nature and chain dynamics of these materials. The free-volume theory, originally developed by and Turnbull, posits that of penetrants in s arises from the availability of transient voids or free volume within the polymer matrix, allowing molecules to jump between sites when the void exceeds a critical size. This theory has been adapted for polymers, where D is expressed as D = A \exp\left(-\frac{B}{v_f}\right), with v_f representing the average free volume per molecule, A a constant related to jump frequency, and B tied to the minimum void size needed for . In glassy polymers, the limited free volume restricts penetrant mobility, leading to lower permeability compared to rubbery states. The temperature (T_g) markedly influences permeability (P), as polymers transition from a flexible, rubbery state above T_g to a rigid, glassy state below it. Below T_g, chain segments are frozen, reducing free volume and causing a sharp decrease in P, often by orders of magnitude for gases like O_2 and CO_2. For instance, in polyimides, permeability drops significantly in the glassy regime due to this constrained dynamics, emphasizing the role of thermal history in design. Synthetic membranes for permeation, particularly asymmetric types, exploit properties to achieve efficient separation. Asymmetric membranes feature a thin, dense selective skin layer supported by a porous substructure, minimizing resistance while maximizing ; is a classic example used in gas separation, such as CO_2/CH_4 mixtures, due to its moderate permeability and selectivity. A key challenge is the selectivity-permeability trade-off, encapsulated by the Robeson upper bound, which plots maximum achievable selectivity against permeability for gas pairs like O_2/N_2, showing that higher permeability typically correlates with lower selectivity owing to solution-diffusion mechanisms. Several factors modulate permeation in these systems. Plasticization occurs when permeants like CO_2 interact with the polymer, inducing swelling that increases free volume and , thereby elevating P but potentially compromising selectivity; for example, CO_2 swells poly(vinyl chloride) (PVC), leading to enhanced transient diffusion coefficients. Physical aging, a relaxation process in glassy polymers, further reduces long-term P by densifying the matrix and decreasing free volume over time, often resulting in 50-90% permeability loss after months, as observed in polymers of intrinsic microporosity (PIMs). Modern advances address these limitations through mixed-matrix membranes (MMMs), incorporating metal-organic frameworks (MOFs) into polymer matrices to enhance selectivity without sacrificing permeability. MOFs provide rigid pores that favor specific gases, disrupting the Robeson trade-off; for instance, incorporation of ZIF-7 nanocrystals into rigid matrices has achieved O₂/N₂ selectivities of around 15, compared to typical values of 5-10 for pure polyimides. These designs leverage the high surface area and tunable apertures of MOFs to enable precise molecular sieving.

References

  1. [1]
    What is Permeation? (A Definitive Guide) - TWI Global
    Permeation is the process of molecular penetration of gases, vapours or fluids through the material membrane of a solid.
  2. [2]
    Permeation FAQ - AMETEK MOCON
    Permeation is the movement of gas and vapor through a barrier such as the wall of a bottle. It is driven by the permeant concentration gradient.Missing: scientific | Show results with:scientific
  3. [3]
    Key factors affecting permeation - SIESA
    Aug 1, 2019 · The liner wall thickness is inversely related to the permeation rate, ie doubling the wall thickness halves the permeation rate. On the wet side ...
  4. [4]
    Maximizing the right stuff: The trade-off between membrane ...
    Jun 16, 2017 · Synthetic membranes are used for desalination, dialysis, sterile filtration, food processing, dehydration of air and other industrial, medical, ...
  5. [5]
    Mechanism and Evaluation of Hydrogen Permeation Barriers
    Adding hydrogen permeation barrier coatings to the surface of stainless steel is one of the most effective approaches to delay or prevent the permeation of ...
  6. [6]
    Permeation-driven flow in poly(dimethylsiloxane) microfluidic devices
    Permeation-driven flow offers a way to concentrate chemicals at specific regions in a channel. For example, a symmetric “T” design could concentrate species at ...
  7. [7]
    Permeation Property - an overview | ScienceDirect Topics
    The passage of small molecules through solid materials is called permeation. Permeability properties are important in many applications and are important in ...
  8. [8]
    Permeation - Corrosionpedia
    Permeation is the rate of molecular diffusion of gases, vapor and fluids through a solid material membrane. It is inversely proportional to the subsurface ...<|control11|><|separator|>
  9. [9]
    Transport properties of cellulose acetate osmotic membranes
    Diffusion and distribution coefficients of water and sodium chloride have been measured in cellulose acetate osmotic membranes.
  10. [10]
    [PDF] Permeation and Measurement Techniques - TAPPI.org
    Chemical Potential. Chemical potential, μ, is the actual driving force for permeation. Chemical potential is a thermodynamic property related to pressure. μ.<|control11|><|separator|>
  11. [11]
    The Permeability of Rubber Compounds | Hot Topics
    Dec 18, 2018 · Temperature is another factor that complicates permeability. Typically, the higher the temperature, the higher the rate of diffusion. Thus, it ...
  12. [12]
    [PDF] FACTORS AFFECTING THE PERMEABILITY OF PE BLOWN FILMS
    These two permeation rates are directly related, and the major factors affecting permeability are shown to be density, film gauge and crystalline orientation.
  13. [13]
    [PDF] PHYSICS OF OUTGASSING
    Permeation of gases is a combination of two physical processes: dissolution and diffusion. First, the gas dissolves in the solid and then it diffuses to the ...
  14. [14]
    Transport of Small Molecules - The Cell - NCBI Bookshelf - NIH
    Two classes of proteins that mediate facilitated diffusion are generally distinguished: carrier proteins and channel proteins. Carrier proteins bind specific ...
  15. [15]
    Molecular Dynamics Simulations of Membrane Permeability - PMC
    This Review illustrates the evaluation of permeability of lipid membranes from molecular dynamics (MD) simulation primarily using water and oxygen as examples.
  16. [16]
    An overview of the transport of liquid molecules through structured ...
    The transport of permeants in polymer films, which is determined by several factors such as the solubility of the permeant and the diffusion coefficient for the ...
  17. [17]
    Permeability - EdTech Books
    Permeability is how easily gases or liquids pass through a material. Low permeability means difficulty, often requiring long time or high pressure.<|control11|><|separator|>
  18. [18]
    24. Diffusion | Introduction to Solid State Chemistry
    Fick, Adolf. “Ueber Diffusion.” Annalen der Physik 170 (1855): 59-86. (Note: this article is in German.) Carslaw, Horatio ...Missing: Über | Show results with:Über
  19. [19]
    [PDF] Lecture 32 Membrane Permeability
    Fick's first law! Page 4. Diffusion Across Vesicle Bilayer. Fick's first law: J = −D. dC dx. Concentration gradient: dC dx. ≈. Cin − Cout. Bilayer thickness.
  20. [20]
    Derivation of Fick's first law - Diffusion - DoITPoMS
    Fick's first law relates this concentration gradient to the flux, J, of atoms within the crystal (that is, the number of atoms passing through unit area in ...<|separator|>
  21. [21]
    Fick's second law - DoITPoMS
    Fick's second law is concerned with concentration gradient changes with time. Diagram of concentration gradient
  22. [22]
    [PDF] THE MATHEMATICS OF DIFFUSION - eng . lbl . gov
    The book contains a collection of mathematical solutions of the differential equations of diffusion and methods of obtaining them. They are discussed against a ...<|control11|><|separator|>
  23. [23]
    [PDF] PERMEABILITY DATA AEROSPACE APPLICATIONS
    Factors affecting per- meation, such as temperature ... At least twenty-eight unit systems are currently in use for expression of permeation rates.
  24. [24]
    Membrane Basics - Liquid Contacting and Gas Separation
    The permeability coefficient is a parameter defined as the transport flux of a gas (rate of gas permeation per unit area), per unit transmembrane driving force ...
  25. [25]
    Temperature and Pressure Dependence of Gas Permeation in ... - NIH
    In order to explore the temperature dependence of the gas permeability, the data were correlated with the Arrhenius equation. P = P 0 e x p ( − E p R T ) ...
  26. [26]
    Model for sorption of mixed gases in glassy polymers - Koros - 1980
    A model is presented for analysis of the sorption of mixed gases in glassy polymers at concentrations below which significant plasticization occurs.
  27. [27]
    Dual-Mode Sorption Model - an overview | ScienceDirect Topics
    The thermodynamic model describes penetrant permeability in glassy polymers and the model represents experimental behaviors with and without plasticization ...
  28. [28]
  29. [29]
  30. [30]
  31. [31]
    The prehistory of haemodialysis as a treatment for uraemia - GIN
    Feb 18, 2016 · Thomas Graham: diffusion dialysis and semipermeable membranes. Graham also quoted Dutrochet's prior work. Thomas Graham (1805-1869) [9] ...
  32. [32]
    Fick's Insights on Liquid Diffusion
    ### Summary of Adolf Fick's 1855 Work on Diffusion and Its Application to Biology
  33. [33]
    Permeation, diffusion and solution of gases in organic polymers
    Permeation, diffusion and solution of gases in organic polymers. R. M. Barrer and E. K. Rideal, Trans. Faraday Soc., 1939, 35, 628 DOI: 10.1039/TF9393500628.
  34. [34]
    A sixty-year love affair with membranes Recollections of Richard M ...
    Permeation, solution and diffusion of gases in organic polymers. Trans. Faraday Soc., 35 (1939), p. 628. [4]. Activated diffusion in membranes. Trans. Faraday ...
  35. [35]
    All About Butyl Rubber – Properties, Applications, and Uses
    Oct 10, 2025 · Excellent Air Impermeability. Butyl rubber's molecular structure provides exceptionally low gas permeability, making it one of the most ...
  36. [36]
    Direct determination of permeability, diffusivity and solubility of ...
    This technique introduces a known amount of gas into an evacuated chamber containing the membrane sample, and the pressure decay in the chamber corresponds to ...
  37. [37]
    A novel time lag method for the analysis of mixed gas diffusion in ...
    Sep 1, 2018 · A novel method to determine the individual diffusion coefficients of gases in a mixture during their permeation through polymeric membranes is described.
  38. [38]
    D1434 Standard Test Method for Determining Gas Permeability ...
    Jun 21, 2023 · ASTM D1434 is a test method using a manometric method to determine gas transmission rate, permeance, and permeability of plastic film and ...
  39. [39]
    Isostatic and Quasi-lsostatic Methods for Determining tiie ...
    ABSTRACT: Two test methods have been developed which provide quantitative and reli- able values for the rate of diffusion of organic vapors through polymer ...
  40. [40]
    Measurement of gas-concentration-driven permeation for the ...
    Oct 7, 2020 · ϕ = D S . Permeability is composed of two thermally activated Arrhenius relationships (D and S); thus, it can be expressed by the following ...
  41. [41]
    [PDF] Water Vapor Permeation in Plastics
    Studies at PNNL have found that only certain locations have the weather conditions necessary to cause the observed fogging, and the effect has been found in ...
  42. [42]
    Experimental Study of CO2 Plasticization in Polysulfone Membrane ...
    Since the current study addresses high pressure conditions, non-ideal gas conditions should be considered. The driving force for this case is described as the ...<|control11|><|separator|>
  43. [43]
    The Time Lag in Diffusion | The Journal of Physical Chemistry
    A Novel Time Lag Method to Measure the Permeation of Vapor-Gas Mixtures. Journal of Membrane and Separation Technology 2012, 1 (2) , 94-99. https://doi.org ...
  44. [44]
    Error analysis and optimum design of permeation time-lag ...
    The systematic and random experimental errors associated with the usual method of measuring the permeation time lag of a diffusion system obeying Fick's and ...
  45. [45]
    Sorption, desorption and diffusion processes in membrane permeation
    It is widely accepted that gas permeation is an activated process of diffusion within the membrane and a partitioning process of the gas at the interfaces with ...
  46. [46]
    Finite Element Analysis of Gas Diffusion in Polymer Nanocomposite ...
    The aim of this work is to analyze, through finite element modeling (FEM), the diffusion behavior of 3D systems representative of PLA–CNC nanocomposites.
  47. [47]
    Gas Transport in Mixed Matrix Membranes: Two Methods for Time ...
    The most widely used method to measure the transport properties of dense polymeric membranes is the time lag method in a constant volume/pressure increase ...
  48. [48]
    [PDF] Value Proposition of Ethylene Vinyl Alcohol (EVOH) for Food and ...
    That is, the higher the surface area to volume ratio of the package, the lower the oxygen transmission rate needed to reach the same shelf life for a given ...
  49. [49]
    High oxygen barrier packaging materials from protein-rich single ...
    Oct 6, 2025 · EVOH is widely used in food packaging because it extends the shelf life of oxygen-sensitive foods by minimizing oxygen permeance and thereby ...
  50. [50]
    How RO membrane permeability and other performance factors ...
    Nov 15, 2019 · In recent RO membranes, the water permeability (A value) is about 1–2 LMH/bar for seawater RO (SWRO), and 2–8 LMH/bar for brackish water RO ( ...Missing: desalination | Show results with:desalination
  51. [51]
    Numerical and experimental characterization of gas permeation ...
    Feb 1, 2023 · This paper first establishes a numerical technique to predict the gas permeation through membrane with consideration of mechanical degradation.
  52. [52]
    Use of nanocomposites as permeability reducing admixtures
    Mar 23, 2018 · Macroscopic measurements show that the addition of a small amount of nanoparticles effectively reduces the extent of the water permeation front.
  53. [53]
    [PDF] Two-stage Membrane-based Process Utilizing Highly CO2-selective
    permeable membrane, in order to achieve energy efficient carbon capture of 90% CO2 capture ... moderate CO2 permeance (1,000 GPU) and CO2/N2 selectivity >300.
  54. [54]
    ISO 15105-1:2007 Plastics — Film and sheeting — Determination of ...
    2–5 day deliveryOne method uses a pressure sensor, the other a gas chromatograph, to measure the amount of gas which permeates through a test specimen. General information.Missing: permeation | Show results with:permeation
  55. [55]
    Membrane Filtration Market Size, Share | Growth Report [2032]
    Oct 29, 2025 · The global Membrane Filtration market size is projected to grow from $19.25 billion in 2025 to $29.97 billion by 2032, exhibiting a CAGR of ...
  56. [56]
    Cholesterol inhibits oxygen permeation through biological membranes
    Sep 13, 2024 · The passive diffusion of oxygen through the cell membrane is, however, strongly affected by lipid composition, which may help to rationalize the ...
  57. [57]
    Solute–Vehicle–Skin Interactions and Their Contribution to ...
    Jun 10, 2025 · Drug permeation across the skin primarily occurs via two main routes: (i) the transepidermal route, where drugs are transported via the SC, and ...Missing: gut | Show results with:gut
  58. [58]
    Mechanistic bottom-up estimation of passive drug absorption from ...
    Objective: The fraction of drug absorbed (Fa) from the intestine is an important parameter to characterize the pharmacokinetics of a drug.
  59. [59]
    Towards a better mechanistic comprehension of drug permeation ...
    Oct 25, 2021 · The process of passive drug absorption from the gastrointestinal tract is still poorly understood and modelled.
  60. [60]
    Human aquaporins: regulators of transcellular water flow - PubMed
    AQPs are crucial for the regulation of water homeostasis, providing selective pores for the rapid movement of water across diverse cell membranes.Missing: permeation | Show results with:permeation
  61. [61]
    Ion Channels and the Electrical Properties of Membranes - NCBI - NIH
    Ion Channels Are Ion-Selective and Fluctuate Between Open and Closed States. Two important properties distinguish ion channels from simple aqueous pores.Missing: permeation | Show results with:permeation
  62. [62]
    Ion channels and ion selectivity Essays in Biochemistry - PMC
    Selective ion channels are essential for the excitability of biological membranes: the action potential is a transient phenomenon that reflects the rapid ...
  63. [63]
    [PDF] soil vapor extraction (sve) treatment technology resource guide
    These volatile contaminants transfer in a vapor phase to the unsaturated zone where soil vapor extraction (SVE) can then capture and remove them. In ...Missing: permeation | Show results with:permeation
  64. [64]
    Permeable Reactive Barrier Technologies for Contaminant ...
    Reactive material is placed in the subsurface to intercept a plume of contaminated ground water which must move through it as it flows, typically under its ...
  65. [65]
    [PDF] Evaluation of Permeable Reactive Barrier Performance
    Permeable reactive barriers (PRB) are developing into an entire new class of technologies for groundwater remediation. A permeable barrier is a porous “barrier” ...Missing: permeation | Show results with:permeation
  66. [66]
    Biofouling of Water Treatment Membranes - PubMed Central - NIH
    Biofouling is a critical issue in membrane water and wastewater treatment as it greatly compromises the efficiency of the treatment processes.
  67. [67]
    [PDF] Municipal Solid Waste Landfills - EPA
    Emissions Analysis: In 2023, this proposal would achieve reductions of 79 Mg NMOC, 12,000. Mg methane, and 308,000 Mg CO2-equivalents1 compared to the baseline.Missing: permeation | Show results with:permeation
  68. [68]
    Geoenvironmental assessment of climate impacts from landfill gas ...
    Methane emissions from landfills contribute 64 Tg CO2-eq. to global climate change (IPCC 2021). While effects of methane have been considered in climate change ...Missing: permeation | Show results with:permeation
  69. [69]
    Assessment of Sieverts Law Assumptions and 'n' Values in ... - NIH
    Oct 12, 2021 · Sieverts' Law suggests that the flux of hydrogen through such membranes is proportional to the difference between the feed and permeate side ...
  70. [70]
    (PDF) Gases in Metals - ResearchGate
    and Tis the temperature in K. Hydrogen solubility in both liquid and solid. pure aluminum follows Sievert's law. That is,. the solubility of hydrogen is ...
  71. [71]
    Diffusion and solubility of hydrogen in palladium and palladium
    Diffusion and solubility of hydrogen in palladium and palladium--silver alloys ... Sieverts Law Empirical Exponent for Pd-Based Membranes: Critical Analysis in ...
  72. [72]
    [PDF] Permeation Rate Equations for Hydrogen and Deuterium in a ...
    Despite that temperature dependence relationship, all calculated permeation rates from the Yoshida constants generate a value less than the experimental value.
  73. [73]
    An Experimental Performance Study of a Catalytic Membrane ...
    Oct 18, 2021 · The values for the effective activation energy were found to be in the range of 12–30 kJ mol−1, which are considerably lower than that of the ...
  74. [74]
    Tritium permeation experiments; design choices and ...
    Tritium permeation is expected to be a major challenge in many locations in the fuel cycle of future fusion reactors. Countless permeation experiments have ...
  75. [75]
    The hydrogen permeation and hydrogen-induced cracking ...
    The influence of H2S partial pressure, gas velocity and chemical composition of the pipes on the hydrogen permeation behavior and HIC occurrence are studied by ...
  76. [76]
    [PDF] AM3 El"W?rnLEMENT - NASA Technical Reports Server
    This equation is known as Sieverts law (3). ... Thee are two main methods for measuring gas solubility. i n metals: i ) Sampling Method. i i ) Hot Volune Method.
  77. [77]
    Hydrogen and Deuterium Solubility, Diffusivity and Permeability from ...
    Jan 21, 2023 · The Sieverts law is valid up to H(D)/M < 0.2 in its ideal form; the absolute values of the hydrogenation/deuteration enthalpy are ΔH(H2) = 85 ± ...<|separator|>
  78. [78]
    The effect of glass transition on gas permeabilities
    The change of activation energy of gas permeability constant at the glass transition temperature (65°C) was observed with N2, Ar, O2, and CO2, but no change ...
  79. [79]
    Effect of physical aging on the gas transport and sorption in PIM-1 ...
    This process, referred to as 'physical aging', results in a reduced gas permeability coupled often with an increased selectivity.