Fact-checked by Grok 2 weeks ago

Semipermeable membrane

A semipermeable membrane, also known as a selectively permeable membrane, is a thin barrier that permits the passage of certain molecules and ions—such as , , and —while restricting others, like larger polar molecules and charged particles, based on factors including size, charge, and lipid solubility. This selective permeability is fundamental to various functions, enabling processes like , where moves across the membrane from areas of lower solute concentration to higher ones to maintain . The structure of a biological semipermeable membrane, such as the plasma membrane of eukaryotic cells, consists of a phospholipid bilayer embedded with proteins, , and carbohydrates, forming a that allows dynamic interactions. Phospholipids arrange with hydrophilic heads facing outward toward aqueous environments and hydrophobic tails inward, creating a nonpolar core that repels water-soluble substances but facilitates the of nonpolar molecules. Integral membrane proteins, including channels and carriers, further enhance selectivity by providing specific pathways for and solute transport, while peripheral proteins assist in signaling and structural support. Semipermeable membranes play critical roles in biological systems for homeostasis, nutrient uptake, waste removal, and cellular communication, as well as in artificial applications like water purification and medical filtration. For instance, in physiology, they regulate osmotic balance in kidneys to filter blood and concentrate urine, support nerve impulse transmission via ion gradients, and protect cells from environmental fluctuations in tonicity, preventing swelling or shrinkage. Disruptions in membrane permeability can lead to conditions like edema or cell lysis, underscoring their essentiality in health and disease.

Fundamentals

Definition and Properties

A semipermeable membrane is a selective barrier that permits the passage of certain molecules or ions while restricting others, primarily based on properties such as molecular size, , or within the membrane material. This distinguishes it from fully permeable barriers, which allow unrestricted flow of all substances, and impermeable barriers, which prevent any passage. The selective nature enables the membrane to regulate , maintaining distinct compositions on either side. The concept originated from the work of Scottish chemist Thomas Graham, who in 1861 described experiments using as a barrier for separating crystalloids from colloids through , a process he termed . Graham's observations highlighted the membrane's ability to allow small solutes like salts to pass while retaining larger particles, laying the foundation for understanding selective . Over time, the terminology shifted from "dialysis membrane" to "semipermeable membrane" as researchers recognized its broader applicability beyond just setups. Key properties of semipermeable membranes include , which governs the exclusion of larger molecules based on pore size; surface hydrophilicity or hydrophobicity, influencing the and partitioning of solutes; thickness, often ranging from 5 to 10 nm in biological contexts; and composition, typically involving amphiphilic in natural forms or synthetic polymers like . These attributes collectively determine the membrane's selectivity and efficiency in transport. Natural examples, such as bilayers, exemplify these traits in . The rate of substance transport across a semipermeable membrane is quantified by the permeability equation, which describes the steady-state under a concentration . The total J (in mol/s) is given by: J = P \cdot A \cdot \Delta C where P is the permeability coefficient (in m/s), A is the membrane area (in m²), and \Delta C is the concentration difference across the (in mol/m³). This equation arises from Fick's first law of , j = -D \frac{dC}{dx}, where j is the areal (mol/m²·s), D is the coefficient (m²/s), and \frac{dC}{dx} is the concentration . For a thin, homogeneous of thickness L, assuming linear and solute partitioning, the approximates \Delta C / L, and the permeability coefficient incorporates the K (dimensionless, reflecting in the ) as P = \frac{D K}{L}, yielding the areal j = P \Delta C; multiplying by area gives the total . The units ensure dimensional consistency: m/s × m² × mol/m³ = mol/s. This model provides a foundational for predicting behavior across diverse materials.

Mechanisms of Permeability

Semipermeable membranes enable selective through passive and active mechanisms, allowing solutes to cross based on concentration gradients or input while restricting others. Passive mechanisms rely on the natural tendency of molecules to move from higher to lower concentrations without energy expenditure, primarily through , , and . across the membrane follows Fick's first law, which states that the J of a solute is proportional to the negative concentration gradient across the membrane: J = -D \frac{dC}{dx}, where D is the diffusion coefficient, C is the concentration, and x is the . The diffusion coefficient D quantifies the solute's mobility and depends on factors such as molecular , , and the membrane's properties; smaller, nonpolar molecules like oxygen diffuse more readily through bilayers due to higher D values, approximately $10^{-5} cm²/s in biological membranes. Facilitated diffusion enhances passive movement of polar or charged solutes through protein channels or carriers without direct energy input, relying on concentration gradients. For instance, s are integral membrane proteins forming water-specific pores that allow rapid of water molecules, increasing permeability by orders of magnitude compared to unassisted diffusion; each channel can conduct up to 3 billion water molecules per second. Osmosis, a specific form of passive transport for water, occurs when a concentration gradient of solutes drives water across the membrane toward the higher solute side. The osmotic pressure \pi required to prevent this flow is given by the van 't Hoff equation: \pi = iMRT, where i is the van 't Hoff factor accounting for solute dissociation, M is the molar concentration, R is the gas constant, and T is the absolute temperature. This equation derives from the analogy between osmotic pressure in dilute solutions and the pressure of an ideal gas, as van 't Hoff established by treating solute particles as exerting pressure on the membrane similar to gas molecules on a container. In semipermeable systems, water permeability via osmosis can reach rates up to 10⁻³ cm/s in hydrated membranes, far exceeding simple diffusion for other solutes. Active mechanisms counteract gradients using cellular energy, such as primary . Primary , such as ATP-driven pumps, uses of ATP to move ions against gradients; the sodium-potassium pump (Na⁺/K⁺-) exemplifies this by exchanging three Na⁺ ions out for two K⁺ ions in per ATP molecule hydrolyzed, maintaining electrochemical balances essential for membrane function. Permeability is modulated by several factors, including molecular , electrostatic interactions, and solubility in the membrane matrix. Many semipermeable membranes exhibit a cutoff, typically allowing free passage of solutes below 500-1000 while restricting larger ones, as seen in systems where pore diameters determine exclusion limits. Electrostatic interactions between charged solutes and membrane components, such as charged lipid headgroups or fixed charges in polymeric membranes, can either enhance or hinder transport; for example, repulsive forces between like charges reduce permeability for anions in negatively charged membranes. The , defined as the ratio of solute concentration in the membrane phase to the aqueous phase, governs solubility-driven permeability; higher lipophilicity (log P > 0) favors partitioning into hydrophobic membrane cores, increasing overall flux according to the solubility-diffusion model. Experimental measurement of permeability employs techniques like (FRAP) and dedicated permeability assays. FRAP involves bleaching a region of fluorescently labeled components with a and monitoring recovery via lateral of unbleached molecules, yielding diffusion coefficients and permeability estimates; this method has quantified mobility in bilayers at rates of 1-10 µm²/s. Permeability assays, such as parallel artificial permeability assays (PAMPA), measure solute flux across lipid-impregnated barriers using concentration gradients tracked by UV spectroscopy or , providing apparent permeability coefficients (P_app) in the range of 10^{-6} to 10^{-4} cm/s for drug-like compounds. These approaches ensure quantitative assessment of transport kinetics in both model and native membranes.

Biological Contexts

Phospholipid Bilayer

The phospholipid bilayer forms the foundational structure of biological semipermeable membranes, primarily composed of amphipathic such as , which feature hydrophilic phosphate heads and hydrophobic tails. These self-organize into a bilayer with heads facing the aqueous environments on both sides and tails sequestered in the interior, while molecules intercalate between phospholipids to modulate and stability, preventing excessive rigidity or permeability at physiological temperatures. Integral proteins, embedded within the bilayer and often comprising up to 50% of the membrane's mass, contribute to its functional complexity by spanning the hydrophobic core. The organization of the bilayer is best described by the , proposed by Singer and Nicolson in 1972, which portrays the membrane as a dynamic, two-dimensional fluid where and proteins diffuse laterally within the plane. This model emphasizes the bilayer's fluidity, with lipid rafts—specialized microdomains enriched in and —facilitating localized protein interactions, and a striking in lipid distribution, such as predominantly localized to the inner leaflet. The arises from enzymatic regulation and maintains the membrane's integrity and signaling capabilities. Physically, the bilayer exhibits a thickness of approximately 4-5 nm, determined by the length of the chains and influenced by chain saturation. Lateral diffusion rates for and proteins typically range around 10^{-8} cm²/s, reflecting the state that allows rapid molecular movement without disrupting the barrier. The bilayer undergoes phase transitions from a to a liquid crystalline , often between 20-40°C depending on composition, which affects packing density and permeability; broadens this transition to sustain fluidity across temperature variations. Self-assembly of the phospholipid bilayer is driven by the , where amphipathic molecules minimize unfavorable interactions with water by aggregating tails inward, leading to spontaneous bilayer formation that lowers the system's (ΔG = ΔH - TΔS), primarily through an increase in (ΔS) as water molecules are released from ordered hydration shells around hydrophobic tails, with minimal enthalpic change (ΔH). This ensures the bilayer's stability in aqueous environments without external input.

Cellular Transport and Communication

Semipermeable membranes in cells facilitate the selective of molecules essential for uptake, waste elimination, and maintenance of internal . Simple allows small nonpolar molecules, such as oxygen (O₂) and (CO₂), to cross the phospholipid bilayer directly down their concentration gradients without energy input or protein assistance. In contrast, facilitated employs carrier proteins for polar or larger molecules that cannot freely permeate the ; for instance, glucose enters cells via (GLUT) proteins, which undergo conformational changes to bind and translocate glucose across the membrane. Vesicular transport mechanisms, including and , enable the movement of macromolecules and particles by enclosing them in membrane-bound vesicles that fuse with or bud from the plasma membrane, bypassing the selective barrier for bulk transfer. Beyond basic exchange, semipermeable membranes play a pivotal role in intercellular communication by hosting receptor proteins and specialized junctions. G-protein coupled receptors (GPCRs), embedded in the membrane, detect extracellular signals like hormones and s, triggering intracellular cascades via G-protein activation upon binding. Gap junctions, formed by proteins, create direct cytoplasmic channels between adjacent cells, permitting the passage of ions and small metabolites to synchronize activities such as in contraction. In neuronal communication, synaptic signaling occurs across the presynaptic and postsynaptic membranes, where neurotransmitter release from vesicles in the presynaptic diffuses across the synaptic cleft to bind receptors on the postsynaptic membrane, modulating ion channels and propagating electrical signals. A key example of across semipermeable membranes is the sodium-potassium pump (Na⁺/K⁺-ATPase), which maintains ion gradients crucial for and cellular volume. This hydrolyzes ATP to actively export three sodium ions (Na⁺) from the while importing two ions (K⁺), operating through a cycle of and that alternates the protein's for ions and its in the membrane. In , exemplifies membrane-mediated communication, where membrane-bound sensor kinases detect accumulated autoinducer molecules, triggering coordinated gene expression for behaviors like formation once a is reached. Energy considerations in ion transport across semipermeable membranes are governed by electrochemical gradients, which combine chemical concentration differences and electrical potential. The Nernst equation quantifies the equilibrium potential (E) for a specific ion, beyond which net diffusion ceases: E = \frac{RT}{zF} \ln \left( \frac{[\text{ion}]_{\text{out}}}{[\text{ion}]_{\text{in}}} \right) Here, R is the gas constant, T is temperature in Kelvin, z is the ion's valence, F is Faraday's constant, and [ion] denotes concentration outside and inside the cell; for potassium (K⁺, z=1), typical mammalian values yield an E_K of about -90 mV, driving passive K⁺ efflux that contributes to the resting membrane potential. Active pumps like Na⁺/K⁺-ATPase counteract these gradients, consuming ATP to sustain nonequilibrium states essential for signaling and excitability.

Osmotic Pressure Management

In biological systems, semipermeable membranes facilitate , the passive movement of water across the membrane from regions of lower solute concentration to higher solute concentration, driven by differences. In plant cells, this influx of water generates against the rigid , which can reach up to 1.0 in well-hydrated tissues, maintaining structural integrity and driving cell expansion. In hypotonic environments, where external solute concentration is lower than inside the , water enters the , increasing volume and turgor; conversely, in hypertonic conditions, water exits, reducing turgor and potentially leading to in plants. Cells manage osmotic stress through regulatory mechanisms that adjust permeability and intracellular solute levels. Aquaporins, water channel proteins embedded in semipermeable membranes, are dynamically regulated during osmotic shock; for instance, their trafficking to the plasma membrane increases under hypotonic stress to enhance water influx, while or endosomal internalization reduces activity during hypertonic stress to conserve water. Additionally, plants accumulate compatible solutes such as during drought-induced osmotic stress, which lowers cellular without disrupting , thereby restoring osmotic balance and protecting against . levels can rise dramatically, up to 100-fold in some species, aiding turgor maintenance and scavenging generated by stress. Pathological disruptions occur when osmotic imbalances overwhelm these mechanisms. In hypotonic environments, excessive water influx causes animal cells, such as red blood cells, to swell and lyse (), rupturing the membrane due to unchecked volume expansion. In hypertonic conditions, water efflux leads to , where red blood cells shrink and develop a spiky, contracted appearance as the condenses. In the , semipermeable membranes in the loop of Henle employ a countercurrent multiplier system to manage osmotic gradients; the descending limb is permeable to water, allowing equilibration with the hypertonic medullary , while the ascending limb actively extrudes solutes, amplifying the gradient up to 1200 mOsm/L at the loop's tip to facilitate concentration without cellular damage. Quantitative analysis of osmotic effects on cellular volume relies on the van't Hoff equation, which relates osmotic pressure \pi to solute amount n, temperature T, and volume V: \pi = \frac{nRT}{V} Rearranging gives the ideal volume V = \frac{nRT}{\pi}, assuming dilute solutions and ideal behavior. For red blood cells in physiological saline (approximately 300 mOsm/L, \pi \approx 7.6 atm at 37°C), volume remains stable at about 90-100 \mum³. In hypotonic media (e.g., 150 mOsm/L, \pi \approx 3.8 atm), predicted volume doubles to roughly 180-200 \mum³, leading to swelling and potential lysis if exceeding 1.4 times original volume; in hypertonic media (e.g., 600 mOsm/L, \pi \approx 15.2 atm), volume halves to 45-50 \mum³, causing . These changes highlight how semipermeable membranes dictate volume under osmotic stress.

Artificial and Synthetic Applications

Reverse Osmosis

(RO) is a pressure-driven that utilizes a semipermeable membrane to purify by forcing molecules, typically , from a higher-concentration to a lower-concentration one, counter to the natural osmotic gradient. This requires applying hydraulic pressure exceeding the of the feed , with typical operating pressures ranging from 10 to 100 depending on the feed salinity—for instance, 55 to 85 for seawater . The process effectively removes dissolved salts, organics, and , producing high-purity permeate while concentrating contaminants in the reject stream. RO membranes are predominantly thin-film composite (TFC) structures, consisting of a selective active layer deposited on a microporous support layer, which provides mechanical stability and high flux. These membranes achieve salt rejection rates of 95-99% for (NaCl) under standard conditions, enabling efficient of brackish or feeds. The layer's nanoscale pores and negative charge selectively permit passage while repelling hydrated ions, though performance can vary with , temperature, and feed composition. In applications, RO dominates seawater desalination, accounting for approximately 70% of global installed capacity as of , or about 77 million cubic meters per day out of a total around 110 million cubic meters per day. It is also widely used in to reclaim from industrial effluents and municipal sources, recovering up to 75-85% of the feed in optimized systems. has improved significantly, with specific energy consumption for seawater RO typically around 3 kWh per cubic meter, achieved through energy recovery devices like pressure exchangers that energy from the high-pressure reject. A major challenge in RO systems is , which reduces and increases energy use over time. Scaling occurs due to mineral precipitation, such as or silica, on the membrane surface under high recovery conditions, while results from microbial growth and extracellular polymeric substances that form biofilms. Maintenance involves periodic chemical cleaning, including soaks at low pH (typically 2-4) to dissolve inorganic scales without damaging the layer, often alternated with alkaline cleaners for organic and biological foulants. Pretreatment strategies, like antiscalant dosing and , are essential to mitigate and extend membrane life to 5-7 years. As of 2025, advancements in RO technology include the development of novel thin-film membranes and improved systems, enabling specific energy consumption as low as 2.5 kWh per cubic meter in cutting-edge plants.

Dialysis and Filtration

relies on the passive of solutes across a semipermeable membrane driven by concentration gradients, separating small molecules from larger ones without requiring external pressure. In laboratory settings, this process typically employs regenerated or tubing with pore sizes ranging from 10 to 20 Å, allowing molecules below 500 Da, such as salts and , to pass while retaining proteins and macromolecules. This mechanism aligns with the fundamental principles of through semipermeable barriers, where solutes move from higher to lower concentrations until is achieved. In medical contexts, hemodialysis uses artificial semipermeable membranes in dialyzers to treat kidney failure by removing waste products from blood. Blood flows through one side of the membrane while dialysate flows countercurrently on the other, facilitating the diffusion of small solutes like urea, with typical clearance rates of approximately 200 mL/min for urea under standard conditions of blood flow around 300-400 mL/min and dialysate flow of 500 mL/min. Peritoneal dialysis, conversely, leverages the natural semipermeable peritoneum—the abdominal cavity's lining—as the membrane, where dialysate is introduced into the peritoneal space to draw out solutes via diffusion across this biological barrier. This approach enables continuous ambulatory peritoneal dialysis, performed multiple times daily, providing a gentler alternative for patients unsuitable for hemodialysis. Laboratory applications of dialysis include equilibrium dialysis, which determines protein-ligand binding affinities by separating bound and free across a semipermeable membrane until concentrations equilibrate, allowing of constants. , a related , employs semipermeable membranes under mild pressure or to concentrate samples by retaining macromolecules while permitting and small solutes to pass, commonly used in workflows. To maintain performance, membranes undergo regeneration through chemical cleaning to remove and deposits, such as using dilute solutions (e.g., 0.5-1% ) followed by rinsing to eliminate contaminants without significantly degrading the structure. Backflushing, where fluid is reversed through the membrane, aids in dislodging and extending , particularly in reusable systems. These techniques can prolong the operational life of membranes to 1-3 years in reprocessing programs, depending on material and usage frequency, by mitigating and preserving permeability.

Other Membrane Types

Nanofiltration Membranes

Nanofiltration membranes represent an intermediate pressure-driven filtration technology positioned between and , characterized by pore sizes typically ranging from 1 to 10 . These membranes selectively reject multivalent ions such as Ca²⁺ and SO₄²⁻ while permitting the passage of monovalent ions like Na⁺ and Cl⁻, enabling targeted removal of organics and salts without complete demineralization. Flux rates for nanofiltration processes generally fall between 20 and 50 L/m²/h under operational pressures of 5 to 15 bar, depending on feed composition and design. The selectivity of nanofiltration membranes is often enhanced by incorporating charged polymers, such as sulfonated polyethersulfone or sulfonated polyether ether ketone (SPEEK), which introduce electrostatic repulsion mechanisms. This charge-based functionality relies on Donnan exclusion, where the fixed charges on the surface repel similarly charged , complementing steric hindrance from the to achieve high rejection rates for divalent species. Such materials allow for tunable permeability, making nanofiltration suitable for applications requiring precise fractionation. In practical applications, nanofiltration is widely employed for by removing hardness-causing divalent ions, as well as in pharmaceutical purification to separate active ingredients from impurities and solvents. In the sector, it facilitates the concentration and demineralization of , notably through removal from permeate, enhancing product yield and quality. The global market for nanofiltration membranes was estimated at approximately $1.2 billion in 2025, according to one analysis, though estimates vary across reports. Compared to , nanofiltration offers significant advantages, including lower operating pressures of 5 to 15 bar that reduce energy consumption by up to 50% in suitable scenarios. This enables partial softening and organic removal without the need for full , preserving beneficial monovalent salts and lowering overall costs for targeted treatments.

Gas Permeation Membranes

Gas permeation membranes are semipermeable barriers designed specifically for separating gas mixtures based on differences in gas and , enabling efficient industrial purification processes. These membranes operate primarily through the solution- , where gas molecules first dissolve into the upstream surface of the according to their coefficient S (often described by ), then diffuse across the driven by a , with the rate governed by the coefficient D. The overall permeability P is given by the product P = D \times S, which quantifies the of a gas through the under a given difference. This model applies to non-porous, dense membranes and is particularly effective for light gases like and CO2, which exhibit high permeability due to their small molecular size facilitating rapid diffusion (D up to orders of magnitude higher than larger gases) and favorable in certain polymers. Polymeric materials dominate gas permeation membranes for their cost-effectiveness and ease of fabrication, with polydimethylsiloxane (PDMS) serving as a representative example for O2/N2 separation due to its exceptionally high gas permeability—often exceeding 500 Barrer for O2—stemming from its rubbery, flexible structure that enhances both solubility and diffusivity. Inorganic membranes, such as those based on zeolites, offer superior selectivity for challenging separations like H2/CH4, achieving ideal selectivities greater than 100 through molecular sieving effects in their uniform micropores (typically 0.3–0.4 nm), which restrict larger CH4 molecules while allowing H2 to permeate freely. For instance, DDR-type zeolite membranes have demonstrated H2/CH4 selectivities up to 207 under practical conditions, far surpassing polymeric limits for this pair. Key applications of gas permeation membranes include natural gas sweetening, where they remove CO2 and H2S to meet specifications, with polymeric systems like processing feeds up to 1200 million standard cubic feet per day (approximately 34 million m³/day) in large installations. Air separation utilizes these membranes to produce high-purity N2 (up to 99.9%) by selectively permeating , supporting applications in inerting and , while hydrogen recovery in refineries employs them to purify from reformate streams, recovering over 90% of with purities exceeding 99%, as an energy-efficient alternative to . Globally, these applications have scaled significantly, with thousands of membrane units operational across industries by the 2020s. Despite their advantages, gas permeation membranes face challenges such as plasticization, where high-pressure exposure to penetrants like CO2 causes polymer chain swelling, increasing and permeability but degrading selectivity by up to 50% over time. This issue is particularly acute in natural gas sweetening under elevated CO2 partial pressures (>10 ), necessitating material modifications like crosslinking to mitigate swelling. Additionally, a fundamental exists between permeability and selectivity, encapsulated by the Robeson upper bound—an empirical plotting selectivity against permeability for gas pairs like CO2/CH4 or /N2, where the 2008 upper bound reflects dual-mode and limitations in glassy polymers, limiting most materials to below this line unless innovative structures like thermally rearranged polymers are employed. Overcoming this bound remains a key focus to enhance efficiency in high-volume separations.

References

  1. [1]
    Cell Membranes - The Cell - NCBI Bookshelf - NIH
    Cell membranes separate the cell from its environment, are made of phospholipid bilayers, and have proteins that perform specialized functions.
  2. [2]
    Physiology, Osmosis - StatPearls - NCBI Bookshelf - NIH
    Mar 13, 2023 · Biological membranes act as semipermeable barriers and allow for the process of osmosis to occur. Osmosis underlies almost all major processes ...
  3. [3]
    Semipermeable Membrane - an overview | ScienceDirect Topics
    A semipermeable membrane is defined as a type of membrane that selectively allows the passage of certain drug compounds while preventing the movement of local ...
  4. [4]
    Transport of Small Molecules - The Cell - NCBI Bookshelf - NIH
    The internal composition of the cell is maintained because the plasma membrane is selectively permeable to small molecules. Most biological molecules are ...
  5. [5]
    The history of dialysis - Fresenius Medical Care
    The first scientific descriptions of these procedures dates back to the 19th century and came from the Scottish chemist Thomas Graham, who became known as the ...
  6. [6]
    Dialysis - Chemistry LibreTexts
    Jan 29, 2023 · A semipermeable membrane is a membrane that lets some molecules to pass through it while not letting others (Figure 1 ). Examples of ...Missing: history | Show results with:history
  7. [7]
    Thomas Graham (1805–1869) — The “Father” of Dialysis
    ... In 1854, Thomas Graham described the ability to separate solutes through a semipermeable membrane via diffusion principles. (Cameron, 2012) . ...Missing: definition | Show results with:definition
  8. [8]
    Semipermeable Membrane - an overview | ScienceDirect Topics
    The barrier to solute movement is largely provided by the membrane's hydrophobic core, a very thin (∼40 Å thick), oily layer. The inherent permeability of this ...
  9. [9]
    Cell Membrane - an overview | ScienceDirect Topics
    Cell membrane is a bio-complex with a thickness of 5–10 nm, and is composed of various bio-macromolecules (Lamparter and Galic, 2020; Singer and Nicolson ...
  10. [10]
    What is the permeability of the cell membrane? - Bionumbers book
    For transport across membranes, these ideas have been codified in the simple equation (for neutral solute) j= – p·(cin-cout), where j is the net flux into the ...Missing: PA ΔC
  11. [11]
    Intrinsic Membrane Permeability to Small Molecules
    Assuming that Kp is identical at both interfaces leads to the common definition of membrane permeability Pm = Kp·Dm/d. (14). where Δcs is the concentration ...
  12. [12]
    Molecular Simulation/Membrane permeability - Wikibooks
    The flux of a solute across a membrane is calculated from its permeability coefficient, P, and the concentration gradient across the bilayer (ΔC).Missing: PA units semipermeable
  13. [13]
    Molecular simulation of nonfacilitated membrane permeation
    For a constant set of conditions and membrane composition, the permeability coefficient provides a measure of the intrinsic membrane permeability of a solute.1.2. Permeability Models · 3. Practical Aspects · 5.1. Polarizable Models
  14. [14]
    Membrane Permeability - an overview | ScienceDirect Topics
    Membrane permeability is defined according to Fick's first law as a composite parameter that lumps the drug membrane diffusivity, the drug membrane ...
  15. [15]
    The origin and present status of Fick's diffusion law - ACS Publications
    Virtually all experimental papers on diffusion are concerned, in the first instance, with the determination of diffusion coefficients defined in a manner ...
  16. [16]
    [PDF] Osmotic pressure and chemical equilibrium - Nobel Prize
    1901 J.H.VAN'T HOFF remembers the vital role played by osmotic pressure in plant and animal life. The membranes of the cell are in fact permeable to water but ...
  17. [17]
    Aquaporins - PMC - PubMed Central - NIH
    The primary function of most aquaporins is to transport water across cell membranes in response to osmotic gradients created by active solute transport. Because ...
  18. [18]
    [PDF] Peter Agre - AQUAPORIN WATER CHANNELS - Nobel Prize
    Discovery of the aquaporins has provided the explanation for selective trans- port of water across the plasma membranes of cells without allowing protons. 202.
  19. [19]
    [PDF] Jens C. Skou - Nobel Lecture
    It was shown that the active efflux of sodium was coupled to an influx of potassium, a pump, and that the substrate for the transport was energy rich phosphate ...
  20. [20]
    Membrane Permeability - an overview | ScienceDirect Topics
    According to the solubility–diffusion model, the membrane permeability coefficient can be related to the partition coefficient between water and the highest ΔG ...
  21. [21]
    Local Partition Coefficients Govern Solute Permeability of ... - NIH
    The permeability of membranes from natural lipid mixtures to both hydrophilic and hydrophobic solutes monotonously decreases with cholesterol concentration.
  22. [22]
    Predicting a Drug's Membrane Permeability: A Computational Model ...
    Apr 28, 2017 · LogP is a crucial factor governing passive membrane partitioning; an increase in LogP enhances permeability.
  23. [23]
    What's past is prologue: FRAP keeps delivering 50 years later - NIH
    Fluorescence recovery after photobleaching (FRAP) has emerged as one of the most widely utilized techniques to quantify binding and diffusion kinetics of ...
  24. [24]
    A Method for Determination and Simulation of Permeability ... - NIH
    Feb 23, 2018 · In this article, a method is presented to directly measure the permeability of a barrier within a membrane insert system, where a skin model ...
  25. [25]
    Real-Time Parallel Artificial Membrane Permeability Assay Based ...
    Nov 2, 2020 · Parallel artificial membrane permeability assay (PAMPA) is a screening tool for the evaluation of drug permeability across various ...
  26. [26]
    Structure of the Plasma Membrane - The Cell - NCBI Bookshelf
    Cholesterol will not form a membrane by itself, but inserts into a bilayer of phospholipids with its polar hydroxyl group close to the phospholipid head groups ...
  27. [27]
    The Fluid Mosaic Model of the Structure of Cell Membranes - Science
    A fluid mosaic model is presented for the gross organization and structure of the proteins and lipids of biological membranes.
  28. [28]
    The Lipid Bilayer - Molecular Biology of the Cell - NCBI Bookshelf
    This change of state is called a phase transition, and the temperature at which it occurs is lower (that is, the membrane becomes more difficult to freeze) if ...Missing: rates | Show results with:rates
  29. [29]
    Thermodynamics of Phospholipid Self-Assembly - PMC - NIH
    The driving force for self-assembly of all biological membranes is the hydrophobic effect, which is contributed by the aliphatic lipid chains. The polar ...
  30. [30]
    Glucose transporters: physiological and pathological roles - PMC
    Facilitative glucose transporters (GLUTs) The GLUTs transport glucose across the plasma membrane by means of a facilitated diffusion mechanism.
  31. [31]
    Biochemistry, G Protein Coupled Receptors - StatPearls - NCBI - NIH
    The G protein is associated with the plasma membrane at the cytoplasmic side, connecting the GPCR to either enzymes or ion channels. In some cases, G proteins ...
  32. [32]
    Gap-junction-mediated cell-to-cell communication - PubMed
    These channels are densely packed into plasma-membrane spatial microdomains termed "gap junctions" and allow cells to exchange ions and small molecules directly ...
  33. [33]
    Physiology, Synapse - StatPearls - NCBI Bookshelf
    Mar 27, 2023 · These membranes possess channels formed by proteins known as connexins, which allow the direct passage of current from 1 neuron to the next and ...
  34. [34]
    Physiology, Sodium Potassium Pump - StatPearls - NCBI Bookshelf
    Mar 13, 2023 · [3][4] The Na+K+-ATPase pump helps to maintain osmotic equilibrium and membrane potential in cells. The sodium and potassium move against the ...
  35. [35]
    Bacterial Quorum Sensing: Its Role in Virulence and Possibilities for ...
    Quorum sensing is a process of cell–cell communication that allows bacteria to share information about cell density and adjust gene expression accordingly.
  36. [36]
    Ion Channels and the Electrical Properties of Membranes - NCBI - NIH
    A simple but very important formula, the Nernst equation, expresses the equilibrium condition quantitatively and, as explained in Panel 11-2, makes it possible ...
  37. [37]
    Structure and Function of Na,K-ATPase-The Sodium-Potassium Pump
    Dec 29, 2021 · Na,K-ATPase is an ubiquitous enzyme actively transporting Na-ions out of the cell in exchange for K-ions, thereby maintaining their concentration gradients ...
  38. [38]
    Loss of Stability: A New Look at the Physics of Cell Wall Behavior ...
    Typical plant cell turgor pressures in the range of 0.3 to 1.0 MPa translate into between 10 and 100 MPa of tensile stress in the walls.
  39. [39]
    Novel Regulation of Aquaporins during Osmotic Stress - PMC
    Endosomal trafficking of plant AQP may play a role in the regulation and turnover of these proteins at their target membrane under conditions of osmotic stress.
  40. [40]
    Regulation of aquaporins in plants under stress - Biological Research
    Jan 16, 2018 · The presence of AQPs in the plasmalemmasomes suggests that they may play a significant role in the osmotic balance and turgor maintenance of ...Background · Structure And Function Of... · Stress Response Of...<|separator|>
  41. [41]
    Role of proline under changing environments: A review - PMC - NIH
    The glutamate pathway accounts for major proline accumulation during osmotic stress. The proline is synthesized from glutamatic acid via intermediate ∆'- ...
  42. [42]
    Spatially variant red blood cell crenation in alternating current non ...
    Red blood cells crenate in response to local medium hypertonicity, which indicates local ion concentration gradients in non-uniform electric fields, but not in ...
  43. [43]
    The Physiology of Urinary Concentration: an Update - PubMed Central
    Countercurrent multiplication in the loops of Henle generates the cortico-medullary osmotic gradient necessary for water reabsorption, and countercurrent ...
  44. [44]
    Reverse Osmosis - an overview | ScienceDirect Topics
    Reverse osmosis is a membrane-based, pressure-driven process where water moves through a semipermeable membrane from a concentrated to a weaker solution. It is ...
  45. [45]
    Reverse Osmosis Desalination Process - Lenntech
    Before entering the seawater reverse osmosis membranes, clarified seawater is pressurized by the High Pressure Pump typically between 55 and 85 bars, depending ...
  46. [46]
    Reverse Osmosis - The Basics - Puretec Industrial Water
    Reverse osmosis (RO) removes contaminants from water by pushing it through a semi-permeable membrane under pressure, removing dissolved solids.
  47. [47]
    Advanced fabrication and characterization of thin-film composite ...
    Apr 30, 2025 · The maximum salt rejection and water flux for the prepared thin film composite reverse osmosis membrane were 98.6% and 19.1 L/m2 h, respectively ...
  48. [48]
    High-Performance Polyamide Reverse Osmosis Membrane ... - MDPI
    Feb 14, 2023 · The performance of RFRO membranes was evaluated by a cross-flow desalination test with a 2000 ppm NaCl solution as feed. Figure 4a shows the ...
  49. [49]
    Recent advances in thin film composite (TFC) membrane development
    Feb 17, 2025 · At the pH value of 11, the PA membrane showed a high rejection of 99.5% for Na2SO4, 32% for NaCl, and 97.8% for methyl orange, but low water ...
  50. [50]
    A comprehensive review of reverse osmosis desalination
    In 2021, the total installed desalination capacity reached approximately 115 million m3/d (Mm³/d), with around 88 million m3/d (77 % of the total) attributed to ...
  51. [51]
    A comprehensive review of energy consumption of seawater reverse ...
    Aug 9, 2025 · It has been reported that the specific energy consumption (SEC) of the seawater reverse osmosis (RO) process is 2.5-4.0 kWh/m 3 , much less ...
  52. [52]
    Reverse osmosis membrane biofouling: causes, consequences and ...
    Sep 15, 2022 · An alkaline clean is often performed before or after an acid clean when targeting removal of both organic and inorganic foulants.
  53. [53]
    [PDF] Evaluation of fouling and chemical cleaning of reverse osmosis ...
    Jan 8, 2025 · Then the RO membrane was cleaned with citric acid for 45 min followed by rinsing with the RO permeate before commencing the permeability test.
  54. [54]
    Can I use citric acid to descale/clean a reverse osmosis (RO ...
    Citric acid is only useful for calcium carbonate scaling and re-tightening membrane pores after high pH cleaning. It can worsen membrane performance and is not ...
  55. [55]
    Nanofiltration Membranes for the Removal of Heavy Metals ... - NIH
    Sep 12, 2023 · This sophisticated filtering method uses semi-permeable membranes with pore sizes of 1–10 nm, permitting the selective removal of heavy metal ...
  56. [56]
    Nanofiltration as an advanced wastewater treatment technique
    Apr 15, 2025 · The unique pore size (1–10 nm) of NF membranes enables selective permeability [56]. This structure allows for the rejection of multivalent ions ...Missing: definition | Show results with:definition
  57. [57]
    Study on the nanofiltration membrane fouling control and cleaning ...
    The initial membrane flux of the nanofiltration membrane was 28.15 L/(m2. h), which decreased to 22.04/(m2. h) after four rounds of pickling, a decrease of 22%.Missing: m²/ | Show results with:m²/
  58. [58]
    PES/SPAES blend membranes for nanofiltration - ScienceDirect.com
    The addition of SPEEK leads to a more open membrane structure and consequently higher solvent fluxes. Due to the Donnan exclusion, the blend membranes showed ...
  59. [59]
    Optimizing Nanofiltration Membrane Layer-by-Layer Modification
    Jun 26, 2025 · (6) In NF membranes, the rejection of ions relies on dielectric exclusion, Donnan (charge) exclusion, and size (steric) exclusion.
  60. [60]
    What Is Nanofiltration In Water Treatment? - Molewater
    Sep 23, 2025 · Nanofiltration's ability to remove water hardness lies between that of softeners and reverse osmosis. It can partially remove calcium and ...
  61. [61]
    NF Lactose Concentration Case Study - Synder Filtration
    Nanofiltration technology is commonly used after ultrafiltration to further concentrate acid whey UF permeate for whey processing applications within the dairy ...
  62. [62]
    Nanofiltration Membrane Market Size, and Growth Report, 2032
    The nanofiltration membrane market size is estimated at USD 1208.7 million for 2025, and it will grow by 11.0% during 2026–2032, to reach USD 2504.7 million by ...
  63. [63]
    Nanofiltration: How Compares to Reverse Osmosis - Hydramem
    Nanofiltration operates at lower pressures than reverse osmosis, making it more energy-efficient for applications that don't require complete desalination.
  64. [64]
    Nanofiltration (NF) Membranes: An In-Depth Guide - J Mark Systems
    Dec 24, 2024 · By operating at lower pressures, they consume less energy than RO systems, translating to reduced operating costs and lower environmental impact ...
  65. [65]
    The solution-diffusion model: a review - ScienceDirect.com
    The solution-diffusion model has emerged over the past 20 years as the most widely accepted explanation of transport in dialysis, reverse osmosis, gas ...
  66. [66]
    Membrane Basics - Liquid Contacting and Gas Separation
    O2/N2 Enrichment ... Silicone, chemically known as polydimethylsiloxane (PDMS), is among the most gas permeable dense polymeric membrane materials available.
  67. [67]
    [PDF] Adsorption and Separation in Small-Pore Zeolites - DiVA portal
    The high permeance was paired with a H2/CH4 selectivity of 207, markedly higher than previously reported for DDR membranes. Furthermore, a mass transfer model ...
  68. [68]
    Membrane gas separation applications in natural gas processing
    This paper considers the current and future potential of polymeric membranes in acidic gas removal, heavy hydrocarbon recovery, water dehydration as well as ...
  69. [69]
    Future Directions of Membrane Gas Separation Technology
    Another, even larger application of hydrogen-permeable membranes exists for hydrogen recovery in refineries. The demand for hydrogen in refineries is ...Missing: sweetening | Show results with:sweetening<|separator|>
  70. [70]
    Physical Aging, Plasticization and Their Effects on Gas Permeation ...
    Aug 29, 2015 · The first phase is documented and characterized by a rapid permeability decline over ∼15 days and driven by excess nonequilibrium free volume ...
  71. [71]
    The upper bound revisited - ScienceDirect.com
    The upper bound correlation can be used to qualitatively determine where the permeability process changes from solution-diffusion to Knudsen diffusion.