Fact-checked by Grok 2 weeks ago

Membrane

A membrane is a thin layer of semi-permeable material that separates substances in a fluid stream by allowing selective based on , charge, or other properties, typically under an applied driving force such as or concentration gradient. In , these barriers are engineered from materials like polymers, ceramics, or composites to enable efficient separation processes. They are widely used in applications including , , gas processing, and industrial separations, offering advantages in and over traditional methods.

Fundamentals

Definition and Basic Principles

A membrane is a thin, selective barrier that permits the of certain across it from one to another while retaining others, based on differences in size, charge, or . This selectivity enables separation processes by exploiting physical or chemical properties of the components in a , with the membrane acting as a semipermeable that responds to applied driving forces. The historical development of synthetic membranes accelerated in the mid-20th century, particularly for applications. In 1959–1960, Sidney Loeb and Srinivasa Sourirajan pioneered asymmetric membranes at UCLA, featuring a thin, dense skin layer over a porous substructure that enhanced flux and salt rejection in . This breakthrough, building on earlier work with films in the 1950s, marked the first viable synthetic membranes for large-scale and laid the foundation for modern . Fundamental principles of membrane operation rely on gradients in across the barrier, which drive the selective of . These gradients arise from pressure differences in pressure-driven processes (e.g., ), concentration differences in diffusion-based separations, or chemical potential disparities in osmotic processes. For flow through porous membranes, the volumetric flux J follows , expressed as J = \frac{\Delta P}{\mu} \cdot \frac{k}{L}, where \Delta P is the pressure drop across the membrane, \mu is the fluid viscosity, k is the permeability coefficient of the porous structure, and L is the membrane thickness. Membranes are categorized by pore size, which determines the separation mechanism. Microporous membranes have pore diameters ranging from 0.1 to 10 μm, enabling sieving of larger particles. Mesoporous membranes feature pores from 2 to 50 nm, suitable for finer separations like macromolecules. Non-porous membranes, lacking discrete pores, rely on a solution-diffusion mechanism where species dissolve into and diffuse through the dense polymer matrix.

Material Composition and Fabrication

Membranes are primarily constructed from polymers, ceramics, and composite materials, each selected for their specific permeability, selectivity, and durability in applications. Polymeric membranes commonly utilize materials such as for (RO) due to its hydrophilicity and chlorine resistance, as the active layer in thin-film composites for RO and nanofiltration (NF) offering high solute rejection rates exceeding 90%, and for (UF) providing robust across a pH range of 1–13. Ceramics, including alumina and zirconia, are favored for high-temperature operations and harsh environments, with alumina exhibiting thermal stability up to 1500°C and zirconia offering enhanced toughness. Composite membranes, such as thin-film composites, integrate a thin polymeric selective layer (e.g., ) on a porous support (e.g., ), combining the selectivity of polymers with the mechanical support of ceramics or additional layers to improve overall performance. Fabrication techniques for these membranes vary by material type to achieve desired asymmetric or symmetric structures. For polymeric membranes, phase inversion is widely employed, including nonsolvent-induced (wet method) where a dope is cast and immersed in a non-solvent bath to form a dense layer over a porous substructure, and dry phase inversion involving prior to immersion. Process parameters like time, exceeding 120 seconds, influence size by densifying the skin layer and reducing microvoids. Interfacial fabricates thin-film composites by reacting aqueous amines (e.g., m-phenylenediamine) with acyl chlorides on a support, yielding ultrathin selective layers (50–200 nm). membranes are produced via , where inorganic powders (e.g., alumina) are compacted and heated to 1000–1500°C to fuse particles into a porous network, or track-etching for polymers like , involving ion bombardment to create latent tracks followed by chemical for uniform nanopores. Surface modifications enhance membrane properties, particularly to mitigate through increased hydrophilicity. Grafting () onto polymeric surfaces via or UV initiation reduces protein adsorption and by lowering the water contact angle and improving wettability. These modifications maintain key mechanical properties, such as tensile exceeding 1 GPa for polysulfone-based polymers, ensuring structural integrity under pressure. Ceramics demonstrate superior thermal stability, with materials like alumina stable up to 1500°C, allowing operation up to 350–500°C without degradation in high-temperature filtrations, depending on the module design. Post-2010 advancements have incorporated nanomaterials like into polymeric matrices to boost selectivity and permeability, with GO layers enabling precise molecular sieving in NF and RO membranes. Zeolite integration in mixed-matrix composites enhances ion rejection by providing uniform nanopores, as seen in zeolite-embedded membranes achieving over 95% salt removal in . In 2025, researchers at developed electrically conductive membranes, improving salt separation and reducing energy consumption in processes. These developments prioritize durability and efficiency while addressing traditional limitations in flux and fouling resistance.

Classification of Membrane Processes

Microfiltration

is a pressure-driven membrane that utilizes porous membranes to remove suspended particles from liquids through a primarily sieving . The membranes typically feature sizes ranging from 0.1 to 10 μm, which effectively retain larger contaminants such as , colloids, and while allowing smaller particles and dissolved substances to pass through as permeate. This size-exclusion process dominates the separation, distinguishing it from other membrane techniques that may involve additional interactions like adsorption or charge effects. The operation of occurs at low transmembrane pressures (), typically between 0.1 and 2 , making it energy-efficient and suitable for applications where is undesirable. Flux rates, defined as the volume of permeate per membrane area per hour, commonly range from 100 to 1000 L/m²h, depending on factors such as feed composition, membrane material, and cross-flow velocity; these rates support high throughput without excessive input from pumping, as osmotic pressures are negligible due to the large sizes. Key applications of include the clarification of beverages such as wine and , where it removes , haze-forming particles, and microorganisms to achieve product stability without , and as a pretreatment step for more selective processes like or in systems. Historically, gained commercial traction in the for large-scale , enabling cross-flow configurations that reduced and supported industrial adoption for solids removal. Despite its advantages, has limitations, including its inability to retain dissolved solutes, small ions, or viruses smaller than the , necessitating downstream treatments for comprehensive purification. is primarily associated with pumping to maintain cross-flow and , rather than overcoming osmotic barriers, though can increase operational costs if not managed. In comparison to , targets coarser separations above 0.1 μm, providing less refined rejection of macromolecules.

Ultrafiltration

Ultrafiltration (UF) is a pressure-driven that effectively concentrates and purifies solutions by retaining macromolecules, fine colloids, proteins, viruses, and emulsions while allowing smaller solutes and to pass through as permeate. The process operates through a combination of sieving, which excludes particles larger than the membrane pores, and , which influences the transport of smaller across the membrane under concentration gradients. UF membranes typically feature pore sizes ranging from 0.001 to 0.1 μm (1 to 100 ), enabling selective retention based on molecular dimensions. The selectivity of UF membranes is characterized by the (MWCO), which represents the molecular weight at which a solute is retained by approximately 90%, typically spanning 1 to 1000 for standard applications. This allows for the separation of biomolecules such as proteins (e.g., globular proteins around 10-100 ) and viruses (20-300 nm in size), while permitting the passage of salts, sugars, and low-molecular-weight compounds. The rejection R, a key metric of membrane performance, is calculated as R = 1 - \frac{C_p}{C_f}, where C_p is the solute concentration in the permeate and C_f is the concentration in the feed; values approaching 1 indicate high retention. UF systems commonly operate under transmembrane pressure (TMP) ranging from 1 to 10 bar, with permeate flux typically between 50 and 200 L/m²h, depending on feed properties, membrane material, and cross-flow velocity. To enhance purification, UF is frequently employed in diafiltration mode, where buffer is added to the retentate to dilute and remove permeable solutes, achieving efficient solute exchange without significant loss of retained species. In the dairy industry, UF has been pivotal for separation since the 1970s, enabling the of whey protein concentrates by retaining proteins while permeating and minerals, thus transforming cheese from a waste stream into a valuable ingredient. In biotechnology, UF facilitates cell harvesting by concentrating microbial or mammalian cells from broths, supporting in and recombinant protein production.

Nanofiltration

is a pressure-driven that serves as an intermediate method between and , characterized by pore sizes typically ranging from 0.5 to 2 . This nanoscale pore structure enables the selective retention of multivalent ions, such as Ca²⁺ and SO₄²⁻, as well as organic molecules with molecular weights exceeding 200 , while allowing partial passage of monovalent ions and smaller solutes. The operates through a involving steric hindrance, which excludes solutes larger than the effective pore radius via size-based sieving, and Donnan exclusion, an electrostatic repulsion arising from the charged nature of the membrane surface interacting with ions of opposite charge. Typical operating conditions for NF include transmembrane pressures (TMP) of 5 to 20 bar, which are lower than those required for , resulting in reduced energy consumption while achieving fluxes of 20 to 50 L/m²h under standard conditions. Rejection rates exemplify this selectivity: NF membranes typically achieve over 90% rejection of divalent salts like MgSO₄, compared to less than 50% for monovalent salts such as NaCl, making it suitable for targeted separation without complete . NF technology emerged in the as a practical solution for partial demineralization, with early developments including composite membranes from interfacial , such as piperazine-based polyamides introduced by FilmTec for use. Key applications include by removing hardness-causing divalent cations and anions, as well as removal from effluents, where the process recovers valuable colorants while treating . Most commercial NF membranes are thin-film composite structures with a negatively charged polyamide active layer, which enhances Donnan exclusion for cations but introduces a between water permeability and solute selectivity—thinner or more open structures boost at the expense of rejection efficiency. This balance is critical in design, as optimizing the and uniformity allows for tailored performance in specific applications.

Reverse Osmosis

Reverse osmosis () is a pressure-driven membrane that produces high-purity by separating dissolved solutes, particularly ions and salts, from a feed using a semi-permeable, non-porous membrane. Unlike porous methods, RO operates via the , where molecules dissolve into the membrane material, diffuse across it under an applied , and desorb on the permeate side, while solutes are largely rejected due to their lower and in the polymer matrix. This results in rejection rates exceeding 99% for monovalent ions like sodium and in typical or applications, with no reliance on discrete pores for separation. The process requires a transmembrane (TMP) of 10-80 to overcome the of the feed solution, which drives natural toward solute concentration and must be reversed for permeation. (π) is calculated using the van't Hoff : \pi = iCRT where i is the van't Hoff factor accounting for , C is the solute , R is the (8.314 J/mol·K), and T is the absolute in ; for with approximately 0.6 M total , π is around 25-30 at 25°C. Typical water flux through commercial RO membranes ranges from 10-50 L/m²·h, depending on feed (lower for at 15-25 L/m²·h and higher for up to 40 L/m²·h), membrane type, and operating conditions like and . The first large-scale RO desalination plant was commissioned in 1965 in , treating brackish groundwater at a capacity of about 19 m³/day (5,000 gallons per day) using membranes, marking the transition from laboratory demonstrations to practical water production. A key operational challenge in RO is , where rejected solutes accumulate at the membrane surface, elevating local and reducing effective driving force, which can lower by 20-50% and increase salt passage if not mitigated through high cross-flow velocities. RO membranes are also susceptible to , which exacerbates performance decline over time. Energy efficiency in modern RO systems has improved significantly through pressure exchangers, devices that transfer hydraulic from the high-pressure brine reject stream to the incoming feed, recovering up to 98% of the and reducing specific energy consumption to approximately 3 kWh/m³ for at 50% . This is a substantial decrease from early systems exceeding 10 kWh/m³, enabling RO to become the dominant technology for large-scale plants worldwide.

Forward Osmosis and Emerging Variants

Forward osmosis (FO) is an osmotically driven membrane process that utilizes a semi-permeable membrane to transport water from a feed to a draw across an osmotic pressure gradient. The driving force arises from the higher of the draw solution, such as a 2 M NaCl solution, which pulls water through the membrane without requiring external hydraulic pressure. The water flux in FO is described by the equation J_w = A (\Delta \pi - \Delta P) where J_w is the water flux, A is the water permeability coefficient of the membrane, \Delta \pi is the osmotic pressure difference across the membrane, and \Delta P is the applied hydraulic pressure difference (typically negligible in FO). This process offers advantages over pressure-driven methods, including reduced fouling propensity due to lower shear forces and the absence of high operating pressures, enabling the use of thinner or more delicate membranes. A key limitation in FO is internal concentration polarization (ICP), which occurs within the porous support layer of the membrane and dilutes the effective osmotic driving force by up to 50%, thereby reducing overall flux. Commercialization of FO has accelerated since the 2000s, with applications in zero-liquid discharge systems for wastewater treatment and resource recovery, where it concentrates effluents while producing high-quality permeate. However, a major challenge remains the recovery and regeneration of the draw solution, often achieved through integration with reverse osmosis to separate water from the diluted draw, which adds complexity and energy costs to the process. Compared to reverse osmosis, FO can achieve energy savings of 20-50% in hybrid configurations for desalination by leveraging natural osmotic gradients. Emerging variants of non-pressure-driven membrane processes include (MD), a thermally driven separation that relies on a gradient across a hydrophobic microporous membrane to transport from a heated feed. In MD, the membrane (typically made from materials like or ) prevents liquid penetration while allowing vapor passage, operating effectively at moderate temperatures of 40-80°C using low-grade heat sources such as or . This makes MD suitable for and concentration of heat-sensitive solutions, with high solute rejection rates exceeding 99% for non-volatiles. Another variant is , which separates organic compounds from aqueous mixtures using a non-porous selective membrane based on differences in and , driven by a partial pressure gradient on the permeate side. For organic recovery, hydrophobic membranes preferentially permeate volatile organics like or , with the driving force enhanced by applying or using a sweep gas (e.g., ) to remove the permeate vapor and maintain low . excels in treating industrial wastewater containing low concentrations of organics, offering energy-efficient separation for applications such as removal and purification, with selectivities often surpassing traditional .

Membrane Configurations and Modules

Plate-and-Frame Modules

Plate-and-frame modules represent one of the earliest and simplest configurations for housing flat-sheet membranes in separation processes. These modules feature flat membrane sheets stacked between rigid support plates, forming a compact akin to a filter press, with spacers inserted between membranes to create thin feed channels and permeate collection pathways. The channel thickness typically ranges from 0.5 to 2 mm, facilitating cross-flow or dead-end operation while enabling straightforward access for membrane installation and removal. Developed in the early , plate-and-frame modules were among the first commercial membrane systems, drawing from conventional filter press designs to enable initial large-scale applications in processes like and . Early implementations often utilized frames for their chemical resistance and durability, supporting the stacking of multiple membrane leaves in parallel arrangements. These modules gained prominence in and pilot-scale testing due to their and adaptability for various membrane types. Key advantages of plate-and-frame modules include their low cost for small-scale operations, full accessibility for thorough cleaning and membrane replacement, and robustness in handling viscous or high-solids feeds, which reduces compared to more compact designs. The packing density, typically around 100-400 m²/m³, supports moderate throughput in applications such as and , particularly in and . However, these modules suffer from disadvantages like high hold-up volume, which increases operational costs, and susceptibility to channeling—uneven flow distribution that can compromise . The need for multiple seals also contributes to higher drops and challenges, limiting their use in large-scale, high-pressure systems. Despite these limitations, plate-and-frame configurations remain suitable for processes requiring frequent and where is prioritized over high density.

Spiral-Wound Modules

Spiral-wound modules consist of flat-sheet membranes formed into open-ended envelopes that are separated by mesh-like feed spacers and wound spirally around a central perforated permeate collection , creating a compact cylindrical . The feed spacers, typically made of netting, have thicknesses ranging from 0.2 to 1 mm to maintain height and promote turbulent . Standard industrial elements feature an 8-inch (20.3 cm) diameter and are housed in pressure vessels up to 12 feet (3.66 m) long, accommodating multiple elements with a total effective membrane area reaching up to 400 m² per module. This configuration allows for efficient permeate drainage through the central while the feed flows axially along the spiral . These modules offer high packing densities of 300 to 1000 m²/m³, enabling large membrane areas in a relatively small volume, which makes them particularly cost-effective for reverse osmosis (RO) and nanofiltration (NF) processes. They have become the standard configuration in desalination plants due to their scalability and economic advantages in handling high-pressure operations. However, cleaning is challenging as the spiral design does not support backwashing, often requiring chemical circulation or replacement of elements. Additionally, the feed spacers can induce fouling by creating stagnation zones and promoting bio-growth, while pressure drop (ΔP) across the module increases with feed velocity (v) and decreases with higher spacer voidage, potentially limiting throughput. Spiral-wound modules were developed in the , and later commercialized through FilmTec Corporation, which introduced advanced thin-film composite elements. To mitigate spacer-induced shadow effects—where filaments reduce effective membrane area exposure—specialized spacers with optimized geometry have been employed in high-salinity applications to enhance uniformity. These modules are typically operated in cross-flow mode to minimize .

Hollow-Fiber and Tubular Modules

Hollow-fiber membrane modules consist of thousands of thin, cylindrical fibers, each with an outer diameter typically ranging from 0.1 to 1 mm and lengths of 0.5 to 2 m, bundled together within a . These fibers are potted at their ends using to create a fluid-tight seal, allowing the feed stream to flow either inside the fiber lumens (inside-out configuration) or outside the fibers in the shell side (outside-in configuration). This design enables high packing densities exceeding 1000 m²/m³, maximizing membrane surface area within a compact volume. The first commercial hollow-fiber modules were developed by in the 1960s, marking a pivotal advancement in for and separation processes. Tubular membrane modules, in contrast, feature larger-diameter tubes with inner diameters of 5 to 25 mm, housed within a and often self-supporting without additional internal frameworks. These modules are particularly suited for processing viscous feeds or those with high solids content, such as in applications for , where the robust tube structure facilitates high cross-flow velocities to minimize accumulation. Both configurations offer significant advantages, including high effective surface area relative to module volume and reduced material requirements compared to flat-sheet alternatives, enabling efficient large-scale operations. However, a key disadvantage is the vulnerability of hollow fibers to mechanical failure, such as breakage, at very high pressures (e.g., exceeding ), which may be encountered in some high-pressure applications. Tubular modules mitigate some of these issues through their sturdier construction but at the cost of lower packing densities.

Operational Principles

Flux, Pressure, and Permeability

In membrane processes, represents the rate of permeate transport across the membrane surface, quantified as the volume flow per unit membrane area per unit time. It is typically expressed by the equation J = \frac{V}{A t}, where V is the permeate , A is the effective membrane area, and t is time, with common units of liters per square meter per hour (L/m²h). is primarily driven by the applied difference and opposed by the total resistance to flow, encompassing intrinsic membrane properties and process conditions such as feed and concentration. Trans (TMP) serves as the primary driving force for in pressure-driven membrane operations, calculated as the average pressure difference across the membrane: \text{[TMP](/page/TMP)} = \frac{P_{\text{feed}} + P_{\text{retentate}}}{2} - P_{\text{permeate}}, where P_{\text{feed}}, P_{\text{retentate}}, and P_{\text{permeate}} are the pressures in the feed, retentate, and permeate streams, respectively. In (RO), TMP must exceed the difference (\Delta \pi) between feed and permeate to achieve net flux, as osmotic pressure arises from solute concentration gradients opposing solvent movement. Membrane permeability quantifies the intrinsic ease of transport for specific species through the membrane material, independent of external fouling or effects. In systems, permeability coefficient A and salt permeability coefficient B are key parameters in the solution-diffusion model, where flux is given by J_w = A (\Delta P - \Delta \pi) and salt flux by J_s = B \Delta C, with \Delta P as the hydraulic pressure difference, \Delta \pi as the difference, and \Delta C as the solute concentration difference across the membrane. These s are determined experimentally using clean membrane tests, such as pure permeability assessments under varying TMP to isolate A via of flux against net driving pressure. To evaluate overall performance and isolate contributions to , the -in-series model decomposes total flux into additive components via : J = \frac{\Delta P}{\mu (R_m + \sum R_i)}, where R_m is the intrinsic membrane , R_i represents additional (e.g., from or ), and \mu is the of the permeate. The intrinsic R_m is determined from low-pressure pure tests by plotting flux J versus \frac{\Delta P}{\mu}, yielding a linear relationship with slope \frac{1}{R_m}. This approach enables predictions, including in RO applications where such measurements guide module scaling by linking permeability to required TMP for target fluxes.

Dead-End and Cross-Flow Filtration Modes

In dead-end filtration, the feed stream is directed perpendicular to the membrane surface, forcing all liquid through the pores while retained particles accumulate directly on the membrane, forming a layer that progressively increases resistance to flow. This mode operates in a batch-like manner, with no separate retentate stream, making it suitable for applications requiring high permeate recovery, such as laboratory-scale processing or polishing of feeds with low concentrations (typically below 1%). The setup is , involving minimal and no recirculation, but it necessitates frequent or cake removal due to rapid from cake buildup. Cross-flow filtration, also known as tangential flow filtration, contrasts by directing the feed parallel to the membrane surface at velocities typically ranging from 1 to 5 m/s, generating forces that minimize cake layer formation and deposition. A recirculation loop returns the retentate to the feed tank, enabling continuous operation and handling of higher solids loads, though it requires additional pumping that accounts for a significant portion of the system's use. This mode was developed in the as an alternative to dead-end to mitigate in industrial-scale processes, such as and bioprocessing. Performance differences between the modes hinge on operational trade-offs: dead-end filtration excels in scenarios demanding near-complete (up to 95% permeate) from dilute feeds but limits throughput due to , while cross-flow supports sustained, continuous processing with lower rates but reduced maintenance intervals. The transition between viable operation and in both modes is governed by the critical concept, defined as the maximum permeate below which no time-dependent flux decline occurs, allowing to be avoided by starting at sub-critical conditions. In dead-end setups, flow diagrams illustrate perpendicular feed arrows terminating at the membrane with accumulation, whereas cross-flow diagrams show parallel feed paths alongside the surface, with permeate extraction perpendicularly and retentate looping back.

Fouling Phenomena

Types and Mechanisms of Fouling

Membrane fouling in and systems is primarily categorized into four main types: particulate fouling, organic fouling, , and inorganic . Each type arises from specific interactions between the feed solution components and the membrane surface, leading to reduced permeability and . Particulate fouling occurs through the accumulation of , such as colloids and fine particles, on the membrane surface, forming a cake layer that increases hydraulic resistance. This process typically follows a cake mechanism, where particles deposit externally without penetrating the pores, as observed in and applications. Organic fouling involves the adsorption of natural organic matter, including and , onto the membrane, often driven by hydrophobic interactions and van der Waals forces. These foulants can alter the membrane's surface properties, promoting further deposition in feeds with high . results from the attachment and proliferation of microorganisms, leading to formation composed of cells and extracellular polymeric substances that severely impede flow. This type has gained significant attention since the early 2000s, particularly in membrane bioreactors (MBRs), where microbial communities thrive under low shear conditions. Inorganic scaling, or mineral precipitation, arises from the supersaturation of sparingly soluble salts like (CaCO₃) near the membrane surface due to , forming crystalline deposits that constrict pores or coat the surface. The underlying mechanisms of fouling are commonly described using Hermia's blocking models, which characterize flux decline under constant pressure as \frac{d^2 t}{d V^2} = k \left( \frac{d V}{d t} \right)^n, where V is the filtrate volume, t is time, k is a constant, and n determines the mechanism: n = 2 for complete pore blocking (foulants seal pore entrances without penetration), n = 1 for intermediate blocking (foulants deposit sequentially on the surface, partially obstructing pores or landing on prior deposits), n = 1.5 for standard blocking or pore constriction (foulants deposit inside pores, uniformly reducing radius), and n = 0 for cake filtration (foulants form an external layer). These models provide a framework for identifying dominant fouling behaviors based on experimental flux data. Several factors influence the onset and progression of fouling. Feed composition plays a critical role; for instance, high (TDS) and divalent cations like Ca²⁺ accelerate by promoting , while elevated organic content exacerbates adsorption-based . Hydrodynamic conditions, such as low cross-flow velocity, reduce shear stress and favor and cake layer buildup by allowing foulants to settle more readily. propensity is often quantified using the (SDI), which measures the rate of particulate plugging in a standardized test; values below 3 are recommended for feeds to minimize risks. Fouling can be classified by time scales into reversible and irreversible forms. Reversible fouling involves loosely attached deposits, such as physical cake layers from particulates, which may partially detach under flow changes, whereas irreversible fouling features strong chemical bonds or embedded biofilms that persist over extended operations.

Fouling Prediction and Modeling

Fouling prediction in membrane systems relies on mathematical models that quantify the accumulation of foulants and its impact on flux decline. The resistance-in-series (RIS) model is a foundational approach, treating total filtration resistance as the additive sum of intrinsic membrane resistance (R_m), cake layer resistance (R_c), and irreversible fouling resistance (R_f), expressed as R_{total} = R_m + R_c + R_f. This phenomenological model, derived from Darcy's law, allows estimation of flux J via J = \frac{\Delta P}{\mu R_{total}}, where \Delta P is transmembrane pressure and \mu is fluid viscosity, and has been widely applied to predict fouling in ultrafiltration and microfiltration processes. For organic foulants, the gel polarization model extends this by describing a gel layer formation at the membrane surface when solute concentration reaches a critical gel-point, leading to a limiting flux independent of further pressure increase; this model, originally developed for protein solutions, highlights diffusive back-transport balancing convective deposition under cross-flow conditions. Computational fluid dynamics (CFD) simulations complement these by modeling shear stress distribution along membrane surfaces, revealing spatial variations in fouling propensity due to flow hydrodynamics in modules like spiral-wound elements. Empirical indices provide practical tools for fouling assessment based on standardized lab tests. The modified fouling index (MFI), specifically MFI-0.45 for silt-like , is determined by dead-end through a 0.45 \mum membrane at constant , plotting filtration time per volume (t/V) against time (t); the of the linear portion yields MFI values, with higher indices indicating greater potential (e.g., MFI > 3 suggests rapid fouling in feeds). For inorganic , the Langelier saturation index (LSI) predicts precipitation, calculated as LSI = [pH](/page/PH) - pH_s, where pH_s is the saturation pH derived from water chemistry factors like , calcium concentration, , and ; LSI > 0 signals and risk, commonly used to forecast inorganic in membranes. Advanced predictive methods incorporate data-driven techniques for complex, multifoulant scenarios. Since 2015, algorithms, such as artificial neural networks and support vector machines, have enabled fouling prediction by on operational data like , , and parameters, achieving up to 95% accuracy in forecasting transmembrane rise in membrane bioreactors. The unified membrane fouling index (UMFI) integrates multiple foulants by normalizing resistance development over filtration cycles in bench-scale tests, distinguishing reversible and irreversible components to provide a comprehensive fouling propensity score for low-pressure membranes treating natural waters. Model validation typically involves lab-scale experiments, where feed solutions are circulated tangentially to mimic module conditions, allowing parameter fitting (e.g., cake in RIS) by comparing predicted versus measured decline. These tests confirm model applicability under controlled rates but reveal limitations in dynamic, full-scale operations, such as unsteady or multifoulant interactions, where predictions deviate by 20-50% due to unmodeled dynamics or temperature fluctuations.

Fouling Control Strategies

Cleaning Methods

Cleaning methods for fouled membranes aim to restore permeability and flux by removing accumulated deposits, targeting reversible fouling such as loose particulates, colloids, and biofilms while minimizing damage to the membrane material. These techniques are broadly categorized into physical, chemical, and biological approaches, often applied in sequence during clean-in-place (CIP) operations to achieve optimal recovery without disassembly. Physical methods provide immediate shear-based removal suitable for surface-level foulants, while chemical and biological methods address more adherent layers like inorganic scales or organic matrices. Physical cleaning techniques rely on hydraulic or mechanical forces to dislodge without chemical agents, making them ideal for frequent maintenance of reversible . Backflushing involves reversing the permeate at pressures of 1-2 to create a negative trans-membrane that expels trapped particles from interiors. For membranes, sponge cleaning circulates slightly oversized sponge balls through the modules at controlled velocities to mechanically scour the inner surfaces, effectively removing soft deposits like biofilms or organics. In submerged membrane systems, air scouring introduces coarse bubbles via diffusers to generate turbulent along the membrane exterior, enhancing of cake layers in membrane bioreactors. These methods typically restore 80-95% of initial for reversible fouling but are less effective against irreversible scales. Chemical cleaning employs solutions to dissolve or degrade specific foulants, often following physical methods for deeper penetration. Acidic cleaners, such as (HCl) at 2-3, target inorganic scales and metal oxides by solubilizing precipitates like or silica. Alkaline solutions, including (NaOH) at 11-12, hydrolyze organic foulants and proteins, breaking down matrices. CIP cycles with these agents last 30-60 minutes, involving circulation at low velocities (0.3-1 m/s) followed by rinsing to prevent residue buildup. Biological and emerging techniques offer targeted degradation for complex biofilms, complementing traditional methods. Enzyme treatments, particularly proteases like , hydrolyze extracellular polymeric substances in biofilms, disrupting microbial adhesion and facilitating removal at neutral and ambient temperatures. applies waves at 20-40 kHz to induce bubbles that generate micro-jets, eroding foulant layers without chemicals, though limits its routine use. Electrical fields, an emerging approach, apply pulsed currents to alter foulant and promote desorption, showing promise in hybrid systems. Additionally, and models are being integrated for predictive monitoring and strategies (as of 2025). Cleaning protocols are triggered by operational indicators, such as a flux decline exceeding 10-15% from baseline or a 5-10% rise in salt passage, to balance performance with membrane longevity. Frequency varies by application—daily for high-fouling feeds like —but aims for 80-95% in reversible cases through sequential physical-chemical cycles. Post-cleaning involves normalized and to ensure efficacy before resuming operation.

Design and Operational Mitigations

Design strategies for minimizing begin with feed pretreatment, where (MF) serves as a guard process ahead of (RO) systems to remove particulates, colloids, and larger foulants that could otherwise deposit on the RO membrane surface. This pretreatment significantly reduces the fouling potential, as measured by silt density index (SDI) and reductions, in applications, extending membrane life and operational intervals. Spacer optimization within membrane modules incorporates promoters, such as net-type or ladder-like structures, to enhance dynamics and forces at the membrane surface, thereby disrupting boundary layers and limiting foulant accumulation. These promoters can increase coefficients by 20-50% while balancing , as demonstrated in spiral-wound modules. Surface modifications, including zwitterionic coatings, further mitigate by creating layers that repel proteins and organics; for instance, such coatings on polyethersulfone membranes have reduced rates by approximately 50% during protein filtration tests. Operational mitigations complement design by adjusting process parameters to maintain low fouling propensity. Pulsed flow regimes, alternating between high and low velocities, disrupt and shear away nascent foulant layers, improving flux stability by 15-30% in systems compared to steady flow. Employing high cross-flow velocities, typically exceeding 4 m/s in configurations, enhances scouring effects to prevent cake layer formation, particularly for viscous feeds like dairy wastewater. adjustment in the feed stream, targeting a range of 6-7, controls silica and , reducing scaling risks in RO processes by inhibiting amorphous silica deposition on the membrane. Continuous online monitoring of transmembrane (TMP) rise enables early detection of fouling onset, allowing operators to intervene before irreversible layers form, as TMP increases of 0.1-0.5 per day signal the need for parameter tweaks in submerged membrane bioreactors. Antiscalants, such as polyacrylate-based polymers, are dosed at 2-5 ppm to achieve threshold inhibition exceeding 95% for common scales like and in RO feeds. These additives distort and maintain supersaturated solutions without , with polyacrylates showing superior performance in high-silica waters up to 200 ppm. Innovative approaches include hybrid () systems integrated with pressure-driven membranes, which operate at lower (often <1 bar) to reduce compaction and fouling inducement from high pressures, achieving up to 90% rejection of organics while minimizing energy use. Overall, these preventive measures can save 20-30% on cleaning-related operational expenditures (OPEX) by extending run times and reducing downtime, serving as a primary line of defense with cleaning reserved as a backup.

Sustainability in Membrane Technology

Environmental Impacts and Waste Management

Membrane processes, particularly reverse osmosis (RO), exhibit significant environmental impacts due to their high energy demands. Seawater RO desalination typically requires 2.5-4 kWh per cubic meter of produced water as of 2024, accounting for approximately 0.4% of global electricity consumption. This energy intensity arises from the high pressures needed to overcome osmotic barriers, contributing to greenhouse gas emissions in regions reliant on fossil fuel-based power. Additionally, the discharge of hypersaline brine, with total dissolved solids (TDS) concentrations of 40-70 g/L—roughly double that of ambient seawater—poses risks to marine ecosystems by altering local salinity gradients, reducing dissolved oxygen levels, and stressing benthic organisms and fisheries. Waste management challenges in membrane operations stem from the finite lifespan of polymeric elements, which typically endure 3-5 years under operational stresses like fouling and chemical cleaning. With the desalination industry expanding, by 2025, over 2 million end-of-life RO modules are expected to be generated annually worldwide. Discarded membranes contribute to solid waste at desalination facilities, exacerbating landfill burdens and plastic pollution. Degradation of these polyamide-based materials can release microplastic fragments into the environment, particularly during improper disposal or incineration, potentially entering marine food webs and compounding global plastic pollution. To mitigate these issues, zero-liquid discharge (ZLD) systems integrate membrane processes with thermal treatments like evaporation and crystallization, recovering nearly all water while minimizing effluent volumes. Brine mining has emerged as a valorization strategy since the 2010s, extracting valuable minerals such as lithium from concentrated streams to offset environmental costs and support the green energy transition. Regulatory frameworks, including the , enforce limits on brine discharge to prevent ecological harm, requiring controls on salinity increases in receiving waters through dilution or treatment protocols. Approaches like recycling end-of-life membranes offer supplementary pathways for waste reduction, though detailed methods are addressed elsewhere.

Recycling and Reuse Techniques

Recycling and reuse techniques for end-of-life membranes, particularly reverse osmosis (RO) types, focus on recovering materials or repurposing modules to extend their utility and reduce waste. Chemical methods, such as oxidative treatment with sodium hypochlorite solutions, effectively strip the degraded polyamide active layer from thin-film composite (TFC) RO membranes, exposing the underlying polysulfone support layer for reuse as ultrafiltration (UF) or microfiltration (MF) membranes. This process involves immersing spiral-wound elements in a 0.5-2% sodium hypochlorite solution at ambient temperature for several hours, followed by rinsing and a secondary cleaning with 0.2% citric acid (pH 2.5) for 40 minutes to remove residual inorganic foulants like calcium carbonate. The resulting membranes exhibit significantly increased permeability—up to 100-200 L/m²·h·bar compared to original RO values of 1-5 L/m²·h·bar—while retaining over 90% rejection for larger particles, enabling their application in lower-pressure filtration for wastewater or brackish water treatment. Alternative chemical approaches include sonication, which degrades the polyamide layer through ultrasonic cavitation, converting RO membranes to MF in as little as 15 minutes. When combined with a 5000 ppm potassium permanganate (KMnO₄) solution, this method achieves a 70-fold permeability increase (to 172.6 L/m²·h·bar) and reduces salt rejection to below 2%, with humic acid separation efficiency around 82%, making it suitable for organic-laden feeds. For material recovery, solvent extraction using dimethylformamide (DMF) delaminates TFC RO membranes, dissolving the polysulfone middle layer while leaving the polyester nonwoven and polyamide layers intact for separate collection. The dissolved polysulfone is precipitated with water, filtered, and purified via chlorine treatment, yielding high-purity polymers (verified by FT-IR, DSC, and TGA) comparable to virgin materials, with molecular weights matching commercial grades. Thermal techniques like pyrolysis offer a route to energy and material recovery from non-reusable membranes. Conducted at 500-600°C in an inert atmosphere, pyrolysis decomposes polymeric components into syngas, pyrolysis oil, and char, with typical yields of 28 wt% oil, 17 wt% non-condensable gases, and 22 wt% char from TFC RO membranes. The char can serve as a carbon precursor for applications like activated carbon or dots, while the oil and gases provide recoverable energy, achieving up to 86% thermal energy recovery and 48.5% mass reduction. For RO-specific processing, disassembly of spiral-wound elements involves manual or mechanical removal of the fiberglass outer casing and glue, followed by layer separation to access the support fabric, which post-2015 pilot studies have recovered at efficiencies exceeding 80% for repurposing as UF supports after active layer stripping. Grinding the remaining components, such as the polyester support or casing, produces filler aggregates for concrete, providing a low-tech recycling option without chemical inputs. Refurbished RO modules are commonly downcycled to lower-grade applications, such as converting spent seawater RO to brackish water NF or UF for industrial effluent treatment, where flux rates of 50-100 L/m²·h and 70-90% rejection for divalent ions suffice. Global recycling rates for end-of-life RO membranes remain low, primarily due to landfilling practices, though pilots indicate potential for improved recovery with sorting and modular disassembly. Economic viability is enhanced in these pilots, with processing costs of €25-75 per module for passive oxidation systems versus €400-800 for new modules, though scaling challenges persist. Key challenges include thorough removal of contaminants like heavy metals (e.g., iron, lead) embedded during operation, which requires additional leaching steps to prevent leaching in reuse applications, and economic hurdles where processing costs ($50-100 per ton) compete with cheap landfilling but lag behind new membrane prices (~$500 per ton equivalent). Despite these, high-impact pilots since 2015, such as those converting brackish water RO to UF, demonstrate 80% support layer recovery and lifecycle cost reductions of 43-52% compared to active chemical methods. Overall, these techniques promote circularity in membrane technology, generalizable beyond RO to other polymeric types like nanofiltration.

Applications

Water Purification and Desalination

Membrane technologies play a pivotal role in water purification and desalination, enabling the production of potable water from brackish and seawater sources to address global water scarcity. dominates this field, accounting for approximately 70% of global desalination capacity worldwide. As of 2024, over 21,000 desalination plants operate globally, producing more than 140 million cubic meters of fresh water per day. In desalination processes, RO is the primary method, often preceded by hybrid microfiltration (MF)-RO systems for pretreatment to remove particulates, colloids, and microorganisms, thereby minimizing fouling on RO membranes. MF pretreatment achieves silt density index (SDI) values below 3, ensuring stable RO operation with recovery rates of 45-55% and reduced chemical usage compared to conventional filtration. These hybrid configurations enhance overall system reliability, particularly in handling variable seawater quality. A notable case study is the Perth Seawater Desalination Plant in Australia, operational since 2006, which has a capacity of 144 megalitres per day (ML/day) and achieves an energy consumption of 3.5 kilowatt-hours per cubic meter (kWh/m³). The plant employs a two-pass RO system to meet stringent water quality standards, including effective boron removal, as single-pass RO typically rejects only 50-70% of boron, necessitating the second pass at elevated pH for over 90% overall rejection. Fouling remains a challenge in such plants, mitigated through regular cleaning and optimized pretreatment. Economically, seawater RO desalination involves capital costs of $1-2 per cubic meter of daily capacity and operational expenditures (OPEX) of $0.5-1 per cubic meter of produced water, with membrane replacement accounting for about 20% of total costs due to their role in core separation. Recent advances in seawater RO (SWRO) include energy recovery devices (ERDs) such as isobaric systems, which achieve up to 97% efficiency by recycling pressure from the brine stream, reducing specific energy consumption to 3-4 kWh/m³ in modern plants.

Industrial Separations and Gas Processing

Membrane technology plays a pivotal role in industrial separations, particularly for gas processing and non-aqueous liquid streams, offering energy-efficient alternatives to traditional distillation and absorption methods. In gas separation, polymeric membranes exploit differences in gas permeability and selectivity to purify streams, with applications in air separation and natural gas upgrading. For liquid separations, ultrafiltration (UF) and pervaporation enable concentration and purification in food and biofuel production, respectively, by leveraging size exclusion and solution-diffusion mechanisms. These processes have scaled commercially since the late 20th century, driven by advancements in membrane materials and module designs. Polymeric membranes for air separation primarily target O₂ enrichment from N₂, achieving selectivities (α_{O_2/N_2}) up to approximately 20 in advanced formulations like facilitated transport systems incorporating metal ions or polar groups to enhance O₂ affinity. These membranes, often based on poly(ethylene oxide) or polyimides, operate under moderate pressures (4-10 bar) and provide a cost-effective route for oxygen production in medical and industrial uses, surpassing cryogenic distillation in energy use by 50-70%. Commercial adoption began in the 1990s, with hollow-fiber configurations dominating due to high packing density. In natural gas processing, polymeric membranes remove CO₂ from CH₄ to meet pipeline specifications (<2-4% CO₂), with commercialization accelerating post-1980s using cellulose acetate and polyimide materials. Early systems exhibited CO₂ permeance of 1-10 GPU (gas permeation units), balancing permeability and selectivity (α_{CO_2/CH_4} ~20-40) while resisting plasticization from high CO₂ partial pressures. These membranes reduce CO₂ content in sour gas streams, enabling enhanced oil recovery and biogas upgrading, with over 1,000 units installed globally by the 2000s for capacities up to 100 MMSCFD. For liquid separations, ultrafiltration is widely applied in the food industry to clarify and concentrate juices, retaining macromolecules like pectins and proteins while allowing sugars and to permeate, achieving solute recovery rates of 85-95% and producing concentrates with 10-20% total solids for products like apple or orange juice. This process preserves nutritional quality and aroma compounds better than thermal evaporation, with flux rates of 20-50 L/m²·h under cross-flow conditions. In biofuel production, pervaporation using membranes dehydrates - mixtures to >99% purity, exploiting hydrophobic pores (e.g., silicalite-1) for selective removal at low temperatures (60-80°C), reducing energy costs by 40% compared to . LTA membranes, in particular, demonstrate separation factors >100 for /, enabling fuel-grade from broths. A seminal from the involves the Prudhoe Bay natural gas facility in , where early membrane systems reduced CO₂ levels from approximately 12% to 4%, facilitating gas reinjection and transport while minimizing corrosion risks; this deployment by (now part of ) marked one of the first large-scale applications, processing 50 MMSCFD with >90% CO₂ recovery efficiency. In biopharmaceuticals, UF combined with (UF/DF) removes viruses from protein solutions, achieving >4-6 for viruses >20-30 nm using polysulfone membranes (MWCO 30-300 kDa), ensuring sterility in production without harsh chemicals. These applications highlight membranes' robustness in high-value separations. The gas separation constitutes about 20% of the overall membrane industry, valued at roughly USD 1.5 billion in 2024, with polymeric types holding 60-70% share due to cost-effectiveness. Growth has accelerated post-Paris Agreement (2015), particularly in (CCS), where membranes enable 90% CO₂ capture from flue gases at <20% of amine system costs, projecting a market expansion to USD 2.2 billion by 2030 at 7-8% CAGR driven by net-zero targets. Hollow-fiber modules, briefly referencing spiral-wound and plate-frame configurations, dominate installations for their scalability in these sectors.

References

  1. [1]
    Cell Membrane (Plasma Membrane)
    The cell membrane, also called the plasma membrane, is found in all cells and separates the interior of the cell from the outside environment.
  2. [2]
    Cell Membranes - The Cell - NCBI Bookshelf - NIH
    Biological membranes consist of proteins inserted into a lipid bilayer. Integral membrane proteins are embedded in the membrane, usually via α-helical ...
  3. [3]
    Membrane Structure - Molecular Biology of the Cell - NCBI Bookshelf
    The plasma membrane encloses the cell, defines its boundaries, and maintains the essential differences between the cytosol and the extracellular environment.
  4. [4]
    Membranes and Transport - Biological Principles
    The cell membrane is a fundamental and defining feature of cells. It is composed of a phospholipid bilayer, with the hydrophilic phosphate “head” groups facing ...
  5. [5]
    Cell Structure - SEER Training Modules - National Cancer Institute
    The cell membrane is a double layer of phospholipid molecules. Proteins in the cell membrane provide structural support, form channels for passage of materials, ...<|control11|><|separator|>
  6. [6]
    membrane definition
    A flexible layer surrounding a cell, organelle (such as the nucleus), or other bodily structure. The movement of molecules across a membrane is strictly ...
  7. [7]
    [PDF] Biological Membranes - Projects at Harvard
    A defining feature of the cell is a membrane, the cytoplasmic membrane, that surrounds the cell and separates the inside of the cell, the cytoplasm, from other ...
  8. [8]
    [PDF] Section One: Chapter 5: Cell Boundaries, Membranes and Transport
    The cholesterol in the membrane helps to fill any voids in the lipid insulator and presumably cholesterol blocks the openings between phospholipid tails through ...
  9. [9]
    MEMBRANE TRANSPORT – BIOLOGY BASICS
    Membrane transport is defined as passage of a substance across a membrane. The cell membrane is selectively permeable.
  10. [10]
    Membranes – Visual Encyclopedia of Chemical Engineering ...
    Apr 6, 2022 · A membrane is a thin barrier that permits the transport of certain species across it from one fluid to another.
  11. [11]
    [PDF] Membrane Filtration
    A membrane or, more properly, a semipermeable membrane, is a thin layer of material capable of separating substances when a driving force is applied across the ...
  12. [12]
    Separation processes - processdesign
    Feb 23, 2016 · This separations process uses selectively permeable membranes to separate components in a mixture. Typically, one of the components will freely ...<|separator|>
  13. [13]
    [PDF] Reverse Osmosis: Introduction - S. Loeb
    Historical Development​​ In 1959–1960 Loeb and Sourirajan (L–S) found that anisotropic cellulose acetate membranes, those possessing a skin surmounting a porous ...
  14. [14]
    History | CBE - Chemical and Biomolecular Engineering @ UCLA
    At that time, Samuel Yuster and two of his students, Sidney Loeb and Srinivasa Sourirajan, produced a functional synthetic RO membrane from cellulose acetate ...
  15. [15]
    [PDF] Membranes for Water Treatment: Reverse Osmosis and Nanofiltration
    Water treatment processes employ several types of membranes1. They include microfiltration (MF), ultrafiltration (UF), reverse osmosis (RO), and nanofiltration ...
  16. [16]
    Pressure-Driven Membrane Separation Technologies - ITRC
    The types of driving forces include mechanical pressure, concentration or chemical potential, and temperature or electrical potential (Mortazavi 2008). The ...
  17. [17]
    [PDF] Understand the Basics of Membrane Filtration
    The chemical potential gradient, or driving force, in membrane separation processes can arise from a pressure, concentration, or temperature difference.
  18. [18]
    [PDF] Incorporating Darcy's law for pure solvent flow through porous tubes
    Mar 28, 2012 · Pressure-driven fluid flow through a porous tube occurs in such diverse applications as filtration, aeration, sparging, foaming, membrane ...
  19. [19]
    A review of polymeric membranes and processes for potable water ...
    Membranes, particularly polymeric membranes, play a crucial role in the purification of municipal wastewater to potable quality, and are the core part of almost ...
  20. [20]
    Thermal Properties of Technical Ceramics - CoorsTek
    ... thermal resistance of 1500 °C or more. Zirconia Toughened Alumina is a specialized alumina designed for high thermal shock resistance and increased toughness.
  21. [21]
    Ceramic-polymer composite membranes: Synthesis methods and ...
    Feb 1, 2024 · Composite membranes are more cost-effective, physically and chemically stable as compared to pristine ceramic and polymeric membrane ...
  22. [22]
    [PDF] Polymers and Solvents Used in Membrane Fabrication - UKnowledge
    Apr 23, 2021 · This review examines the advances in sustainable membrane development and per- formance. In particular, these advances for polymeric membranes ...<|control11|><|separator|>
  23. [23]
    Review—Track-Etched Nanoporous Polymer Membranes as Sensors
    Jan 21, 2020 · In this review, we look at the background on tracking technology, chemical etching of these tracks for the fabrication of nanopores with varying geometries.<|control11|><|separator|>
  24. [24]
    Surface modification of polypropylene membrane by polyethylene ...
    Sep 1, 2014 · The results clearly indicate that plasma graft-polymerization of PEG is an effective way for antifouling improvement of polypropylene membranes.
  25. [25]
    Polysulfone Mixed-Matrix Membranes Comprising Poly(ethylene ...
    Jul 27, 2021 · The observed increase in tensile modulus and strength is attributed to the favorable interaction between the PSF matrix and the well-dispersed ...
  26. [26]
    Recent Developments of Graphene Oxide-Based Membranes - MDPI
    In this review paper, the latest research progress in GO-based membranes focused on adjusting membrane structure and enhancing their mechanical strength as ...
  27. [27]
    Ultrapermeable 2D-channeled graphene-wrapped zeolite molecular ...
    May 18, 2022 · Recent study of membranes has reported progress in MFI zeolite membranes in which crack-free MFI membranes were prepared using exfoliated ...
  28. [28]
    Microfiltration - an overview | ScienceDirect Topics
    Microfiltration refers to filtration of a liquid suspension through a membrane with pores of approximately 0.1–10 μm in diameter, to retain microorganisms ...Missing: retention | Show results with:retention
  29. [29]
    Innovative Trends in Modified Membranes: A Mini Review of ...
    Microfiltration and Ultrafiltration. In microfiltration (MF) processes, where the pore sizes range between 0.1–10 μm, the primary transport mechanism is size ...
  30. [30]
    Microfiltration - an overview | ScienceDirect Topics
    Microfiltration ... transmembrane pressure (TMP) on both surface of the membrane is low, demanding a TMP of less than 2 bar, but ranging from 0.1 to 2 bar.
  31. [31]
    [PDF] Best Practices for Optimization and Scale-Up of Microfiltration TFF ...
    a noteworthy feature is the low transmembrane pressures. (TMP) (in the range of 1–3 psi) required for the process operation. Secondly, a microfiltration TFF ...Missing: bar m²h
  32. [32]
    BREAKTHROUGH FLUX PERFORMANCE OF 1000+ LMH - Cerafiltec
    It is the first time that a ceramic flat sheet membrane is operating at flux rates of 1,000 LMH and well beyond. The transmembrane pressure is very low at ...Missing: microfiltration 100-1000
  33. [33]
    Fouling and Chemical Cleaning of Microfiltration Membranes - NIH
    Mar 10, 2021 · This review is carried out to enable a better understanding of the membrane cleaning process for an effective membrane separation process.Missing: commercialization | Show results with:commercialization
  34. [34]
    Microfiltration: Enhancing Water Purity and Safety
    Jun 10, 2024 · However, it also has limitations, including its ineffectiveness in filtering out chemicals and certain viruses.Missing: solutes | Show results with:solutes
  35. [35]
    Application of Membrane Filtration to Cold Sterilization of Drinks and ...
    Mar 18, 2023 · Ultrafiltration membrane has a pore size ranging from 0.001 to < 0.1 µm (McCarron et al., 2016) while microfiltration membrane has a pore size ...
  36. [36]
    Molecular Weight Cut-off - an overview | ScienceDirect Topics
    Molecular weight cut-off (MWCO) is defined as the molecular weight of an electroneutral solute at which its sieving rejection is approximately 90%. It is ...Missing: formula | Show results with:formula
  37. [37]
    [PDF] Reverse osmosis filtration for space mission wastewater: membrane ...
    concentrations were used to calculate the rejection R, according to. R = 1 −. Cp. Cf. (2) where Cp is the concentration in the permeate and Cf the ...
  38. [38]
    [PDF] HOLLOW FIBER ULTRAFILTRATION - QUA Group
    Filtrate flux range. Maximum feed pressure. Recommended Operating Pressure ... 50 to 150 l/m2h (30 to 90 gfd). 4.8 bar (70 psig). Up to 3.0 bar (44 psig).
  39. [39]
    Exploring the use of ultrafiltration-diafiltration for the concentration ...
    Diafiltration (DF) consists of adding water to the UF retentate to facilitate the transmission of low molecular and non-desired compounds (mainly minerals and ...
  40. [40]
  41. [41]
    Ultrafiltration processes in biotechnology - PubMed
    These uses range from concentration and harvesting of cells to product isolation to production by continuous fermentation and cell culture. As biotechnology ...
  42. [42]
    Recent Advances on the Positively-Charged Nanofiltration ... - NIH
    Jun 22, 2025 · NF membranes, positioned between ultrafiltration (UF) and reverse osmosis (RO) in filtration precision, feature pore sizes of 0.5–2 nm [13].
  43. [43]
  44. [44]
  45. [45]
    [PDF] 1 History of Nanofiltration Membranes from 1960 to 1990 - Wiley-VCH
    This chapter describes the developments in nanofiltration (NF) membranes from the 1960s to the early 1990s that brought NF technology to its current status.
  46. [46]
  47. [47]
    Understanding Salt Passage Vs. Salt Rejection In Reverse Osmosis ...
    Aug 29, 2014 · RO membranes are used to remove dissolved ions in a process that does not rely on distinct pores for filtration. Instead, RO applies diffusion ...<|separator|>
  48. [48]
    Water transport in reverse osmosis membranes is governed by pore ...
    Apr 14, 2023 · The widely accepted theory or mechanism to describe water and salt transport in RO membranes is the solution-diffusion (SD) model, which was ...
  49. [49]
    [PDF] Rejection mechanisms for contaminants in polymeric reverse osmosis
    Aug 10, 2015 · In RO membrane separation, a high pressure is applied to the contaminated feed solution on one side of the membrane that opposes the osmotic ...
  50. [50]
    RO Sizing Calculator (4040/8040) – Quick Element ... - Stark-Water
    How this RO sizing calculator works. This free RO sizing calculator estimates the number of RO elements and the operating flux for 4040 and 8040 skids.Missing: 10-80 bar 10-50 m²h
  51. [51]
    Pressure and Flow Rate Optimization in Reverse Osmosis - XRAY
    Sep 12, 2025 · Reverse osmosis systems operate under precise pressure differentials, typically requiring 6-15 bar for brackish water treatment and up to 70 ...Missing: 10-80 m²h
  52. [52]
  53. [53]
    Quantifying and reducing concentration polarization in reverse ...
    May 15, 2023 · CP is a challenge as it reduces the driving force for transport and leads to a lower overall solute rejection. Compared to membrane fouling, ...
  54. [54]
    PX Q400: Highly Efficient Energy Recovery Device
    Learn how the highly efficient PX Q400 energy recovery device saves energy and operating costs in a typical desalination system.Missing: m³ | Show results with:m³
  55. [55]
    Seawater Desalination Energy Recovery Systems
    Aug 15, 2024 · SECs can be as low as 3 kWh/m3, compared to over 6 kWh/m3 for plants using older technologies.Missing: m³ | Show results with:m³
  56. [56]
  57. [57]
  58. [58]
  59. [59]
    Fluid dynamics of spacer filled rectangular and curvilinear channels
    ... (0.5-2 mm) [123, 124]. Hence, to investigate the effect of spacer geometry on local flow regime in plate-and-frame or SWM modules, CFD simulations can be ...
  60. [60]
    Plate and Frame Membranes - Synder Filtration
    Disadvantages. Low packing density is one major problem for plate and frame membrane systems. Low efficiency compared to other configurations, and high pressure ...Missing: historical | Show results with:historical
  61. [61]
    [PDF] 1 The Basics - Wiley-VCH
    Two main advantages of plate-and-frame modules are the ease of cleaning and replacement of defective membranes and the ability to han- dle viscous feeds. A low ...<|control11|><|separator|>
  62. [62]
    Chapter 5: Design of Membrane Modules for Gas Separations - Books
    Jul 21, 2011 · Plate-and-frame modules were the earliest version of membranes ... The advantages of plate-and-frame modules are:11. Exchange ability of ...
  63. [63]
    Open-channel with plate-and-frame modules - Alfa Laval
    A plate-and-frame installation normally has a higher permeate capacity per unit of membrane area compared with what can be achieved using spiral-wound membrane ...
  64. [64]
    Plate and Frame Membrane Module - ResearchGate
    In addition, they exhibit disadvantages such as: need of several sealings, high pressure drop and low packing density [191]. Currently, this kind of module ...
  65. [65]
    Plate Module - an overview | ScienceDirect Topics
    Compactness and high mechanical strength are the advantages of flat-plate modules. However, their mass and heat transfer rates are limited, as is the contact ...
  66. [66]
    Spiral Wound Membranes - Synder Filtration
    Construction. Spiral-wound elements consist of membranes, feed spacers, permeate spacers, and a permeate tube. First, a membrane is laid out and folded ...
  67. [67]
    Feed Channel Spacers for Spiral-Wound Membranes in Reverse ...
    Material: polypropylene. ; Thickness: 14 – 38 mil. ; Hole Shape: diamond, square. ; Width: 30–55". ; Weight: 80–230 g/m2.
  68. [68]
    Performance Comparison of Spiral-Wound and Plate-and-Frame ...
    Oct 30, 2020 · Element size, diameter: 8, 4, or 2.5 inch, length: 40 inch, various ; Packing density (m2/m3), 300–1000, 100–400 ; Membrane area (m2/element), 37– ...
  69. [69]
    [PDF] FilmTec™ Reverse Osmosis Membranes Technical Manual - DuPont
    Sep 18, 2025 · Figure 13: Construction of spiral-wound FilmTec™ RO Membrane Element ... between 1981 and 1984, the FilmTec Corporation became a wholly owned ...
  70. [70]
    7 Types of RO Membranes - J Mark Systems
    Aug 28, 2023 · The spiral-wound design allows for a large membrane surface area in a compact space, making it efficient for industrial-scale RO systems. Spiral ...
  71. [71]
    [PDF] Membrane Filtration
    Because spiral-wound membranes generally do not permit reverse flow, NF and RO membrane systems are not backwashed. For these systems, membrane fouling is ...
  72. [72]
    Performance enhancement of spiral-wound reverse osmosis ...
    Feb 1, 2022 · Membrane biofouling is a major hindrance limiting the efficiency of spiral wound membrane (SWM) module for water treatment. Feed spacer is an ...
  73. [73]
    Mass transfer and pressure loss in spiral wound modules
    This paper presents some experimental data for pressure drop and mass-transfer characteristics in spiral wound elements and in spacer-filled flow channels.
  74. [74]
    A review of spiral wound membrane modules and processes for ...
    This review provides an insight into the use of pressure-driven membrane processes for groundwater treatment, with focus on the spiral wound membrane module.<|separator|>
  75. [75]
    Increasing Performance of Spiral-Wound Modules (SWMs) by ...
    Sep 12, 2023 · The goal of this study was first to improve filtration performance and cleanability by utilising pulsed flow in a modified pilot-scale filtration plant.
  76. [76]
    Spiral Wound Module - an overview | ScienceDirect Topics
    Spiral wound module is an advanced form of plate and frame model in which two membranes with their active side are placed facing each other separated by a feed ...
  77. [77]
    Hollow‐Fiber Membranes - ResearchGate
    This article briefly discusses the history and theory of hollow‐fiber membrane technology. Methods of manufacture are reviewed together with processing ...
  78. [78]
    [PDF] Water Desalination ReporT - University of Connecticut
    1964 – DuPont textile group begins R&D on nylon-6 fiber RO membrane. 1967 – First commercially successful hollow fiber RO module by DuPont. 1969 – DuPont ...
  79. [79]
    Tubular membranes - Synder Filtration
    Tubular membrane modules are tube-like structures with porous walls, operating via tangential cross-flow and are ideal for high solids / oily applications.Missing: viscous | Show results with:viscous
  80. [80]
    Flux
    J = flux, L/hr/m2 (gal/d/ft2) or Lmh (gfd) · R = recovery of the membrane unit as a percentage · Q = concentrateConcentrate: · TMP = transmembrane pressure, psi.Missing: 0.1-2 bar 100-1000 m²h low-
  81. [81]
    Flummoxed by flux: the indeterminate principles of haemodialysis
    Dec 27, 2021 · 'flux', a finite parameter defined as the volume of fluid (or permeate) transferred per unit area of membrane surface per unit time.
  82. [82]
    Transmembrane Pressure (TMP) Calculator - Online Tools - Cytiva
    The calculation of TMP uses the formula: (Inlet pressure + Outlet pressure)/2 - Permeate pressure.
  83. [83]
    Related Equations | McGraw-Hill Education - Access Engineering
    PT = the transmembrane pressure (TMP) (kPa [psi]), and; PO = the osmotic pressure of the feed solution (kPa [psi]). The above equation clearly shows ...
  84. [84]
    A Brief Review on the Resistance-in-Series Model in Membrane ...
    The present work focused on the role of the cake layer in the fouling analysis and its quantification through the application of the RIS model.
  85. [85]
    Dead-End Filtration - an overview | ScienceDirect Topics
    Dead-end filtration is defined as a process where all incoming water is pushed through a membrane, retaining certain components and solids on the membrane ...
  86. [86]
    [PDF] Crossflow Filtration: Literature Review - Savannah River Site
    Unlike dead-end filtration, which runs a feed perpendicular to a filtering medium, the flow of the feed is parallel to the filter media which helps clear away ...
  87. [87]
  88. [88]
    [PDF] A QUICK GUIDE FOR MEMBRANE SELECTION IN TANGENTIAL ...
    on the membrane type, tangential flow is translated into cross flow velocity for ceramic (2–5 m/s), shear rate for hollow fiber (2000–10000 s-1) and cross flow.
  89. [89]
    Dead End Filtration vs. Cross Flow Filtration: A Complete Comparison
    Sep 21, 2025 · Dead End Filtration results in rapid accumulation of filter cake, leading to higher pressure drop and more frequent cleaning or replacement.
  90. [90]
    Critical flux concept for microfiltration fouling - ScienceDirect.com
    The critical-flux hypothesis is that on start-up there exists a flux below which a decline of flux with time does not occur.
  91. [91]
    A Comprehensive Review on Membrane Fouling - MDPI
    This paper presents a review of the different membrane fouling types, fouling-inducing factors, predictive methods, diagnostic techniques, and mitigation ...
  92. [92]
    Fouling in reverse osmosis membranes: monitoring, characterization ...
    A review on sources, types, mechanisms, characteristics, impacts and control strategies of fouling in RO membrane systems. Desalination Water Treat. 2020 ...
  93. [93]
    Silt Density Index (SDI) - Lenntech
    Silt Density Index testing is a widely accepted method for estimating the rate at which colloidal and particle fouling will occur in water purification systems.
  94. [94]
  95. [95]
  96. [96]
    Effect of backflushing conditions on ultrafiltration of board industry ...
    Earlier tests have shown that backflushing can improve the flux with 11–28% when done for 1 s every minute at a pressure of four bar from the permeate side.
  97. [97]
    SS Tubular Membrane Systems - Spintek Filtration
    The membrane surface can be mechanically cleaned by passing special sponge balls through the tubes. The flow can be reversed for multiple passes if needed.
  98. [98]
    Air scouring! How is it relevant? - Theway Membranes
    Apr 17, 2020 · Air scouring in membranes is the cleaning of the membrane surface on the feed side of the membrane by rapidly abrading, scraping the membrane surface using ...
  99. [99]
  100. [100]
    [PDF] Cleaning Procedures for FilmTec™ Elements - DuPont
    The pH of cleaning solutions used with FilmTec™ Elements can be in the range of 1 –. 13 (see Cleaning Chemicals (Form No. 45-D01622-en) ), and therefore, non- ...
  101. [101]
    Enzymatic cleaning of biofouled thin-film composite reverse osmosis ...
    Subtilisin protease- and lipase-based enzyme cleaning at 100 ppm and for 18 h contact time restored the hydrophobicity of the TFC RO surface to its virgin ...
  102. [102]
    Improved ultrasonic cleaning of membranes with tandem frequency ...
    Oct 1, 2012 · They found that lower frequencies (here 28 kHz) was more effective in cleaning fouled membranes. Similar results were obtained by Lamminen ...
  103. [103]
    RO Membrane Cleaning Procedure: Chemicals, SOP & Recovery ...
    Oct 20, 2025 · 2) When to clean: objective triggers. Trigger, Typical threshold, Catatan. Normalized permeate flow drop, 10–15% below baseline, Use normalized ...
  104. [104]
    Study of Turbulence Promoters in Prolonging Membrane Life - PMC
    Apr 8, 2021 · Installing a turbulence promoter is a promising means of improving the hydraulic conditions inside the membrane chamber.
  105. [105]
  106. [106]
    Fouling behavior of zwitterionic membranes: Impact of electrostatic ...
    ... membrane fouling rate was reduced by approximately 50%. The energy consumption rate for the electrically assisted membrane fouling control was only 22.2 ...
  107. [107]
    Fouling Mitigation in Membrane Distillation Using Pulsation Flow ...
    Sep 15, 2023 · The pulsatile flow reduces the temperature and concentration polarization in the region where the feed and membrane interfaces reside. It turns ...
  108. [108]
    Effect of pressure and cross-flow velocity on membrane behaviour in ...
    Effect of transmembrane pressure (TMP) (10 to 50 bar) and cross-flow rate (1-3 l/min) on the permeate flux are discussed in view of membrane fouling. Higher ...
  109. [109]
    Prevention and management of silica scaling in membrane ...
    Silica scaling in membrane distillation can be reduced with feed pH adjustment. At similar pH and concentration, scaling rates correlate to polymerization ...
  110. [110]
    Online monitoring of MBR fouling by transmembrane pressure and ...
    This study assessed in a critical light the use of TMP and K for membrane fouling monitoring. TMP and K were directly measured on a pilot-scale submerged MBR ...
  111. [111]
    Antiscalant - an overview | ScienceDirect Topics
    Antiscalants are chemical additives that inhibit the precipitation and adherence of scales to membrane ... dosage is in the range of 2–5 ppm. SHMP can prevent ...
  112. [112]
    Antiscalants in RO membrane scaling control | Request PDF
    To prevent scaling of membrane, antiscalant dosing typically may range from 2 to 5 ppm, with a cost of approximately $1.2-1.5 per m 3 of brine treated, ...<|separator|>
  113. [113]
    Pump-less forward osmosis and low-pressure membrane hybrid ...
    This study demonstrates a pump-less forward osmosis low-pressure membrane (FO-LPM) hybrid system for the effective removal of selected PhACs.Missing: lower TMP
  114. [114]
    Cost of fouling in full-scale reverse osmosis and nanofiltration ...
    Mar 15, 2021 · Some studies have reported that the cost of biofouling in a water production membrane system is around 20 to 30% of operating expenditure (OPEX) ...
  115. [115]
    [PDF] Water Desalination using Renewable Energy - IRENA
    Seawater desalination via MSF consumes typically 80.6 kWh of heat energy (290 MJ thermal energy per kg) plus 2.5 to 3.5 kWh of electricity per m3 of water, ...
  116. [116]
    How much energy does desalinisation use? Is it “absurdly cheap”?
    Sep 18, 2024 · Reverse osmosis desalinization uses 2.5 to 3.5 kWh per cubic meter of water, with a theoretical minimum of 1 kWh. Thermal desalinization uses 3 ...
  117. [117]
    Characteristics of Desalination Brine and Its Impacts on Marine ...
    Apr 25, 2022 · The capacity of RO for feed water containing total dissolved solids (TDS) of up to 70,000 mg/L, RO may desalinate with a maximum water recovery ...
  118. [118]
    A better understanding of seawater reverse osmosis brine
    Similarly, total dissolved solids (TDS) can also have an adverse impact on aquatic life and water quality. It is one of the parameters which determines the ...
  119. [119]
    How Often Should You Change Your RO Membranes?
    Nov 12, 2024 · How Often Should You Change Your RO Membranes? Generally, RO membranes can last up to about 1 to 5 years, but the exact lifespan depends ...
  120. [120]
    Environmental impact of desalination technologies: A review
    Dec 15, 2020 · Membrane desalination, using reverse osmosis (RO), accounts for approximately 69% of global desalination capacity and about 85% of the total ...
  121. [121]
    The Global Rise of Zero Liquid Discharge for Wastewater ...
    Jun 8, 2016 · We highlight the evolution of ZLD from thermal to membrane-based processes ... evaporation/crystallization to achieve zero-liquid discharge J.Introduction · Conventional ZLD Systems · beyond Thermal Evaporators... · Outlook
  122. [122]
    Lithium recovery from brine: Recent developments and challenges
    Apr 15, 2022 · This review aims to highlight the recent developments in this field starting from the significance of extracting Li from brine followed by discussing recent ...Missing: 2010s | Show results with:2010s
  123. [123]
    [PDF] Guidelines for the regulation of desalination - ProDes
    Sep 9, 2010 · Regarding the discharge of the concentrate, salt or other pollutants from desalination processes are not explicitly listed in the Water ...
  124. [124]
    Recycling of end-of-life reverse osmosis membranes by oxidative ...
    An additional step of chemical cleaning was conducted with 0.2% (w) citric acid solution (pH 2.5) for 40 min. Thus, removing at least part of the membrane ...<|control11|><|separator|>
  125. [125]
    Recycling of end-of-life reverse osmosis membranes: Comparative ...
    In this work two pilot designs for recycling end-of-life (EoL) RO membranes are evaluated with LCA and cost effectiveness analysis.
  126. [126]
    New approach of recycling end-of-life reverse osmosis membranes ...
    A new approach based on sonication method to convert discarded RO membrane to MF application. · Ultrasonication method is able to convert RO membrane into MF ...
  127. [127]
    Solvent-based recovery of high purity polysulfone and polyester ...
    A facile, solvent-based recycling method is presented to recover high value-added polymers from the end-of-life, Thin Film Composite (TFC) based Reverse ...
  128. [128]
    Membrane Recycling and Fabrication Using Recycled Waste - MDPI
    The pyrolysis process conducted at 600 °C resulted in 28 wt% oil, 17 wt% non-condensable gases, and 22 wt% char production.
  129. [129]
    Recycling and energy recovery solutions of end-of-life reverse ...
    Aug 9, 2025 · Pontié et al. concluded from the TGA study that pyrolysis could lead to waste membrane mass reduction of 48.5% and thermal energy recovery of 86 ...
  130. [130]
    [PDF] Towards new opportunities for reuse, recycling and disposal of used ...
    Jun 21, 2012 · RO membranes [14–17]. Although direct reuse of old membranes without ... As a result, the reported recycling rate was as low as 10% by ...
  131. [131]
    Transforming end-of-life SWRO desalination membranes into ...
    Feb 7, 2025 · In this study, we explore the possibility of reusing end-of-life seawater reverse osmosis (RO) membranes to treat brackish water and industrial effluent.
  132. [132]
    Unlocking the potential of recycling RO membranes | Aquatech
    Nov 8, 2024 · The process uses a sodium hypochlorite oxidation step to strip the compromised polyamide layer, thereby transforming spent RO membranes into ...
  133. [133]
    Recycling and energy recovery solutions of end-of-life reverse ...
    Dec 1, 2017 · The objective of the present work is first to estimate the degradation level of an old RO membrane taken from Mauritania (North West Africa)
  134. [134]
    Advancements in reverse osmosis desalination - ScienceDirect.com
    Apr 1, 2025 · There are around 16,000 operational desalination plants producing around 95 million m3/day of desalinated water [10]. Desalination systems could ...
  135. [135]
    [PDF] Pretreatment in Reverse Osmosis Seawater Desalination: A Short ...
    Pretreatment for seawater RO includes conventional methods like coagulation and granular media filtration, and ultrafiltration/microfiltration (UF/MF).
  136. [136]
    Low energy consumption in the Perth seawater desalination plant
    The plant has a total capacity of 143,000 m 3 /d operating with 45% recovery efficiency and SW feed salinity of 34,000 ppm [144] .
  137. [137]
    (PDF) High performance boron removal from seawater by two-pass ...
    Aug 6, 2025 · Boron removal in seawater RO desalination systems is usually done by a second desalination pass, operating at high pH (pH≥10). This method ...Missing: Perth | Show results with:Perth
  138. [138]
    The Costs of Constructing a Desalination Plant and Facility - Medium
    Jan 7, 2024 · Total Opex averages $0.50–1/m3 (World Bank, 2019). Financing a Desalination Plant Project. Construction and long-term funding options impact ...
  139. [139]
    [PDF] CAPEX, OPEX & Technological Game Changers to Come
    Costs now average around $0.2–0.75/m³, including both brackish and seawater desalination. RO Desalination Technology. 1. Significant development in membrane.
  140. [140]
    Energy Recovery in SWRO Desalination: Current Status and New ...
    Apr 2, 2020 · Seawater RO (SWRO) offers several advantages over other desalination methods including high efficiency and selectivity, easy control and scale- ...Abstract · Introduction · Energy Recovery Devices · RO Process Configurations
  141. [141]
    Energy Recovery Devices in Seawater Reverse Osmosis ...
    The energy requirements of conventional SWRO plants are presently as low as 1.6 kWh/m3, making the process more cost effective and energy efficient than other ...
  142. [142]
    [PDF] Polymeric membrane materials for nitrogen production from air - HAL
    Oct 25, 2021 · A very large number of advanced membrane materials have been recently reported, showing increasing permeability and/or selectivity for air ...
  143. [143]
    Highly Selective Oxygen/Nitrogen Separation Membrane ...
    Sep 3, 2019 · Therefore, developing membranes with both excellent O2 permeability and O2/N2 selectivity for efficient air separation is highly desired.
  144. [144]
    Facilitated transport membranes for CO2/CH4 separation - State of ...
    The results showed that the permeability of both CO2 and CH4 increased with the increase of operating temperature (15–75 ​°C), which was caused by the increase ...
  145. [145]
    Gas Membranes for CO2/CH4 (Biogas) Separation: A Review
    This article is a review of the existing research on polymers as used in the separation of carbon dioxide/methane (CO2/CH4), generally focusing on polyimide and ...
  146. [146]
    [PDF] Optimizing juice processing - Alfa Laval
    Prior to ultrafiltration, the juice needs to be treated with enzymes to ... ○ High juice recovery, approximately 98–99%. ○ Enzyme treatment can be ...
  147. [147]
    Ultrafiltration and Reverse Osmosis Recovery of Limonene from ...
    UF membrane rejections were 78–97% for mixtures with initial limonene concentrations from 0.04–0.6%v/v. RO membrane rejection of limonene ranged from 87–99% for ...
  148. [148]
    High-performance pervaporation membranes for ethanol dehydration
    To achieve fuel-grade purity (≥99.5 ​wt%), raw bioethanol needs to be purified. Adopting pervaporation membrane for bioethanol enrichment can greatly reduce ...
  149. [149]
    (PDF) CO2 removal processes for Alaska LNG - ResearchGate
    A discussion on CO2 removal processes for Alaska LNG covers screening studies; gas separation membranes; the Morphysorb technology, which is based on use of ...Missing: 1980s | Show results with:1980s
  150. [150]
    Removal of Particles by Ultrafiltration Systems
    Microfiltration permits the almost complete removal of bacteria but, in contrast to ultrafiltration, is unreliable regarding the removal of viruses. For this ...
  151. [151]
    Gas Separation Membrane Market | Industry Report, 2030
    The global gas separation membrane market size was valued at USD 1.5 billion in 2024 and is projected to reach USD 2.16 billion by 2030, growing at a CAGR ...
  152. [152]
    Gas Separation Membranes 2026-2036: Materials, Markets, Players ...
    IDTechEx forecasts emerging membrane markets (biogas upgrading, post-combustion capture, blue hydrogen) will see the largest growth in revenue by 2036.