Fact-checked by Grok 2 weeks ago

Membrane technology

Membrane technology encompasses a suite of pressure-driven or diffusion-based separation processes that utilize thin, selective semi-permeable barriers to partition mixtures into concentrated and purified streams, exploiting disparities in molecular size, , , or charge. These membranes, often polymeric or , function without phase changes, offering energy efficiencies superior to thermal methods like in many applications. Key variants include for particle removal, for macromolecules, nanofiltration for multivalent ions, and for , with gas permeation membranes enabling separations like air into oxygen and . Pioneered in the mid-20th century with asymmetric cellulose acetate films for reverse osmosis, the field has scaled to industrial dominance in water purification, producing potable water from brackish sources and enabling large-scale seawater desalination plants that supply over 100 million cubic meters daily worldwide. Achievements include fouling-resistant designs and hybrid systems integrating membranes with bioreactors for wastewater reclamation, reducing energy demands by up to 50% compared to evaporation in food processing and pharmaceuticals. In gas processing, polymer membranes achieve purities exceeding 99% for natural gas sweetening, mitigating CO2 emissions cost-effectively. Persistent challenges involve from adsorbing solutes, which diminishes and necessitates frequent cleaning or replacement, alongside limits from high pressures in that elevate capital costs. Recent advances in thin-film nanocomposites and electrospun fibers address durability, yet economic viability hinges on material innovations to counter degradation under harsh feeds like in . These technical hurdles underscore membrane technology's causal reliance on precise material science for real-world efficacy, rather than overstated claims detached from operational data.

Historical Development

Early Concepts and Initial Applications

The selective permeability of biological membranes, which regulate the transport of solutes and solvents across cellular boundaries, served as a foundational for artificial membrane concepts. These natural structures, observed through early and physiological studies, demonstrated phenomena like and that later guided synthetic mimics. In 1748, French physicist Jean-Antoine Nollet conducted the first documented laboratory demonstration of , using a pig's bladder stretched over a vessel containing and immersing it in , whereby permeated the toward the higher solute concentration, causing visible swelling. This experiment established the principle of semipermeable barriers, though practical artificial alternatives like were explored in the for studies, revealing limitations in durability and control over pore uniformity. By the early , advancements shifted toward synthetic materials for . In , Heinrich Bechhold developed collodion (nitrocellulose) membranes with graded sizes, enabling the separation of colloids and , and he coined the term "" while devising a bubble pressure method to assess dimensions. These isotropic structures, however, exhibited low selectivity and due to uneven distribution and susceptibility to , constraining broader utility. In the 1930s, William J. Elford refined graded membranes for bacteriological applications, particularly virus separation, by controlling preparation variables to achieve average pore diameters from 12 to 240 nm, allowing size-based retention of filterable agents like those causing ectromelia. Concurrently, initial applications emerged, with Georg Haas performing the first human in 1924 using hirudin-anticoagulated blood circulated through semipermeable celloidin tubes, followed by substitution in 1927, though low efficiency from membrane fragility limited clinical . Early gas separation trials remained rudimentary, hampered by poor permselectivity in available materials.

Post-War Advancements and Commercialization

Following , membrane technology accelerated due to U.S. government initiatives addressing water shortages through research. The Office of , created in 1952 under the Saline Water Conversion Act, expanded funding in the 1950s and 1960s to develop practical seawater conversion methods, including . This support facilitated breakthroughs at institutions like UCLA, where Sidney Loeb and Srinivasa Sourirajan developed the first integrally skinned asymmetric membranes in 1959–1960. These membranes featured a thin, selective skin layer atop a porous substructure, enabling water fluxes up to 100 times higher than symmetric membranes while achieving salt rejections exceeding 98%, thus making viable for large-scale . Industrial commercialization followed swiftly, with DuPont launching its Permasep reverse osmosis permeators in 1969 using hollow-fiber configurations based on asymmetric membrane principles. Pilot plants in the late 1960s transitioned to commercial scale-up, culminating in the B-10 permeator introduced in 1974, which processed seawater at capacities of approximately 5.7 m³/day per module and demonstrated economic feasibility for brackish and seawater treatment. Concurrently, the 1970s marked the market entry of ultrafiltration and microfiltration modules, pioneered for applications in food processing and biotechnology, such as whey protein concentration and microbial removal, with early systems operating at low pressures (1–5 bar) to separate particles from 0.1 to 10 µm. Despite these advances, early reverse osmosis systems encountered practical constraints, including rapid fouling and the necessity for elevated trans-membrane pressures (800–1000 psi) to balance flux and rejection, resulting in energy demands of 15–25 kWh/m³—substantial relative to thermal distillation's heat equivalents but highlighting inherent trade-offs where higher rejection demanded reduced permeability per Darcy's law analogs. Membrane lifetimes were initially limited to months, necessitating frequent replacements and elevating operational costs until material refinements improved durability. These limitations underscored causal dependencies on material asymmetry for performance, driving iterative engineering toward modular designs that facilitated broader adoption in municipal and industrial water treatment by the late 1970s.

Recent Milestones (1980s–Present)

In the , membrane technology saw the commercialization of separation systems, such as Monsanto's hydrogen separation unit, which marked the first large-scale application for recovering from refinery streams with selectivity driven by polymeric materials like . Concurrently, nanofiltration () membranes matured from experimental loose variants developed in the , achieving commercial viability by the late for applications like through divalent ion retention exceeding 90% while allowing monovalent passage. emerged as a viable dehydration process in the same decade, leveraging hydrophobic membranes to separate water from organic solvents via vapor pressure differences, with early pilots demonstrating flux rates up to 1-2 kg/m²·h for ethanol-water mixtures. The 1990s and 2000s focused on enhancing membrane flux and durability, particularly through composite thin-film developments. Thin-film composite (TFC) layers on porous supports enabled higher permeate fluxes—often 20-50% improvements over asymmetric —in and NF, reducing energy demands for to below 4 kWh/m³ in optimized systems. These advancements supported broader industrial adoption, including EU-backed research into polymeric gas separation for , where selectivity for CO₂/CH₄ reached 20-30 via rubbery polymers like poly(ethylene oxide). mitigation remained a persistent challenge, with early anti-fouling coatings extending operational life but not eliminating the need for frequent cleaning cycles that increased operational costs by 10-20%. From the 2010s onward, (FO) gained traction for , utilizing osmotic draw solutions to achieve lower propensity than pressure-driven processes, with lab-scale systems recovering up to 80% of water from municipal effluents at fluxes of 10-20 L/m²·h. (MD) pilots in the 2020s targeted in oilfields, demonstrating salt rejections over 99.9% in direct contact configurations, though scalability was limited by thermal efficiency losses from temperature polarization reducing effective fluxes to 5-15 L/m²·h. In June 2025, prototyped an electrically conductive membrane, incorporating conductive polymers to enable in-situ electrochemical cleaning, enhancing salt rejection to levels comparable to traditional polyamides while mitigating organic accumulation. Market integration reflected these efficiency gains, with the global membranes sector expanding from roughly $4 billion in 2010 to approximately $8.3 billion by 2024, propelled by demand in and bioprocessing amid industrial growth in . Projections indicate continued rise to $8.8 billion in 2025, tempered by fouling-related downtime that can comprise 15-25% of operational budgets in high-salinity feeds.

Fundamental Principles

Mass Transfer Mechanisms

In membrane technology, mass transfer across the selective layer occurs predominantly through solution-diffusion or pore-flow mechanisms, each dominant in distinct membrane architectures based on empirical permeation studies. The solution-diffusion mechanism governs transport in dense, non-porous membranes, where solutes or solvents sorb into the polymer matrix at the feed interface, diffuse molecularly through the material, and desorb at the permeate side, with flux proportional to diffusion coefficients measured via time-lag permeation experiments adhering to Fick's laws. In contrast, the pore-flow mechanism prevails in porous membranes, involving sieving and viscous convection of species smaller than pore dimensions under applied pressure gradients, as evidenced by Hagen-Poiseuille flow correlations in microporous structures validated through scanning electron microscopy and flux measurements. These mechanisms are driven by thermodynamic forces, including concentration gradients that induce diffusive flux per Fick's first law, hydrostatic pressure differentials that propel convective flow, and overarching gradients that represent the net motive force for selective permeation, as derived from irreversible and confirmed in non-equilibrium simulations. Empirical permeability tests, such as constant-volume variable-pressure methods, quantify these via species-specific coefficients (typically 10^{-6} to 10^{-10} cm²/s in polymers) and hydraulic permeabilities, revealing that solution- favors selectivity in glassy polymers while pore-flow enhances flux in but risks lower resolution. A key empirical constraint arises in gas separation membranes, where solution-diffusion transport exhibits an inherent permeability-selectivity , encapsulated by Robeson's upper bound—a logarithmic correlation between pure-gas permeability and ideal selectivity, initially formulated in 1991 from over 100 polymers and refined in 2008 with expanded datasets incorporating mixed-gas effects and aging, underscoring causal limits from free-volume distribution and chain rigidity rather than idealized continuum assumptions. Recent molecular simulations challenge the universality of solution-diffusion even in , proposing pore-like nanoscale channels as the actual conduits based on anomalies under high pressure, highlighting how institutional modeling preferences may overlook such data-driven revisions.

Key Theoretical Models

The solution-diffusion model posits that transport through dense, non-porous occurs via sequential dissolution of permeants into the membrane phase followed by diffusion driven by a gradient, without invoking pores or convective flow. The permeate J for a given is expressed as J = P (\Delta c), where P is the intrinsic permeability coefficient (product of S and D) and \Delta c represents the concentration difference across the membrane, often adjusted for activity. This framework, originally formulated by Lonsdale, Merten, and in 1965, has been validated against experimental permeation data for and gas separation processes, where measured fluxes align closely with predictions under dilute conditions and low pressures. However, the model's assumption of uniform solubility and negligible polymer chain mobility leads to systematic deviations in high-concentration feeds, as evidenced by empirical studies showing reduced fluxes due to solute-membrane interactions, plasticization, or clustering that alter diffusivity beyond simple Fickian behavior. Recent analyses, incorporating two-phase partitioning between free volume elements and polymer matrix, better reconcile these discrepancies with neutron scattering and positron annihilation data, highlighting the model's limitations as an idealized approximation rather than a universal causal mechanism. For porous membranes in and , hydrodynamic sieving models treat separation as size-based exclusion coupled with viscous Poiseuille flow through tortuous capillaries, augmented by for macroscopic permeability. The Hagen-Poiseuille equation approximates local flux through cylindrical as J = \frac{\Delta P r^2}{8 \mu \tau L}, where \Delta P is the transmembrane , r the radius, \mu the , \tau the , and L the effective thickness; aggregate forms like the Kozeny-Carman relation extend this to polydisperse structures by integrating over size distributions derived from mercury porosimetry. Experimental validation via pure water permeability tests and solute rejection curves confirms for Newtonian fluids at low Reynolds numbers, though introduces cake-layer resistance modeled as an additional term R_c = \alpha m, where \alpha is specific cake resistance and m mass deposited, as quantified in cross-flow trials. Irreversible thermodynamics frameworks, rooted in Onsager's reciprocal relations, address coupled solvent-solute transport via the Kedem-Katchalsky equations, framing volume flux J_v = L_p (\Delta P - \sigma \Delta \pi) and solute flux J_s = \bar{P} (\Delta c) + (1 - \sigma) \bar{C} J_v, where L_p is hydraulic permeability, \sigma the Staverman reflection coefficient (0-1, indicating sieving efficiency), \bar{P} the solute permeability, and \bar{C} the mean concentration. The Spiegler-Kedem extension of 1966 refines this for concentration-dependent profiles within the membrane, solving differential forms to predict rejection as R = \frac{\sigma (1 - F)}{1 - \sigma F} with F a Peclet-like parameter; parameter estimation from transient permeation experiments yields values aligning with nanofiltration data for electrolytes, though friction coefficients require cautious interpretation due to assumptions of local equilibrium. These models empirically capture electrokinetic and Donnan effects in charged systems but overpredict coupling in highly asymmetric membranes without pore-scale corrections.

Driving Forces and Selectivity Factors

The primary driving forces in membrane separations arise from gradients across the membrane, including differences for hydraulic processes, concentration gradients for diffusive processes, and electrical potential differences for charged separations. In -driven systems such as (RO) and , transmembrane hydraulic (ΔP) serves as the dominant force, propelling solvent flux via approximation, J_v = L_p (ΔP - Δπ), where L_p is the hydraulic permeability coefficient, Δπ is , and deviations from ideality occur due to reducing the effective ΔP. For and , chemical gradients drive selective diffusion, while in gas permeation, partial differences (Δp_i) govern flux according to Fickian modified by , with permeance often quantified in units (1 = 10^{-10} cm³(STP)·cm/(cm²·s·cmHg)) to reflect real kinetic limitations beyond thermodynamic ideality. Selectivity, the ability to discriminate between , stems from kinetic and thermodynamic interactions including steric exclusion ( size versus ), Donnan partitioning (electrostatic repulsion at charged interfaces), and solution-diffusion pathways where differential and yield separation factors α_{i/j} = (P_i / P_j) = (D_i S_i) / (D_j S_j), with P as permeability, D as , and S as . In liquid separations, rejection R = 1 - (C_p / C_f) quantifies solute retention, where C_p and C_f are permeate and feed concentrations, respectively, often exceeding 99% for RO of salts under ideal conditions but influenced by feed composition altering interfacial equilibria. In practice, lab-optimized selectivity overestimates field performance due to membrane compaction under sustained pressure, which densifies the matrix and reduces effective permeability by up to 50% in commercial RO modules at 70 over extended operation, as viscoelastic deformation compresses void volumes without full recovery. Polymer swelling from interactions similarly disrupts selectivity by expanding free volume nonuniformly, correlating with decreased rejection in swollen RO membranes as diffusive pathways broaden preferentially for solutes. These deviations highlight causal limits: while models assume rigid structures, empirical long-term tests reveal thermodynamic non-idealities like increased chain entanglement under load, necessitating factors of 20-30% for design projections versus short-term data.

Membrane Classification

By Material Composition

Organic membranes, predominantly polymeric materials such as , polyamides, and , form the basis of most commercial membrane systems due to their cost-effectiveness, , and straightforward into thin, flexible sheets or fibers. These polymers enable widespread adoption in and gas separation, where production costs remain below those of alternatives, though they suffer from susceptibility to chemical degradation, swelling in solvents, and limited operational temperatures typically under 100°C. Inorganic membranes, constructed from ceramics (e.g., alumina, , zirconia), zeolites, or metals, offer markedly higher resistance to thermal extremes—often exceeding 300–500°C—and aggressive chemicals, including strong acids, bases, and oxidants, making them suitable for and high-temperature gas processing. Their rigidity and inertness reduce and enable aggressive cleaning protocols, contributing to operational lifespans of 10–20 years in abrasive or corrosive environments, compared to 3–5 years for many organics under similar stresses. However, their higher costs and brittleness limit broader use, with ceramic variants representing a niche segment valued at approximately USD 10.67 billion in 2024 amid a larger overall market. Hybrid or mixed-matrix membranes integrate inorganic fillers, such as metal-organic frameworks (MOFs) or nanoparticles, into polymer hosts to balance cost with enhanced performance, particularly in selectivity for gas separations like CO₂/CH₄ or /. Since the , amine-functionalized MOF nanoparticles have demonstrated up to 19% improvements in CO₂/H₂ selectivity through better polymer-filler interfacial compatibility, reducing voids and boosting overall permeability without sacrificing mechanical flexibility. These composites address pure limitations in resistance and transport efficiency while avoiding the expense of fully inorganic systems.

By Pore Size and Structure

Membranes are categorized by pore size into microporous types used in (MF), with pore diameters ranging from 0.1 to 10 μm, allowing retention of particles such as and suspended solids. Mesoporous membranes, applied in (UF), feature pores from 2 to 100 nm, enabling separation of macromolecules like proteins and viruses based on size exclusion. Non-porous or dense membranes, employed in (RO) and , exhibit effective pore sizes below 1 nm, relying on solution-diffusion mechanisms rather than physical sieving for solute rejection. Structural morphology further distinguishes isotropic membranes, which maintain uniform characteristics throughout their thickness, from anisotropic (asymmetric) designs featuring a thin, selective layer atop a thicker porous support for enhanced mechanical stability and flux rates. The asymmetric configuration minimizes resistance in the support while concentrating separation in the dense or finely porous , typically 0.1–1 μm thick, to balance permeability and selectivity. Operational ranges are quantified using metrics like (MWCO), defined as the solute molecular weight at which a membrane retains 90% under specified conditions, derived from retention curves plotting solute rejection against molecular size. Pore size distribution, characterizing heterogeneity in pore diameters, is assessed via mercury porosimetry, where intrusion pressure correlates inversely with pore radius per the Washburn equation, yielding cumulative volume data for microporous and mesoporous structures. These parameters, validated against performance, delineate effective separation thresholds without conflating with material-specific transport.

By Operational Mechanism

Membrane separation processes are categorized by their primary driving force, which determines the mechanism of across the selective barrier and influences energy requirements, scalability, and suitability for specific separations. Pressure-driven processes dominate industrial applications due to their versatility in liquid separations, but they are constrained by the need to overcome osmotic pressures governed by the van't Hoff equation, \pi = cRT, where high solute concentrations demand elevated applied pressures, escalating energy costs and risking membrane compaction. In -driven mechanisms, a transmembrane forces and small solutes through porous or dense membranes, separating based on exclusion or solution-diffusion. () and () target particles and macromolecules above 0.1 \mum and 2–100 nm, respectively, at low pressures (0.1–5 bar), while nanofiltration (NF) and () achieve finer rejections down to ions via denser structures under higher pressures (5–80 bar). , for seawater desalination, typically consumes 2–4 kWh/m³ in modern plants with devices, approaching of ~1 kWh/m³ limited by osmotic work, though actual efficiencies vary with and recovery rates. Concentration-gradient-driven processes rely on differences without mechanical pressure, reducing fouling risks but requiring auxiliary steps for solute recovery. exploits passive across semipermeable membranes driven by solute concentration gradients, primarily for biomedical applications like . () draws water from a feed to a hypertonic draw solution via gradients (often 10–50 bar equivalent), offering lower energy use (~0.5–2 kWh/m³ post-reconcentration) and better tolerance to feed impurities compared to , though draw solute regeneration adds complexity. Electrically-driven mechanisms, such as , apply an to migrate ions through alternating cation- and anion-exchange membranes, concentrating salts in alternate compartments while depleting others. suits monovalent ion separations like brackish water , with energy consumption of 0.7–2.5 kWh/m³ for low-salinity feeds, scaling with and stack resistance, but it underperforms for divalent ions or high concentrations due to effects. Thermal mechanisms utilize temperature-induced or gradients for non-isothermal separations. Membrane distillation (MD) evaporates water through hydrophobic pores under a thermal gradient (typically 40–60°C), condensing permeate on the cold side, with thermal energy demands equivalent to 100–200 kWh/m³ electrical equivalent—10–20 times higher than —though viable with low-grade . Pervaporation drives volatile components across dense membranes by vacuum or carrier gas, favoring or organic recovery at 50–150°C, but its energy efficiency lags pressure-driven methods for large-scale aqueous processing due to requirements. Hybrid systems combining mechanisms, like pressure-assisted osmosis, mitigate individual limitations but introduce integration challenges.

Fabrication Methods

Synthesis of Organic Membranes

Organic membranes, predominantly composed of polymers such as polysulfone, polyethersulfone, and polyamide, are fabricated through scalable processes that exploit phase separation or polymerization reactions to achieve desired microstructures for applications like ultrafiltration and reverse osmosis. Phase inversion remains the dominant method for producing asymmetric porous membranes, involving the dissolution of a polymer in a solvent to form a dope solution, followed by casting into a film or fiber and inducing phase separation via solvent evaporation (dry phase inversion) or immersion in a nonsolvent bath (wet phase inversion). This technique, first developed by Sidney Loeb and Srinivasa Sourirajan in the 1960s for cellulose acetate membranes, generates a thin, dense selective skin layer supported by a porous substructure, enabling high permeate flux while maintaining mechanical integrity. Empirical control of casting parameters—such as polymer concentration (typically 15-25 wt%), solvent-nonsolvent exchange rates, and additives like polyvinylpyrrolidone—dictates pore size distribution and defect minimization, with optimized conditions yielding ultrafiltration membranes exhibiting water fluxes exceeding 100 L/m²·h·bar under standard pressures. Interfacial polymerization constitutes a cornerstone for thin-film composite (TFC) membranes, particularly in , where a crosslinked selective layer, approximately 50-200 nm thick, is formed on a microporous support. The process entails saturating the support with an of (e.g., m-phenylenediamine at 0.1-2 wt%) followed by exposure to an immiscible organic phase containing acid chloride (e.g., trimesoyl chloride at 0.05-0.2 wt%), triggering rapid polycondensation at the liquid-liquid to produce an ultrathin, defect-free barrier. Pioneered by John Cadotte in the 1970s as an advancement over early asymmetric designs, this method decouples the selective layer from mechanical support, allowing independent optimization for enhanced performance. Reaction variables, including monomer concentrations, cure temperature (often 60-90°C), and contact time (seconds to minutes), critically influence layer thickness and crosslinking density, with well-controlled syntheses achieving NaCl rejection rates exceeding 98% at fluxes of 20-50 L/m²·h·bar in tests. These empirical outcomes underscore the causal role of kinetic barriers in minimizing voids, as verified through cross-sectional imaging and salt permeation assays in peer-reviewed validations.

Inorganic and Hybrid Membrane Production

Inorganic membranes, primarily ceramics such as alumina (Al₂O₃) and silica (SiO₂), are synthesized via the sol-gel process, which entails of metal alkoxides to form a sol, followed by gelation, coating onto porous supports (e.g., tubular alumina substrates), drying, and at temperatures typically between 400–600°C to yield thin microporous layers with pore sizes in the 1–10 nm range. This method enables high thermal and , with silica-alumina composites achieving H₂ permeance up to 10⁻⁶ mol·m⁻²·s⁻¹·Pa⁻¹ under conditions optimized for defect-free films. Subsequent at 1200–1700°C densifies these structures into tubular forms, enhancing mechanical integrity for high-temperature gas separation applications, where membranes withstand up to 800°C and corrosive environments like those in processing. Zeolite membranes, valued for their crystalline frameworks, are produced through involving precursors, alkali sources, and often organic templating agents (e.g., tetrapropylammonium hydroxide) at 100–200°C for 24–72 hours, followed by to remove templates and activate uniform micropores of approximately 0.3–0.5 diameter. Template-free variants, using seed layers on supports like alumina disks, reduce costs and environmental impact while maintaining selectivity, as in membranes with Si/Al ratios >20 yielding CO₂/H₂ separation factors exceeding 10 under conditions. These pores enable precise molecular sieving based on size exclusion, outperforming amorphous ceramics in kinetic diameter-based separations (e.g., CHA-type zeolites at 0.38 for CO₂ over larger hydrocarbons). Carbon molecular sieve (CMS) membranes derive from inorganic-like carbon structures formed by pyrolyzing polymeric precursors (e.g., polyimide or cellulose) at 600–900°C under inert atmospheres, creating rigid, slit-shaped micropores (<0.7 nm) via controlled carbonization and cross-linking to achieve high selectivity in gas separations. Yields typically reach 20–30% of precursor mass, with hollow fiber geometries post-pyrolysis exhibiting H₂/CO₂ permeabilities >1000 Barrer and selectivities >50 at elevated temperatures up to 200°C, suitable for harsh feeds like natural gas upgrading. Hybrid inorganic-organic membranes integrate nanoparticles such as TiO₂ (5–20 wt%) into matrices (e.g., PVDF or polyethersulfone) during dope preparation via phase inversion or interfacial , followed by and annealing to embed hydrophilic, photocatalytic domains that mitigate . This embedding reduces adhesion by enhancing surface hydrophilicity and ROS generation under UV, with modified membranes showing flux recovery ratios >90% after protein challenges compared to <70% for pristine polymers, and up to 50–70% lower biofilm growth in long-term MBR tests. Such hybrids retain polymeric processability while inheriting inorganic durability, operating stably in aqueous environments with pH 2–12 and temperatures to 80°C.

Advanced Techniques and Modifications

Surface modifications, such as plasma grafting and zwitterionic coatings, have been developed since the early 2000s to enhance antifouling properties of polymeric membranes, reducing protein adsorption and biofouling by forming hydration layers that repel contaminants. Plasma-mediated grafting of zwitterionic polymers onto nanofibrous membranes improves filtration performance while maintaining mechanical durability, as demonstrated in studies showing sustained antifouling over extended operation. Similarly, tannic acid-based coatings on poly(methylpentene) membranes achieve a 70.58% reduction in protein adsorption, enhancing hydrophilicity and biocompatibility for applications like blood-contacting devices. These modifications often yield 20-50% increases in water flux through reduced fouling resistance, though long-term stability can diminish due to coating degradation under high shear or chemical stress. Biomimetic approaches, inspired by aquaporin water channels, integrate selective proteins into synthetic membranes to achieve ultra-high permeability exceeding traditional reverse osmosis limits by orders of magnitude. Post-2000 innovations, such as embedding in block copolymer matrices, enable water fluxes up to 100 times higher than conventional membranes while preserving salt rejection above 99%, as validated in thin-film composite prototypes. However, challenges persist in protein stability and scalability, with empirical data indicating flux declines of 10-30% over months due to denaturation or vesicle rupture under operational pressures. Additive manufacturing via 3D printing allows fabrication of custom membrane geometries with hierarchical porosity, optimizing flow paths and surface area for targeted separations since the 2010s. Techniques like produce multilayered structures mimicking biological tissues, boosting permeability by 30-40% in desalination tests through tailored channel designs that minimize concentration polarization. Trade-offs include reduced mechanical robustness in printed polymers compared to cast membranes, necessitating hybrid reinforcements for industrial viability, as iterative prototypes reveal 15-25% flux gains offset by higher fouling propensity in complex feeds.

Module Designs and Operational Configurations

Membrane Geometries and Forms

Membrane geometries encompass the physical arrangements of membranes in modules, tailored to enhance hydrodynamic efficiency, maximize surface area utilization, and minimize operational challenges such as pressure losses and fouling propensity. Primary configurations include , , , and , each suited to specific feed characteristics and process scales based on empirical hydrodynamic data. Flat-sheet membranes, typically employed in plate-and-frame modules, feature membranes sandwiched between support plates, yielding packing densities of 100–400 m²/m³. These geometries facilitate straightforward membrane replacement and cleaning but exhibit lower hydrodynamic shear rates, rendering them less effective for high-fouling applications. Tubular membranes consist of cylindrical elements with diameters ranging from 5 to 25 mm, achieving packing densities of 30–300 m²/m³. Their design promotes elevated wall shear rates, particularly beneficial for processing viscous or particulate-laden feeds, where higher shear reduces fouling accumulation compared to flat-sheet configurations; studies indicate tubular modules sustain flux longer in such conditions due to diminished concentration polarization. Hollow-fiber membranes, formed by bundles of capillary tubes with inner diameters of 0.1–1 mm, deliver exceptionally high packing densities exceeding 1,000 m²/m³, optimizing compactness for space-constrained systems. Originating in the 1960s for , hollow-fiber modules gained dominance post-1970s, exemplified by 's 1971 polyamide-based innovations, enabling efficient seawater desalination through enhanced surface-to-volume ratios that lower overall pressure drops in large arrays. Spiral-wound modules, constructed by rolling flat-sheet membranes with spacers around a central permeate collection tube, offer packing densities of 300–1,000 m²/m³, balancing scalability and cost for industrial deployments. Feed spacers induce secondary flows that elevate local shear, mitigating fouling, though they contribute to elevated trans-membrane pressure drops—typically 0.1–0.5 bar higher than tubular equivalents under comparable fluxes—necessitating optimized spacer designs for viscous industrial streams. The selection of geometry profoundly affects module hydrodynamics, with higher shear in tubular and hollow-fiber formats correlating to reduced fouling rates in empirical tests; for instance, tubular configurations demonstrate 20–50% lower flux decline in viscous dairy processing versus flat-sheet due to augmented turbulent mixing at the surface.
GeometryPacking Density (m²/m³)Key Hydrodynamic AdvantageTypical Pressure Drop Range (bar)
Flat-sheet100–400Low resistance, easy access0.1–0.3
Tubular30–300High shear for fouling control0.2–1.0
Hollow-fiber>1,000Maximal compactness, low hold-up volume0.1–0.5
Spiral-wound300–1,000Spacer-induced mixing0.3–1.5

Flow Dynamics and Process Modes

In dead-end filtration, the feed stream flows perpendicular to the surface, with permeate passing through while all retained solutes and particles accumulate directly on the , forming a cake layer that progressively increases hydraulic resistance and necessitates periodic cleaning or replacement. This configuration suits low-solids feeds where short filtration cycles are acceptable, but it promotes rapid in particulate-laden suspensions due to unmitigated buildup. Cross-flow filtration, conversely, directs the feed tangentially across the , inducing shear stresses that disrupt boundary layer formation and sweep away deposited materials, thereby sustaining higher permeate fluxes over extended periods. The ( = ρvd/μ, where ρ is fluid density, is , d is , and μ is ) governs the transition from laminar to turbulent regimes; at > 2100, enhanced turbulence thins the layer—the solute-enriched film adjacent to the that reduces effective driving force—thus improving selectivity and flux . Empirical studies in and of suspensions demonstrate that cross-flow configurations reduce rates by factors of 2 to 5 compared to dead-end modes, as evidenced by prolonged operational run lengths in pilot-scale trials with or slurries. Process modes in membrane operations range from single-pass continuous setups, which yield modest recoveries limited by escalating retentate concentration, to staged cascades that recycle retentate across multiple units for enhanced overall . In for brine management, single-pass systems typically achieve 40-50% recovery before osmotic pressures curtail , whereas multi-stage cascades—employing intermediate pressurization and partial permeate withdrawal—enable recoveries exceeding 75% by distributing concentration gradients and minimizing energy penalties from high disposal. Batch modes predominate in dead-end for intermittent processing, while continuous cross-flow supports steady-state industrial applications, often augmented by the (Sc = μ/ρD, ratio of momentum to ) in modeling diffusive within layers to predict extent. These dynamics underscore causal trade-offs: tangential causally suppresses deposition via mechanical forces, yet demands higher pumping power, balancing longevity against operational costs.

Equipment and System Integration

Membrane systems require robust ancillary equipment, including high-pressure pumps capable of delivering feed pressures up to 80 bar for processes and corrosion-resistant housings such as fiberglass-reinforced pressure vessels to encase membrane elements securely. These components ensure stable operation by maintaining uniform flow distribution and preventing mechanical failure under operational stresses. Pretreatment units, integrated upstream, typically incorporate filters, chemical dosing for antiscalants to inhibit precipitation, and coagulants to , thereby extending lifespan by mitigating precursors. Monitoring instrumentation, such as differential pressure sensors for transmembrane pressure () assessment and flux meters for permeate flow tracking, enables real-time detection of performance degradation, with TMP rises signaling onset. System integration often employs modular skids—pre-assembled frames housing pumps, membranes, and controls—for plug-and-play deployment, facilitating from pilot to full-scale operations, as seen in multi-skid units processing up to several million gallons per day in facilities. A key integration challenge in high-pressure systems like (SWRO) is energy consumption, addressed by isobaric energy recovery devices such as PX pressure exchangers, which transfer hydraulic energy from to feed, achieving efficiencies of 93-98% and reducing overall power use by 25-40%. These devices, when paired with variable-frequency-drive pumps, optimize energy recovery while minimizing mixing losses.

Performance Evaluation

Governing Equations and Metrics

![{\displaystyle L_{p}={\frac {J}{\Delta p}}}][float-right] Permeate flux J, defined as the of permeate per unit area (J = Q_p / A), quantifies the of processes and is typically expressed in liters per square meter per hour (LMH). This metric is central to design calculations, where higher flux reduces required area but often correlates with increased risk under constant pressure operation. Solute rejection R, calculated as R = 1 - (C_p / C_f), measures the membrane's ability to retain solutes, with C_p and C_f denoting permeate and feed concentrations, respectively; values approach 100% for effective separation in processes like (). Permeability, such as hydraulic permeability L_p = J / \Delta p, describes intrinsic transport under a transmembrane difference \Delta p, enabling comparison across membrane types while assuming and negligible osmotic effects in . Stage-cut \theta, the ratio of permeate flow to feed flow (\theta = Q_p / Q_f), indicates recovery efficiency, typically limited to 10-50% in single-stage to minimize ; higher values necessitate multi-stage configurations for economic viability. For permeate quality in desalination, salt passage SP = C_p / C_f inversely reflects rejection, with commercial membranes exhibiting 0.1-1% passage under standard conditions, influenced by feed and . Scaling prediction metrics, such as the Stiff-Davis Index for silica (SDI = \mathrm{pH} - \mathrm{pH_s}), assess risk where positive values signal potential on membranes, guiding antiscalant dosing in high-silica feeds like . These indices rely on models validated against pilot data, yet overpredict scaling in dynamic flows due to kinetic barriers. Governing equations often presume steady-state conditions without , deriving from or solution-diffusion models, but operational data reveal transient declines from cake layer formation, invalidating assumptions in prolonged runs and necessitating dynamic modeling for accurate predictions. Field validations, such as in plants, show initial overestimations by 20-50% when ignoring early fouling transients.

Characterization Techniques

Characterization techniques for membranes focus on measuring intrinsic properties such as , structure, and surface charge to assess material performance independent of operational conditions. These methods emphasize non-destructive and reproducible approaches to ensure comparability across studies and standards compliance. Key techniques include for visual structure, porosimetry for dimensions, and electrokinetic measurements for charge properties. Scanning electron microscopy (SEM) and (TEM) provide detailed insights into , including surface , cross-sectional , and distribution. SEM offers high-resolution imaging of surface features and is commonly used for polymeric membranes, requiring such as to enhance conductivity. TEM enables visualization of internal structures at nanoscale , particularly useful for thin or dense layers, though it often involves destructive sectioning. These techniques reveal asymmetries like finger-like voids in phase-inversion membranes, influencing selectivity. For porous membranes, the bubble point method determines maximum pore size by applying increasing gas pressure to a wetted membrane until the first bubble emerges, corresponding to the largest pore via the . Standardized in ASTM F316, this test applies to filters with pores from 0.1 to 15.0 μm and provides a non-destructive check. Mean flow pore tests complement it by assessing average pore size through flow rates at specific pressures. Limitations include sensitivity to wetting fluid and inability to resolve full pore distributions without extensions like capillary flow porometry. Molecular weight cut-off (MWCO) characterizes and nanofiltration membranes by measuring solute rejection, defined as the molecular weight at which 90% retention occurs, often using (PEG) standards or proteins like . Challenge tests involve pressure-driven filtration of graded solutes, with retention calculated from concentration ratios in permeate and feed. This indirect method correlates nominal pore size to separation capability but varies with solute shape and membrane-solute interactions. Dense membranes, such as those for gas separation, undergo pure gas testing to quantify permeability coefficients via pressure-driven flow through the membrane, following solution-diffusion models. is measured for gases like , , and CH4 under controlled conditions, enabling selectivity calculations as ratios of single-gas permeabilities. Multi-lab validations confirm reproducibility, though mixed-gas tests reveal deviations due to competitive . ASTM and similar protocols normalize data to standard temperatures and pressures. Streaming potential measurements evaluate membrane surface charge via , derived from pressure-induced generating electric potentials across the membrane. The Helmholtz-Smoluchowski relates streaming current to , influenced by and . This technique aids in understanding electrostatic interactions in charged membranes, with applications in prediction, though it requires careful cell design to minimize errors from non-uniform . Standardization, such as ASTM F316 for metrics and normalized tests using pure under defined conditions (e.g., 25°C, 1 ), ensures inter-laboratory comparability and focuses on intrinsic properties like hydraulic permeability = J/ΔP. These protocols mitigate variability from fabrication differences, prioritizing empirical validation over theoretical assumptions.

Fouling Mechanisms and Control Strategies

Membrane fouling refers to the accumulation of substances on or within the membrane structure, resulting in increased hydraulic resistance and reduced permeate flux. This process encompasses biological fouling from microbial adhesion and (EPS), organic fouling from natural (NOM) such as humic acids and proteins, and inorganic fouling from mineral scaling like precipitation. Colloidal and contribute via cake layer formation, exacerbating all types through blocking and surface deposition. The primary mechanisms driving foulant attachment involve intermolecular forces described by , which balances attractive van der Waals forces against repulsive electrostatic double-layer interactions to determine deposition kinetics on the membrane surface. In favorable conditions—such as high compressing the electrical double layer or low reducing repulsion—foulants overcome the energy barrier, leading to irreversible adhesion rather than reversible transport. For biofouling, initial bacterial attachment follows DLVO predictions, followed by maturation via EPS matrix growth, which embeds further deposits and resists shear. Organic foulants like proteins adsorb through hydrophobic interactions and hydrogen bonding, while inorganic scales form via and when local exceeds limits, often catalyzed by prior organic layers. These kinetics highlight that is not merely but a thermodynamically driven deposition process, with extended DLVO (XDLVO) incorporating acid-base interactions improving predictions for complex feeds. Control strategies primarily address reversible fouling through physical methods like backwashing, which reverses permeate flow to dislodge cake layers, and enhanced hydrodynamics such as cross-flow or air scouring to minimize stagnation. Chemical cleaning employs acids (e.g., citric or hydrochloric) to dissolve inorganic scales by chelating divalent cations, and alkaline-oxidant solutions (e.g., NaOH with ) to hydrolyze organic and biofoulants, typically restoring 70-90% of initial in short-term cycles. However, empirical long-term operations reveal persistent declines of 20-50% due to irreversible pore constriction and surface modification, as fails to fully reverse deep-layer adhesions or remnants. In plants, fouling-related costs—encompassing , chemical usage, and —account for approximately 24% of operational expenditures (OPEX), underscoring how incomplete mitigation sustains elevated trans-membrane pressures and curtails efficiency. This persistent impact challenges claims of membrane scalability without rigorous accounting for energy penalties from compensatory pumping.

Major Applications

Water Purification and Desalination

Membrane processes dominate modern desalination, with reverse osmosis (RO) comprising approximately 69% of global installed capacity as of recent assessments, primarily for treating seawater and brackish water to produce potable supplies. Nanofiltration (NF) complements RO in brackish water applications by selectively removing divalent ions like calcium and sulfate while permitting monovalent salts to pass more readily, enabling softening without full demineralization. These pressure-driven processes achieve high solute rejection—typically over 99% for salts in seawater RO—yielding permeate purities suitable for drinking water standards, though actual performance depends on feed salinity, recovery rates (often 40-50% for seawater), and membrane integrity. Pretreatment via () and () is essential to protect / membranes from by particulates, organics, and biota in raw , reducing silt density index (SDI) to below 3 and extending membrane life. pores (0.1-10 μm) target larger and , while (0.001-0.1 μm) addresses finer colloids and macromolecules, often replacing conventional coagulation-flocculation-sedimentation with lower chemical use and generation. In large-scale plants, / pretreatment can cut rates by 50-90%, enabling higher and while minimizing cleaning frequency. The Desalination Plant in , commissioned in with a capacity of 144,000 m³/day, exemplifies efficient implementation, utilizing dual-pass RO with devices to achieve consumption below 4.2 kWh/m³ for the full process, including pretreatment and post-treatment. Its first-pass RO stage records 2.4 kWh/m³, among the lowest for large-scale seawater at the time, reflecting optimized management and minimal losses. Contemporary seawater RO systems maintain 2.5-3.5 kWh/m³ total use, far below alternatives (10-15 kWh/m³ equivalent), though still 3-4 times the thermodynamic minimum of ~1 kWh/m³ due to irreversibilities in pumping and mixing. Despite high product purity, RO desalination generates hypersaline —1.5-2 times feed —disposed via ocean outfalls, which elevates local , , and chemical loads (e.g., antiscalants, biocides), potentially harming benthic ecosystems through and . Life cycle assessments quantify these impacts, attributing 20-50% of total environmental footprint to discharge and energy-related emissions, with marine eutrophication and acidification risks amplified in discharge zones lacking dilution. Mitigation via diffuser design or zero-liquid discharge (e.g., evaporation ponds) adds costs but reduces ecological disruption, as evidenced by site-specific monitoring showing recovery radii of 1-5 km for plumes. Operational yields remain constrained by limits, with recovery capped below 60% for to avoid excessive scaling, underscoring trade-offs between output volume and concentrate management.

Gas Separation and Purification

Polymeric membranes dominate industrial gas separation due to their ability to exploit differences in gas permeability via solution-diffusion mechanisms, enabling applications such as oxygen/ separation from air and / separation in . Common materials include glassy polymers like , , and , often configured as hollow modules for high surface area. For air separation, systems like ' , commercialized in the early 1980s, use -based membranes to produce -enriched streams up to 99% purity in single-stage operation, with oxygen permeance typically in the range of 10-50 GPU depending on module design. Compared to cryogenic , which achieves high-purity (>99.5%) and at large scales (>100 /day) but incurs costs of approximately 200-300 kWh/ from and , membrane processes offer lower (around 100-200 kWh/ equivalent) and simpler operation for mid-scale (20-100 /day) or decentralized applications. However, membranes yield lower single-stage purity (e.g., 95-99% N2) and recovery (<50% for ), often requiring multi-stage cascades or hybrid integration for demanding specifications, limiting their economic edge to low-volume scenarios despite higher initial capital costs. For CO2/CH4 separation, polymeric membranes achieve selectivities of 20-40 under mixed-gas conditions, facilitating acid gas removal to meet pipeline specifications (<2-4% CO2), though performance degrades with increasing pressure due to competitive sorption. In the 2020s, hybrid approaches combining polymeric membranes with inorganic fillers (mixed matrix membranes) or absorption processes have advanced post-combustion CO2 capture, targeting flue gas streams with permeance exceeding 100 GPU for CO2 and selectivities >30 for CO2/N2. These innovations aim to surpass pure polymeric limits, yet membrane-based capture constitutes less than 10% of deployed post-combustion capacity, overshadowed by scrubbing due to persistent plasticization—where high CO2 induces chain swelling, reducing selectivity by up to 50%—and the need for high-pressure feeds to achieve viable flux. Empirical gas pair selectivities in commercial polymeric membranes remain below theoretical maxima derived from solution-diffusion parameters, constrained by the permeability-selectivity encapsulated in Robeson's upper bound, which reflects intrinsic free volume limitations and non- mixed-gas effects like competitive . For O2/N2, practical selectivities hover at 3-7 versus potential values >10 in defect-free models, while CO2/CH4 pairs achieve 20-50 against bounds approaching 100, underscoring the gap between lab and industrial realities influenced by fabrication defects and aging.

Bioprocessing and Medical Uses

In , synthetic membranes such as are widely employed in configurations to remove uremic like from blood, achieving removal efficiencies exceeding 90% under optimized conditions. These membranes offer high permeability for small solutes while maintaining , reducing complement activation compared to cellulosic alternatives. However, certain membrane materials can induce production, such as interleukin-1 and tumor necrosis factor-alpha, potentially exacerbating inflammation in patients, as evidenced by studies linking bioincompatible surfaces to activation during . Regulatory evaluations, including those aligned with FDA standards for , emphasize minimizing such responses through material modifications like to enhance adsorption without excessive protein loss. Ultrafiltration plays a critical role in bioprocessing for pharmaceutical applications, enabling protein concentration and buffer exchange via tangential flow filtration, where molecules are separated based on with molecular weight cut-offs typically 1–1000 . In production, / steps concentrate solutions from initial low densities, such as 5 mg/mL to over 40 mg/mL, facilitating high-titer substance while controlling aggregation risks under pressure-limited operations. For processes, membranes harvest cells post-cultivation, replacing in continuous systems to recycle and boost productivity, as demonstrated in propionic acid production from with yields up to 1 g/L/h. Biomass-derived materials like and offer niche antifouling enhancements in bioprocessing membranes, improving hydrophilicity and recovery to near 100% in modified setups, though synthetic polymers predominate due to superior sterility and scalability in sterile environments. Challenges persist in for medical membranes, where induction from less optimized polymers underscores the need for rigorous testing, with synthetic options like showing reduced inflammatory markers relative to older cellulosic types. These applications highlight membrane technology's precision in separating biologics while navigating and limitations.

Industrial Separations

In industrial separations, membrane technologies facilitate the processing of high-volume commodity streams, emphasizing through reduced energy use and minimal chemical inputs, alongside membrane robustness to handle complex, often fouling-prone feeds like those laden with salts, organics, or emulsions. Applications span sectors requiring separation of solvents, salts, and biomolecules from bulk liquids, where pressure-driven or osmotically driven processes outperform traditional methods like or in targeted scenarios. In the oil and gas industry, forward osmosis (FO) and membrane distillation (MD) address produced water treatment, a byproduct generated at ratios of 7–10 barrels per barrel of oil extracted. Pilot-scale FO-reverse osmosis (RO) hybrids applied to raw produced water have demonstrated over 99% rejection of ions and non-volatiles, enabling volume reduction despite high salinity exceeding 100,000 mg/L TDS. MD variants similarly achieve near-complete solute rejection in hypersaline feeds, with field pilots confirming operational stability against oil and grease contaminants as of 2023. These systems support reinjection or reuse, though pretreatment is essential to mitigate scaling from divalent ions like calcium and barium. Ultrafiltration (UF) dominates dairy processing for whey valorization, separating proteins from lactose-rich permeate in cheese production byproducts. Commercial UF systems concentrate whey proteins to 80% purity, enabling recovery of 20–30% of milk's total protein content that would otherwise be discarded, with continuous operations in multi-loop plants handling thousands of cubic meters daily to meet global demand exceeding 200 million tons of annually. This scalability underscores volume-driven economics, as UF reduces discharge while producing high-value concentrates for and nutraceuticals. Pervaporation excels in dehydrating azeotropic mixtures like ethanol-water, bypassing distillation's energy-intensive reflux requirements. Hybrid pervaporation-distillation pilots for bioethanol production have reported up to 50% energy reductions relative to standalone distillation, leveraging selective hydrophilic membranes to permeate at fluxes of 1–5 kg/m²·h while retaining purity above 99.5%. Semi-pilot tests with ceramic membranes confirm lower operational costs, particularly when integrated downstream of distillation columns to break the azeotrope. Fouling by organics, prevalent in feeds like or with hydrocarbons and proteins, limits by 20–50% over extended runs, driven by adsorption and layer formation that elevates trans-membrane . Mitigation via cross-flow configurations and periodic chemical sustains productivity, though organics-heavy streams demand feed to preserve integrity and achieve commercial viability.

Challenges and Criticisms

Technical Limitations and Failure Modes

Membrane compaction, a primary failure mode in pressure-driven systems like () and nanofiltration (), arises from viscoelastic deformation under transmembrane pressure, compressing the porous support layers and active skin, which reduces effective and elevates hydraulic resistance. In membranes, this manifests as an initial decline of approximately 20-40% within hours to days of operation at pressures around 300 psi, with total thickness reductions up to 21%, primarily in the and backing layers. Compaction accelerates with higher pressure, temperature, and cyclic loading, leading to irreversible morphological changes that diminish without recovery upon pressure relief, distinct from reversible elastic deformation. Chemical compromises membrane integrity through oxidative or hydrolytic attacks on chains, particularly in aromatic used for . Chlorine exposure above 0.1 free chlorine rapidly oxidizes the active layer, causing chain scission and cross-linking within 200-1,000 hours at 1 , which elevates salt passage by degrading selectivity while sometimes increasing initial before overall decline. This vulnerability stems from the electron-rich bonds reacting with , catalyzed by transition metals like iron, resulting in performance loss concentrated in upstream elements; combined exacerbate the issue by gradual free chlorine release. In gas separation membranes, such as those employing glassy , similar occurs via plasticization or sorbate-induced swelling, though less quantified in operational statistics. Pinhole defects, often originating from manufacturing inconsistencies like uneven coating or handling-induced scratches, create localized breaches in the selective layer, bypassing rejection mechanisms and allowing solute leakage. In RO systems, undetected pinholes can result in permeate losses of less than 1% flux through compromised areas, but aggregated across elements, they contribute to 1-5% overall purity degradation, detectable via elevated conductivity or tracer tests. These flaws, verified through leak or bubble-point integrity tests, stem from incomplete polymerization or mechanical stress, leading to cascading failure if pressure differentials widen the defects. Empirical membrane lifespans in operational settings average 2-5 years for elements, falling short of laboratory projections of 5-10 years due to unmodeled variables like feed fluctuations, intermittent , and synergistic stressors. In full-scale plants, accelerates after 3-5 years, evidenced by 10-15% normalized loss and 5-10% passage rise, attributable to cumulative compaction, partial , and defect propagation rather than isolated events. For modules, real-world durability similarly trails idealized models, with integrity breaches rising in variable feeds versus controlled lab simulants.

Economic and Scalability Barriers

Membrane systems incur substantial capital expenditures (CAPEX), with large-scale seawater () plants costing $30–$50 million for a of 10 million gallons per day (MGD), or roughly $4,000 per m³/day of installed , driven by membrane modules, pressure vessels, and supporting infrastructure. Operational expenses (OPEX) are dominated by , where without energy recovery devices demands 5–6 kWh/m³ or higher for typical recovery rates around 25–50%, debunking notions of membrane processes as inherently low- absent such add-ons, which recover only a of the hydraulic input. Pretreatment, critical to prevent and extend membrane life, accounts for 30–40% of total CAPEX in systems, encompassing structures, , and chemical dosing to handle feedwater variability. While facilitates scaling from small pilots to industrial plants, deployment hurdles persist, particularly in developing countries where high upfront costs combine with shortages in skilled O&M personnel, leading to underutilization and reliance on imported expertise despite potential for decentralized applications. Compared to thermal alternatives like multi-stage , membranes offer long-term cost advantages through lower energy intensity— uses 3–5 times less per m³—yielding water costs as low as $0.41/m³ in optimized cases versus higher figures. However, for membranes often includes a premium of 20–40% over non-membrane treatments like conventional or in non-desalination contexts, attributable to periodic membrane replacements (every 5–7 years) and elevated energy demands relative to lower-pressure alternatives.

Environmental and Sustainability Concerns

Discarded membrane modules, primarily composed of non-biodegradable polymers like and , generate substantial waste, with approximately 840,000 modules—exceeding 14,000 tonnes of material—discarded annually worldwide as of 2023, a figure projected to rise significantly by 2025 due to expanding capacity. These materials resist degradation and fragment into upon environmental exposure or landfilling, contributing to long-term in and systems, as plastics persist and break down into persistent microparticles rather than fully decomposing. Recycling rates remain below 10%, hampered by technical challenges in separating composite layers and economic disincentives, leading to predominant landfilling or that releases toxins and gases. Brine discharge from membrane-based desalination processes exacerbates hypersalinity in receiving waters, particularly in enclosed basins like the , where cumulative inputs have elevated average by up to 0.5 practical salinity units since the 1970s, stressing ecosystems through osmotic shock to organisms and altered benthic communities. In the Gulf region, hypersaline effluents—often 1.5 to 2 times concentration—have been linked to reduced , heavy metal bioaccumulation in sediments, and seagrass die-offs, with diffuse discharge strategies failing to mitigate basin-wide accumulation due to limited flushing. These impacts persist despite regulatory efforts, as volumes match or exceed desalinated output, concentrating trace contaminants and nutrients that promote localized . The energy intensity of pressure-driven membrane processes, such as requiring 2–6 kWh per cubic meter for treatment, translates to a substantial when powered by fossil fuels, equivalent to emissions of 1.4–4.32 kg CO₂ equivalent per cubic meter and mirroring the fuel consumption of major industrial sectors in non-renewable grids. For instance, in energy-dependent regions like the consumes over 770,000 MWh annually, akin to burning 180,000 tons of fossil fuels yearly, underscoring reliance on non-renewable sources absent widespread renewables . While membrane-enabled water reuse reduces freshwater and offsets some upstream environmental costs, these benefits do not fully negate disposal and burdens, as net savings diminish with scaling and without closed-loop . Bio-based membranes derived from offer theoretical advantages through renewability and lower toxicity, yet their production yields lower compared to synthetics, resulting in higher overall environmental impacts from extended processing and limited for volumes. Synthetic polymers maintain superior and durability at large scales, where biomass alternatives face fouling vulnerabilities and supply chain constraints, rendering hype around "green" variants unsubstantiated for replacing established technologies without trade-offs. Lifecycle assessments confirm that while bio-membranes reduce certain virgin material dependencies, their deployment remains niche, unable to match synthetics' performance in high-throughput applications essential for claims.

Future Prospects

Emerging Innovations (2020–2025)

Electrically conductive membranes for electro-enhanced represent a key advancement, transitioning from lab development to field pilots. In June 2025, Active Membranes, licensing technology from UC Riverside and UCLA, conducted a pilot in Ventura County using membranes coated with a conductive layer. Applying low voltage creates an that repels ions via charge-based mechanisms, reducing and while boosting operational efficiency. The pilot achieved 20-30% higher production rates compared to standard membranes, enhancing treatment of brackish and for agricultural and industrial applications. Hybrid forward osmosis-membrane distillation (FO-MD) configurations have advanced toward zero-liquid discharge in processing, with bench-scale validations demonstrating high . A 2021 study applied FO-MD to , attaining 89% water using a 3.4 M NaCl draw solution, alongside >99.8% rejection of key ions (Na⁺, K⁺, Ca²⁺, Mg²⁺, Cl⁻, SO₄²⁻) and permeate below 105 μS/cm. Water fluxes reached 37 L/m²h in and 25 L/m²h in at 65.5°C, supported by low-grade for . Subsequent 2020s reviews affirm FO-MD hybrids routinely enable 85-90% in saline effluents by combining with vapor separation, facilitating solid concentration for disposal or . Machine learning integration in membrane fabrication has accelerated optimization, focusing on property prediction and structural refinement from 2023-2025. , employing featurization techniques like group contributions and , screen vast libraries to forecast permeability, selectivity, and flux, reducing reliance on trial-and-error synthesis. This has yielded designs such as materials with ~4x higher water flux than benchmarks and optimizations improving rejection by 8%. By minimizing fabrication variability and structural defects through data-driven parameter tuning, these methods support faster lab-to-pilot scaling, as evidenced in and nanofiltration applications.

Research Directions and Potential Breakthroughs

Research in membrane technology increasingly targets overcoming fundamental permeability-selectivity trade-offs through atomically thin materials, such as and other two-dimensional () structures, which theoretically enable flux rates up to ten times higher than conventional polymeric membranes due to reduced transport across nanopores. However, large-scale fabrication remains unproven, with challenges in defect-free and limiting practical deployment beyond lab prototypes, as evidenced by ongoing efforts to optimize lamellar stacking for applications. These materials' potential hinges on resolving causal barriers like precise pore size control to achieve ultimate selectivity limits, without which empirical gains in separation efficiency cannot be realized at industrial scales. Anti-fouling strategies represent another focal direction, emphasizing nanotechnology-enabled self-cleaning surfaces—such as stimuli-responsive coatings or bioinspired structures—that aim to restrict flux decline to under 10% annually by dynamically repelling or degrading foulants. Recent advancements, including chelation-directed interfaces on membranes, demonstrate in-place regeneration under operational conditions, yet real-world tests reveal persistent vulnerabilities to complex effluents, underscoring the need for causal validation of mechanisms over simulated models. Prioritizing empirical metrics like sustained permeate quality in hybrid feeds is essential to transcend hype-driven claims from grant-funded studies. Hybrid membrane architectures integrating adsorption or photocatalysis offer promising trajectories for multifunctional separations, where in-situ contaminant capture or degradation addresses intrinsic flux limitations of standalone membranes. For instance, photocatalytic membrane reactors combine separation with reactive species generation for pollutant mineralization, potentially enhancing overall efficiency in wastewater streams, but require rigorous techno-economic modeling to quantify capital and energy costs against unintegrated alternatives. Future breakthroughs demand first-principles scrutiny of synergistic effects, favoring data-driven simulations over optimistic projections to ensure scalability without subsidizing inefficiencies.

References

  1. [1]
    [PDF] 1 Introduction to Membrane Technology - Wiley-VCH
    1.1 Overview of Membrane Technology. Membrane technology is a general term used for a range of different separation processes.
  2. [2]
    [PDF] Membrane Technology and Applications
    ... overview of membrane science and technology; this book is an attempt to give such an overview. The book starts with a series of general chapters on membrane ...
  3. [3]
    [PDF] Module 19: Membrane Filtration - DEP
    Membrane filtration is a rapidly expanding field in water treatment. There are many different types of filters available in a wide range of pore sizes and ...
  4. [4]
    Membrane Technologies in Wastewater Treatment: A Review - PMC
    Membrane technology has emerged as a favorite choice for reclaiming water from different wastewater streams for re-use. This review looks at the trending ...
  5. [5]
    A Review on the Use of Membrane Technology Systems in ...
    This review aims to delve into the progress of membrane technology for water treatment systems, particularly in developing countries.
  6. [6]
    Challenges, Opportunities and Future Directions of Membrane ... - NIH
    Membrane technology is used for natural gas purification to remove impurities like CO2 and H2S, and is low cost and energy efficient. Hollow fiber membranes ...
  7. [7]
    Challenges and trends in membrane technology implementation for ...
    This review covers membrane technology implementation, performance, challenges, and trends for produced water treatment, driven by environmental regulations ...
  8. [8]
    Grand Challenges in Membrane Modules and Processes - Frontiers
    Challenges in Membrane Processes​​ They offer great solutions to the problems such as water stress, energy production, reductions in mineral resources, etc. in ...
  9. [9]
    Once upon a time the cell membranes: 175 years of cell boundary ...
    Dec 19, 2014 · When the Cell Theory was first formulated in the XIXth century, almost nothing was known about the cell membranes. It was not until well into ...Missing: 19th | Show results with:19th
  10. [10]
    History of Reverse Osmosis | When Was RO Discovered
    Feb 18, 2020 · The process of osmosis through semipermeable membranes was first observed in a laboratory setting in 1748 by Jean-Antoine Nollet, using a pig's bladder as a ...
  11. [11]
  12. [12]
    Novel high performance hollow fiber ultrafiltration membranes spun ...
    The first such membranes were produced in 1907 by Bechhold, who produced collodion membranes thus the term “ultrafiltration”. Since Bechhold's original ...Missing: filtration | Show results with:filtration
  13. [13]
    [PDF] Overview of Membrane Science and Technology
    In 1907, Bechhold devised a technique to prepare nitrocellulose membranes of graded pore size, which he determined by a bubble test [1]. Other early workers ...
  14. [14]
    THE PREPARATION OF THE GRADED COLLODION MEMBRANES ...
    The method described by Elford for the preparation of graded collodion membranes suitable for ultrafiltration was found to give excellent results, and his ...Missing: 1930s | Show results with:1930s
  15. [15]
    A new series of graded collodion membranes suitable for general ...
    Aug 10, 2025 · The first virus removal filters, which were based on graded nitrocellulose membranes, were developed by Elford in the 1930s [2][3][4] [5] [6][7] ...
  16. [16]
    [The early history of the artificial kidney] - PubMed
    Haas, who in 1924 conducted the first dialysis using hirudin, in 1927 conducted the first hemodialysis using heparin. Further hemodialysis treatments followed.
  17. [17]
    Records of the office of Saline Water - National Archives
    Administered research and development programs in the area of desalinization pursuant to the Saline Water Conversion Act (66 Stat. 328), July 3, 1952.Missing: desalination | Show results with:desalination
  18. [18]
    Nor Any Drop to Drink | Science History Institute
    Nov 13, 2018 · By 1955, with the drought worsening, Congress established the Office of Saline Water, quintupled funding for the effort, and in 1958 tasked the ...
  19. [19]
    [PDF] Reverse Osmosis: Introduction - S. Loeb
    In 1959–1960 Loeb and Sourirajan (L–S) found that anisotropic cellulose acetate membranes, those possessing a skin surmounting a porous substructure, could have ...
  20. [20]
    Synthetic Membranes - ACS Publications - American Chemical Society
    The advent of the Loeb-Sourirajan asymmetric membrane some twenty years ago gave birth to an industry now exceeding 200 million dollars in annual sales.
  21. [21]
    Collection: DuPont Permasep Products records
    DuPont introduced its first reverse osmosis permeators for water desalination in 1969 under the trade name "Permasep" as a result of its contusion research in ...
  22. [22]
    [PDF] Engineering Aspects of Reverse Osmosis Module Design - Lenntech
    The first Permasep Hollow Fiber B-10 Permeators from DuPont were introduced commercially in Europe in 1974. These were 4'' elements which had a capacity of 5.7 ...
  23. [23]
    crossflow microfiltration
    In the early Seventies, Ultrafiltration (U.F.) and. Reverse Osmosis (R.O.), were being commercialized in food industry applications. One of the first ...
  24. [24]
    History & Brands | Kovalus Separation Solutions
    In the early 1970s the company began to focus more on membrane related technologies and developed a variety of leading-edge ultrafiltration products and systems ...
  25. [25]
    Effect of Thermodynamic Restriction on Energy Cost Optimization of ...
    However, energy cost remains a major contributor to the total cost of water produced by RO membrane desalination. Reduction of the overall cost of water ...<|control11|><|separator|>
  26. [26]
    A Critical Review on Membrane Technology and Its Application in ...
    The first successfully designed gas separation system for industries, the Monsanto Prism hydrogen separation system, was developed in the 1980s, marking the ...
  27. [27]
    History of Nanofiltration Membranes from 1960 to 1990
    Jul 9, 2021 · The history of nanofiltration (NF) development is unique and began in the 1960s based on the realization that membrane selectivity in the ...
  28. [28]
    Pervaporation as a Successful Tool in the Treatment of Industrial ...
    Apr 14, 2022 · Pervaporation emerged in the 1980s and is now becoming a popular membrane separation technology because of its intrinsic features such as ...
  29. [29]
    A recent progress in thin film composite membrane: A review
    This paper reviews the recent research progress of the TFC membrane science and technology, particularly in the fields of water-related separation processes.
  30. [30]
    Energy-efficient polymeric gas separation membranes for a ...
    Aug 16, 2013 · This review presents the fundamental scientific principles underpinning the operation of polymers for gas separations, including the solution-diffusion model.
  31. [31]
    Membrane Filtration: Past, Present, and Future - 2021 - Journal AWWA
    Oct 4, 2021 · In addition, a key advantage of membrane filtration, compared with media filtration, is superior filtered water quality. Because water quality ...
  32. [32]
    Forward osmosis technology for water treatment - ScienceDirect.com
    Jan 20, 2021 · Forward osmosis is an emerging osmosis-driven membrane process, which has benefits of low energy consumption, low operational pressure, easy equipment, and low ...
  33. [33]
    [PDF] Laboratory and pilot-scale studies of membrane distillation for ...
    implemented membrane-based technologies for the treatment and reuse of flowback. 447 and produced water from shale gas and oil plays: A review, Desalination.
  34. [34]
    New Membrane Technology Could Expand Access to Water for ...
    Jun 30, 2025 · A new technology equips membranes used in the reverse-osmosis desalination process with electrical conductivity, improving their ability to separate salts and ...
  35. [35]
    Membranes Market Size, Share, Growth | Forecast Report [2032]
    The global membranes market size was valued at USD 8.27 billion in 2024. The market is projected to grow from USD 8.79 billion in 2025 to USD 14.21 billion by ...
  36. [36]
    Membranes Market Size, Industry Share Forecast [Latest]
    The global membranes market is expected to reach USD 11.70 billion by 2030, up from USD 7.87 billion in 2025, growing at a CAGR of 8.2% during the forecast ...
  37. [37]
    The solution-diffusion model: a review - ScienceDirect.com
    In this paper we will derive the phenomenological equations for transport in these processes using the solution-diffusion model and starting from the ...
  38. [38]
    Solvent Transport in Disordered and Dynamic Membrane Pores
    Aug 15, 2025 · (4) The basic premise of the SD model is that solvent molecules “dissolve in the membrane material and molecularly diffuse through it as a ...<|separator|>
  39. [39]
    A two-phase model that unifies and extends the classical ... - Science
    Jul 7, 2022 · The pore-flow model, based on fluid mechanics, works for porous membranes, whereas the solution-diffusion model invokes molecular diffusion to ...
  40. [40]
    Water transport in reverse osmosis membranes is governed by pore ...
    Apr 14, 2023 · MD simulations and solvent permeation experiments show that water transport in reverse osmosis membranes is governed by pore flow. INTRODUCTION.
  41. [41]
    [PDF] Permeation and Measurement Techniques - TAPPI.org
    Chemical potential, μ, is the actual driving force for permeation. Chemical potential is a thermodynamic property related to pressure. μ. L μ.
  42. [42]
    Gas permeation through a glassy polymer membrane
    The flux equation for gas transport through a glassy polymer membrane is expressed in terms of a chemical potential gradient. It is an implicit function of ...
  43. [43]
    Comparison of modeling methods for the determination of effective ...
    Sep 25, 2009 · This study presents and discusses various solutions, analytical and semianalytical, to Fick's law for the typical through-diffusion experiment ...
  44. [44]
    [PDF] 29 Chapter 4. Permeability, Diffusivity, and Solubility of Gas and ...
    The gas transport through rubbery polymers is described according to Fick's law for diffusion. The diffusion coefficient is known to be concentration ...
  45. [45]
    The upper bound revisited - ScienceDirect.com
    The empirical upper bound relationship for membrane separation of gases initially published in 1991 has been reviewed with the myriad of data now presently ...
  46. [46]
    Designing exceptional gas-separation polymer membranes using ...
    May 15, 2020 · The 1991 and 2008 Robeson upper bounds are shown as solid black lines. Each data points represents a single polymer; Robeson plots typically ...
  47. [47]
    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 ...
  48. [48]
    Mechanisms and models for water transport in reverse osmosis ...
    Oct 27, 2023 · Following this line of argument, Lonsdale and coworkers proposed the solution–diffusion model which assumes molecular diffusion down a water ...
  49. [49]
    [PDF] The solution-diffusion model
    Sep 7, 2024 · The solution-diffusion model: a review. J. Membr. Sci. 107 (1–2), 1–21. Yang, S., Qu, J., 2012. Computing thermomechanical properties of ...
  50. [50]
    Comparison of classical hydrodynamic models of transport through ...
    Jul 8, 2025 · Classically, three hydrodynamic models of porous media permeability are used: Hagen-Poiseuille (HP), Kozeny-Carman (KC), and Happel's cell (HC) ...
  51. [51]
    [PDF] MEMBRANE SEPARATIONS
    Oct 7, 2016 · Two basic models2 for mass transfer through the membrane will be considered: the hydrodynamic model for porous membranes; the solution-diffusion ...
  52. [52]
    Spiegler-Kedem Model - an overview | ScienceDirect Topics
    Transport of solutes through nanofiltration membranes can be described by the equations of Spiegler and Kedem, which combine both diffusive and convective ...
  53. [53]
    Some remarks on the Kedem-Katchalsky equations for non ...
    The Kedem-Katchalsky equation for the flow of a non-electrolyte through a homogeneous membrane is shown to be a first order expansion of an exact integral.
  54. [54]
    Membranes Separation - an overview | ScienceDirect Topics
    When there is a certain driving force (such as pressure difference, concentration difference or potential difference) on both sides of the membrane, the ...
  55. [55]
    Introduction to membrane science
    This driving force is often a pressure, heat, or concentration gradient, depending on the membrane process that is implemented. Separation processes utilizing ...
  56. [56]
    Membranes – Chemical Engineering Separations: A Handbook for ...
    Introduction to Membrane Processes and Modeling Porous Membranes. \Delta P = pressure driving force across the membrane (pressure). \Delta z = membrane ...
  57. [57]
    Principles of Membrane Technology (Pressure-driven processes)
    Jun 30, 2025 · In order to carry out the process, a driving force is necessary, in this case pressure. There are in essence four over lapping pressure ...
  58. [58]
    MEMBRANE PROCESSES - Thermopedia
    In crossflow filtration, the flow is parallel to the membrane and the drag forces close to the membrane wall reduce the amounts of particulate material ...
  59. [59]
    Membrane Selectivity → Term - Energy → Sustainability Directory
    Apr 15, 2025 · Meaning → Membrane selectivity is the capability of a membrane to preferentially separate components within a mixture. → Term.
  60. [60]
    Selectivity Analysis for Ultrafiltration: Effect of Pore Geometry - NIH
    The results clearly demonstrate that membranes with slit-shaped pores have higher performance, ie, greater selectivity at a given value of the permeability.
  61. [61]
    Rejection coefficient - Oxford Reference
    A measure of the amount of a solute in a feed solution fed to a semi-permeable membrane that does not permeate the membrane in a separation process.
  62. [62]
    Membrane Rejection - an overview | ScienceDirect Topics
    Membrane rejection is defined as the fraction of a specified size solute that is retained by an ultrafiltration membrane, typically characterized by the ...
  63. [63]
    [PDF] Journal of Membrane Science - NSF PAR
    May 30, 2020 · In this work, we demonstrated that compaction decreases the water permeability of com- mercial RO membranes from 2.0 L m 2 h 1 bar 1 at 70 bar ...<|control11|><|separator|>
  64. [64]
    Compaction of Pressure-Driven Water Treatment Membranes: Real ...
    Oct 3, 2024 · The initial high decrease in impedance in step 1 indicates instantaneous compaction that accounts for 93% of the total compaction in this step.
  65. [65]
    Permeability and selectivity of reverse osmosis membranes
    This is because higher swelling has been shown to result in membranes with lower water permeability, as water molecules struggle to diffuse through membrane ...Missing: real | Show results with:real
  66. [66]
    Thin-Film Composite Membrane Compaction - NIH
    A thin-film composite (TFC) membrane support compaction model based on realistic parameters, incorporating viscoelasticity and Monte Carlo flux calculations,Missing: drop | Show results with:drop
  67. [67]
    Effect of operating environment on membrane performance
    May 8, 2020 · Membrane performance can be significantly impacted by the molecular components present in real feed streams.
  68. [68]
    Membrane Materials: Organic v. Inorganic - Synder Filtration
    Some advantages that inorganic membrane possesses are high thermal and chemical stability, inertness to microbiological degradation, and ease of cleaning after ...
  69. [69]
    Emerging membrane technologies for sustainable water treatment
    Jan 17, 2025 · This review examines the recent advances in membrane technologies, focusing on innovations that enhance the efficiency and environmental compatibility
  70. [70]
    Comparison of Organic Membranes and Inorganic Membranes
    Organic Membrane has relatively poor tolerance to environmental factors such as pH, temperature, oxidants, and solvents.Missing: disadvantages durability
  71. [71]
    Inorganic Membranes: Preparation and Application for Water ... - NIH
    Inorganic membranes have the advantage of resisting harsh chemical cleaning, high temperature and wear resistance, high chemical stability, long lifetime, and ...
  72. [72]
    Advantages and disadvantages of inorganic membranes with ...
    The major advantages of inorganic membranes when compared with polymeric membranes, is their high chemical, thermal, and mechanical stability and wide tolerance ...
  73. [73]
    Why Ceramic Membranes Excel in Harsh Chemical Filtration
    Aug 6, 2025 · First and foremost, their lifespan can exceed 5 to 10 years, far outlasting polymeric alternatives. Their high mechanical strength makes them ...
  74. [74]
    Life and operation of ceramic membrane? - ResearchGate
    Oct 17, 2023 · High-quality ceramic membranes can last for 10-20 years or more, while lower-quality ones may have a shorter lifespan.Missing: harsh | Show results with:harsh
  75. [75]
    Ceramic Membrane Market Size | Industry Report, 2030
    The global ceramic membrane market size was estimated at USD 10.67 billion in 2024 and is projected to reach USD 21.17 billion by 2030, growing at a CAGR of 12 ...
  76. [76]
    Enhanced selectivity in mixed matrix membranes for CO2 capture ...
    Jun 5, 2017 · Mixed matrix membranes (MMMs) for gas separation applications have enhanced selectivity when compared with the pure polymer matrix, ...
  77. [77]
    Rational design of mixed-matrix metal-organic framework ... - Science
    Jun 2, 2022 · These mixed-matrix membranes exhibited improved carbon dioxide and methane selectivity relative to many other comparable ones, as well as ...
  78. [78]
    Mixed Matrix Membranes with Surface Functionalized Metal ...
    Apr 12, 2023 · The optimal membrane with a high MOF loading (up to 45 wt.%) displays a propylene permeability of ≈95 barrer and a mixed propylene/propane ...
  79. [79]
    Increasing both selectivity and permeability of mixed-matrix ...
    Aug 1, 2017 · The unique design of the core-shell MOF NPs can enhance both the membrane permeability and selectivity simultaneously.
  80. [80]
    The Classification of Membrane Filters - Hawach
    Oct 9, 2025 · Microfiltration (MF) Membranes: MF membranes have a relatively large pore size ranging from 0.1 to 10 microns. · Ultrafiltration (UF) Membranes: ...
  81. [81]
    What are the 4 Types of Membrane Filtration?
    Jan 7, 2025 · Ultrafiltration (UF) operates at a finer level than microfiltration, with membrane pores in the range of 1 to 100 nanometers. This process can ...1. Microfiltration (mf) · 2. Ultrafiltration (uf) · 3. Nanofiltration (nf)
  82. [82]
    Nanofiltration - Dive into Ceramic Membrane Technology - LiqTech
    NF has a pore size in the range of 0.001-0.01, and RO has a pore size in the range of 0.0001-0.001 μm. As nanofiltration membranes have a larger pore size ...
  83. [83]
    [PDF] Membranes: Introduction
    Structure and morphology: Anisotropic membranes. • Anisotropic membranes combine the high selectivity of a dense membrane with the high permeation rate of a ...
  84. [84]
    Molecular weight cutoff (MWCO) - Lenntech
    Molecular weight cutoff (MWCO) is an important term in membrane filtration. It describes the pore size or rejection capability of a membrane.
  85. [85]
    Pore Size Distributions in Microporous Membranes II. Bulk ...
    The bulk pore size distributions are directly obtained for each filter. The statistical distributions are studied thoroughly and those obtained from both ...
  86. [86]
    Analysis of mercury porosimetry for the evaluation of pore shape ...
    The use of mercury porosimetry to determine pore size and pore distribution is currently applied in a variety of areas where powders or membranes are utilized.
  87. [87]
    MEMBRANE PROCESSES - Thermopedia
    ... driving force provide criteria for a classification of the membrane separation processes, as shown in the table. It should be noted that the boundaries ...
  88. [88]
    Membrane Technology - an overview | ScienceDirect Topics
    In the pressure-driven membrane separation process, the solution is divided into permeate and retentate phases by applying pressure as a driving force to the ...
  89. [89]
    Table 1 .1: Driving forces in membrane processes [2]. - ResearchGate
    Polymeric hollow fiber membrane is produced by a physical process called wet or dry/wet phase inversion; a technique includes many steps and depends on ...
  90. [90]
    Membrane Processes: An Introduction
    These categories are microfiltration (MF), ultrafiltration (UF), nanofiltration (NF), and reverse osmosis (RO).
  91. [91]
    A new world record in SWRO energy efficiency underscores the ...
    Aug 6, 2024 · With an SEC of 1.86 kWh/m3, an experimental SWRO plant on the Canary Islands showcases record-breaking energy efficiency.<|separator|>
  92. [92]
    A comprehensive review of pressure and osmosis driven membrane ...
    Apr 1, 2025 · Pressure-driven processes use hydraulic pressure, like reverse osmosis. Osmotically-driven processes use osmotic pressure differences, like  ...
  93. [93]
    The difference between forward and reverse osmosis - Aquaporin
    Reverse osmosis uses hydraulic pressure, while forward osmosis uses natural osmotic pressure with a saline draw solution, making it less sensitive to fouling.
  94. [94]
    Energy Efficiency of Electro-Driven Brackish Water Desalination
    Feb 21, 2020 · This study systematically compares the energy consumption in electrodialysis (ED) and membrane capacitive deionization (MCDI) over a broad ...Introduction · ED Process Model and... · What Makes ED More Energy...<|control11|><|separator|>
  95. [95]
    Energy efficiency of direct contact membrane distillation
    The study suggested that while RO is still more energy efficient, multistage membrane distillation can be utilized for better heat recovery higher recovery ...
  96. [96]
    Membrane distillation at the water-energy nexus - RSC Publishing
    We begin by exploring how the energy efficiency of MD is limited by the thermal separation of water and dissolved solutes. We then assess the performance of MD ...<|separator|>
  97. [97]
    A review on fabrication, characterization of membrane and the ...
    This review compares polymeric membrane fabrication techniques, characteristics, and factors responsible for effective membrane separation for different ...
  98. [98]
    A Review on Porous Polymeric Membrane Preparation. Part I - NIH
    4.1. Phase Inversion. Presently, most of the commercial polymer membranes are produced via phase inversion methods due to their simple processing, flexible ...
  99. [99]
    Journal of Applied Polymer Science | Wiley Online Library
    Mar 26, 2015 · Introduction. The phase inversion process was first introduced in membrane technology by Loeb and Sourirajan47 in the 1960's and is the base for ...ABSTRACT · INTRODUCTION · THE PROCESS of PHASE... · CONCLUSIONS
  100. [100]
    Phase-Inversion - an overview | ScienceDirect Topics
    This method is also known as the Loeb-Sourirajan technique and was discovered by Sidney Loeb and Srinivasa Sourirajan [11]. They used this technique for ...
  101. [101]
    Fabrication of desalination membranes by interfacial polymerization
    Apr 13, 2021 · ... TFC membranes was largely ascribed to the poor rejection of NaCl salt. For instance, based on projected NaCl rejection, the area beneath ...
  102. [102]
    Thin film composite membrane — Recent development and future ...
    Jan 15, 2015 · The concept of interfacial polymerization has been known for more than 45 years since it was first introduced by Cadotte and Morgan in 1965 [10] ...
  103. [103]
    Fabrication of desalination membranes by interfacial polymerization
    ... Cadotte developed the first thin-film. composite (TFC) polyamide ... thin-. film composite (TFC) polyamide membranes for desalina-. tion, J. Membr ...
  104. [104]
    Impact of pilot-scale PSF substrate surface and pore structural ...
    Sep 1, 2021 · Our best SWRO membrane exhibits an excellent NaCl rejection of 99.5% together with high water permeance of 1.72 L m−2 h−1 bar−1 under seawater ...
  105. [105]
    Nanomaterials-modified reverse osmosis membranes
    Jun 12, 2024 · The PA layer is responsible for providing selectivity, with typical salt rejection of about 99%. TFC membranes have a high selectivity and water ...
  106. [106]
    Silica membranes by the sol-gel process - ScienceDirect.com
    The sol-gel process allows the preparation of thin microporous layers. This technique has been used to prepare silica membranes with controlled thickness ...
  107. [107]
    [PDF] Production and Characterization of a Silica-Alumina Membrane ...
    Dec 20, 2018 · The obtained pore structure shows high potential for membrane manufacture. The silica layer was produced by the sol-gel method and dip-coated on.
  108. [108]
    Preparation of Ceramic Membranes and Their Application in ... - MDPI
    The sol–gel process is an established method of producing ceramic membranes by producing a sol by hydrolysis. The sol–gel process and subsequent heat treatment ...
  109. [109]
    Template-free ZSM-5 membrane preparation on alumina support by ...
    Zeolite membranes stand out for presenting molecular sized uniform pores. The current work aimed to optimize supported template-free ZSM-5 zeolite membrane ...
  110. [110]
    Sustainable and resource-efficient synthesis of zeolite membranes ...
    Jun 26, 2025 · Zeolite membranes exhibit optimal pore sizes and thus facilitate size-based selective separation of molecules, thereby enhancing separation ...
  111. [111]
    Two-step activation of SSZ-13 zeolite membranes for mild template ...
    Typically, SSZ-13 exhibits a CHA-type zeolite topology with a uniform pore size of 0.38 nm [15], which facilitates the fast diffusion of CO2 while excluding C2H ...
  112. [112]
    Cellulose-Based Carbon Molecular Sieve Membranes for Gas ...
    Carbon molecular sieve membranes (CMSM) are prepared by the thermal decomposition, at high temperatures, of polymeric precursor materials under controlled ...
  113. [113]
    Carbon hollow fiber membranes for a molecular sieve with precise ...
    Jan 11, 2021 · Carbon molecular sieve (CMS) membranes have rigid pore structures and are fabricated by controlled carbonization of polymeric precursors at high ...
  114. [114]
    Preparation of carbon molecular sieve membranes with remarkable ...
    Nov 15, 2020 · Carbon hollow fiber membranes (CHFMs) were fabricated based on cellulose hollow fiber precursors spun from a cellulose/ionic liquid system.
  115. [115]
    Preparation of Nano-TiO2-Modified PVDF Membranes with ... - NIH
    Mar 31, 2022 · Titanium dioxide (TiO2) nanoparticles have been applied in membrane antifouling performance modification for years.
  116. [116]
    High-Performance TiO 2 Nanotubes/Poly(aryl ether sulfone) Hybrid ...
    Mar 23, 2019 · To maximize the self-cleaning and anti-fouling properties of hybrid membranes, high surface area TiO2 nanotubes (TNTs) with excellent ...
  117. [117]
    Investigation of antifouling properties of polypropylene/TiO 2 ...
    This paper presents the results of aeration rate effect on antifouling properties of polypropylene (PP)/TiO 2 nanocomposite membrane in membrane bioreactor ( ...
  118. [118]
    Antifouling polymeric nanocomposite membrane based on ... - PeerJ
    Jun 14, 2023 · TiO2 nanoparticles are encouraging nanomaterials for organic–inorganic hybrid TFC membranes due to these properties (Yang et al., 2011).
  119. [119]
    Hybrid Organic/Inorganic Reverse Osmosis (RO) Membrane for ...
    The hybrid thin-film-composite (TFC) aromatic polyamide membranes were prepared by self-assembly of the TiO2 nanoparticles on the polymer chains with COOH ...
  120. [120]
    Recent Progress of Zwitterionic Materials as Antifouling Membranes ...
    Aug 31, 2021 · Facile plasma grafting of zwitterions onto nanofibrous membrane surface for improved antifouling properties and filtration performance.
  121. [121]
    Durable plasma-mediated zwitterionic grafting on polymeric ... - Nature
    Mar 5, 2024 · Zwitterionic chemistries are one of the more promising candidates for antifouling surface modifications as they avoid entropic and energetic ...
  122. [122]
    Preparation of hydrophilic and antifouling coatings via tannic acid ...
    Mar 6, 2025 · Following modification, protein adsorption on the PMP membrane was reduced by 70.58%, and its blood compatibility was significantly enhanced.
  123. [123]
    Maximizing the right stuff: The trade-off between membrane ...
    Jun 16, 2017 · Membrane separations are often limited by available driving force, so increases in membrane material selectivity result in little or no gain in ...
  124. [124]
    Highly permeable and shelf-stable aquaporin biomimetic membrane ...
    Feb 15, 2024 · Aquaporin (AQP) biomimetic membranes are a coming-of-age technology for water purification. Although several studies have reported aquaporin ...Missing: innovations | Show results with:innovations
  125. [125]
    Biomimetic membranes as potential tools for water purification
    May 15, 2019 · Biomimetic membranes, like aquaporin, are emerging as a potential solution for water purification, showing superior performance in water flux ...Missing: innovations | Show results with:innovations
  126. [126]
    Pathways and Challenges for Biomimetic Desalination Membranes ...
    Sep 4, 2020 · In this review, we critically assess the status and potential of biomimetic desalination membranes. We first review channel chemistries and their transport ...Missing: post- | Show results with:post-
  127. [127]
    Biomimetic 3D Prototyping of Hierarchically Porous Multilayered ...
    Jan 23, 2025 · This study introduces a biomimetic approach to 3D printing multilayered hierarchical porous membranes (MHMs) using Direct Ink Writing (DIW) technology.
  128. [128]
    Perspective on 3D printing of separation membranes and ...
    Feb 1, 2017 · Here we provide key background information on 3D printing technologies and applications in membrane engineering.
  129. [129]
    3D Printed and Conventional Membranes—A Review - PMC
    Mar 3, 2022 · This review article compares a conventional method of manufacturing polymer membranes to additive manufacturing. A review of 3D printed membranes is also done.Missing: geometries | Show results with:geometries
  130. [130]
    Membrane Configuration - an overview | ScienceDirect Topics
    Four main types of membrane configurations are used in the food industry. These are: plate-and-frame, spiral wound, tubular and hollow-fiber configurations.
  131. [131]
    Comparative Assessment of Tubular Ceramic, Spiral Wound ... - NIH
    Mar 24, 2021 · Polymeric hollow fiber membranes (HFM) generally combine the advantages of both types of modules, including a large membrane area per module, ...
  132. [132]
    (PDF) Comparative Assessment of Tubular Ceramic, Spiral Wound ...
    Oct 15, 2025 · Comparative Assessment of Tubular Ceramic, Spiral Wound, and Hollow Fiber Membrane Microfiltration Module Systems for Milk Protein Fractionation.
  133. [133]
    State-of-the-Art Organic- and Inorganic-Based Hollow Fiber ... - NIH
    The history of hollow fiber membranes was traced back to the 1960s, when the first hollow fiber membrane was developed for the reverse osmosis (RO) process [13] ...
  134. [134]
    Brief history of water treatment technology | Reverse Osmosis
    Although the phenomenon of infiltration occurs every day in our bodies, it was not until 1748 that French physicist Jean Antoine nollet first discovered it from ...
  135. [135]
    Review of Hollow Fiber Membranes for Gas Separation - MDPI
    Hollow fiber membranes (HFMs) provide advantages such as self-supported, compact design, and a higher surface area-to-volume ratio compared to flat sheet ...
  136. [136]
    Performance of spiral wound ultrafiltration membrane module for ...
    In membrane operations, different membrane modules are available for clarification of juice such as tubular [3], hollow fibre [4], flat sheet [5] and spiral ...
  137. [137]
    On the Fluid Mechanics of Spiral-Wound Membrane Modules
    Aug 5, 2025 · Currently available membrane modules include spiral-wound, hollow-fiber, tubular, and plate and frame elements, employed in reverse osmosis ...
  138. [138]
    Hydrodynamic Techniques to Enhance Membrane Filtration
    Abstract. This article reviews the use of various techniques for membrane filtration, such as Dean and Taylor vortices, pulsatile flows, ...
  139. [139]
    Dead-End Filtration vs. Cross-Flow Filtration — All You Need to Know
    Cross-flow filtration moves fluid tangentially, while dead-end filtration moves it vertically, forming a cake that clogs the filter. Dead-end filtration is ...Missing: technology Reynolds number polarization empirical data
  140. [140]
    [PDF] Dead-end and crossflow ultrafiltration process modelling
    The recirculation of retentate in crossflow filtration may induce a tangential shear velocity which prevents membrane fouling, thus promoting a longer ...
  141. [141]
    Cross Flow Membrane Operations - Synder Filtration
    Cross flow filtration is when the flow is applied tangentially across the membrane surface vs. dead-end which is applied perpendicular to the membrane.
  142. [142]
    Dimensionless numbers for the analysis of air sparging aimed to ...
    This paper introduces a new approach: dividing the fouling number by the Reynolds number, which basically corresponds to the ratio of the Schmidt and Sherwood ...Missing: dynamics | Show results with:dynamics<|control11|><|separator|>
  143. [143]
    [PDF] dead-end and crossflow microfiltration of yeast and bentonite ...
    Fouling of the membrane causes a decrease in filtration efficiency and can result in loss of product. Fouling may be caused by plugging of membrane pores by.<|separator|>
  144. [144]
    Comparison between dead-end (DEF), cross-flow (CF) and dynamic ...
    In cross-flow filtration, the aim is to reduce the impact of fouling. ... cross-flow membrane since fouling and/or concentration polarization are reduced.
  145. [145]
    [PDF] Emerging Technologies for High Recovery Processing
    Company Information: The Hyrec Membrane Concentrator is a cascade RO system with feed to an intermediate stage. The intent of the system was to replace the ...
  146. [146]
    Enhancing energy efficiency and water recovery in two-stage ...
    Aug 11, 2025 · Two-stage reverse osmosis (2SRO) has demonstrated superior energy efficiency and higher clean water productivity compared to single-stage ...
  147. [147]
    Schmidt Number - an overview | ScienceDirect Topics
    The Schmidt number is a dimensionless quantity that relates the viscosity of a fluid to its diffusion coefficient, correlating momentum transport with mass ...
  148. [148]
    Modeling of Cross-Flow Filtration Processes in an Airlift Ceramic ...
    (3-6) The formation of gas−liquid two-phase cross-flow on the membrane surface by air sparging has been shown to reduce concentration polarization and fouling ...
  149. [149]
    How Reverse Osmosis Membranes Work - AXEON Water
    Engineered with stable operation in mind, RO system components like housings and pressurizing pumps prioritize efficient water drainage and can accommodate up ...
  150. [150]
    Membrane Filtration
    Most of our designs include the required pretreatment systems (equipment and chemical) in order to reduce or eliminate fouling and scaling of the membranes.
  151. [151]
    Membrane fouling monitoring by 3ω sensing | Scientific Reports
    Sep 14, 2023 · Currently, membrane fouling monitoring in full scale is mainly monitored by transmembrane pressure measurement (TMP) and flux measurements.
  152. [152]
    Experimental evaluation of a multi-skid reverse osmosis unit ...
    Nov 15, 2016 · A reverse osmosis (RO) desalination unit, composed of three identical skids connected in parallel and using an energy recovery system, ...
  153. [153]
    Modular Water Treatment Systems | IDE Technologies
    Sep 11, 2025 · Discover how modular water treatment systems cut project costs, reduce timelines, and deliver high-quality, sustainable results.
  154. [154]
    Energy Recovery in SWRO Desalination: Current Status and New ...
    Apr 2, 2020 · The PX is most effective at recovery rates up to 50% (Urrea et al., 2019). Energy savings with ERDs can be in the range of 25–40% compared with ...Abstract · Introduction · Energy Recovery Devices · Hybrid Process Configurations
  155. [155]
    Prediction of Permeate Flux in Ultrafiltration Processes - NIH
    In any membrane filtration, the prediction of permeate flux is critical to calculate the membrane surface required, which is an essential parameter for ...
  156. [156]
    Model-based optimization approaches for pressure-driven ...
    Jun 15, 2023 · In this review, we present a unified four-phase model-based optimization framework to optimize these systems, using mechanistic models, empirical models ...
  157. [157]
    Salt and Water Transport in Reverse Osmosis Membranes
    ... Solution-Diffusion Model - a Review. J. Membr. Sci. 1995, 107, 1– 21, DOI: 10.1016/0376-7388(95)00102-I. Google Scholar. 13. The solution-diffusion model: a ...
  158. [158]
    [PDF] Membranes II Model Equations
    We define characteristic dimensionless parameters for a membrane separation : Selectivity: Pressure ratio: Stage cut: Recovery: Purity: Given the operating ...Missing: governing rejection
  159. [159]
    Universal solution to the membrane selectivity challenge
    The target of selectivity of a membrane can range from solid particles, dissolved ions, organic molecules, or pure solvents to gases. We defined a membrane ...
  160. [160]
    Insights into the silica scaling behaviors in membrane distillation ...
    Sep 1, 2024 · Silica scaling imposes a significant limitation on the efficacy of membrane distillation (MD) in the treatment of hypersaline wastewater.
  161. [161]
    Inorganic and Organic Silicon Fouling of Nanofiltration Membranes ...
    Jun 26, 2023 · We additionally provided compelling evidence that silicon deposition was accompanied by mineral scaling (Ca, Al, Mg, and P) and the accumulation ...
  162. [162]
    Potential Pitfalls in Membrane Fouling Evaluation: Merits of Data ...
    We provide a couple of examples to prove that it is possible to obtain misleading conclusions from commonly used representations of fouling data.
  163. [163]
    A Comprehensive Review on Membrane Fouling - MDPI
    The findings showed that an optimized steady-state flux could be obtained by lowering the value of A2 and increasing that of A1. Moreover, a higher feed CFV ...2.2. Organic Fouling · 2.4. Biofouling · 3.1. Pilot Plant Evaluation...
  164. [164]
    Microscopy and Spectroscopy Techniques for Characterization ... - NIH
    This paper reviews common spectroscopy techniques for the characterization of polymeric membranes. The techniques usually study the morphology data, crystal ...
  165. [165]
    [PDF] Overview of main techniques used for membrane characterization
    Dec 4, 2014 · This review is motivated by the improvement of researchers' ability to evaluate the role(s) of membrane properties in determining membrane.
  166. [166]
    Microscopy and Spectroscopy Techniques for Characterization of ...
    For membrane morphology, scanning electron microscopy (SEM), transmission electron microscopy (TEM), and atomic force microscopy (AFM) are widely implemented.
  167. [167]
    Sample preparation matters: Scanning electron microscopic ...
    SEM analysis provides invaluable insights into the surface morphology and cross-sectional architecture of membranes, which are critical determinants of their ...
  168. [168]
    F316 Standard Test Methods for Pore Size Characteristics of ... - ASTM
    Dec 13, 2019 · These test methods cover the determination of two of the pore size properties of membrane filters with maximum pore sizes from 0.1 to 15.0 μm.
  169. [169]
    Bubblepoint - an overview | ScienceDirect Topics
    Bubble point is used to characterise the pore size of MF membranes.18,21 It is also used to test the integrity of membrane assembly including membrane seals.
  170. [170]
    An Efficient Method to Determine Membrane Molecular Weight Cut ...
    Oct 1, 2020 · A membrane's MWCO is defined as the minimum molecular weight of a solute that is 90% retained by the membrane, according to the French Standard ...
  171. [171]
    (PDF) Molecular weight cut-off determination of pressure filtration ...
    Aug 7, 2025 · In this work, a newly modified colorimetric method was established to conduct a simple measurement for a membrane's MWCO.
  172. [172]
    Multi-lab study on the pure-gas permeation of commercial ...
    Oct 5, 2022 · In this multi-lab study, ten independent laboratories collected high-pressure pure-gas permeation data for H2, O2, CH4, and N2 in commercial ...
  173. [173]
    [PDF] Review on Testing Methods for Permeability and Selectivity ...
    This paper gives general guidelines on various methods for determining the membrane permeability and selectivity with the commonly used setup. The paper also.
  174. [174]
    Streaming potential measurements as a characterization method for ...
    Streaming potential measurements have been carried out to determine the zeta potential and the kinetic surface charge density of the membranes.Missing: zeta | Show results with:zeta
  175. [175]
    Zeta potential of ion-conductive membranes by streaming current ...
    During the zeta potential characterization of a series of ion-conductive sulfonated poly(sulfone) membranes, it was found that the measured streaming current ...
  176. [176]
    [PDF] Porosimetric membrane characterization techniques: a review
    Indirect membrane pore size characterization techniques reviewed in this study. Theoretical Basis. Technique. Various Embodiments. Young Laplace Equation.
  177. [177]
    Fouling mitigation strategies for different foulants in membrane ...
    Membrane fouling that is observed in MD can be classified according to the source of feed water: biological fouling, colloidal fouling, organic fouling, and ...
  178. [178]
    Fouling Layer - an overview | ScienceDirect Topics
    Membrane fouling occurs when organic, inorganic materials, suspended and colloidal particles, or microorganisms accumulate, adsorb, or attach to a semipermeable ...
  179. [179]
    Interaction mechanisms associated with organic colloid fouling of ...
    In this paper we have therefore used XDLVO theory to investigate initial UF membrane fouling caused by NOM, and to determine the predominant factors that ...
  180. [180]
    Quantitative Analysis of Membrane Fouling Mechanisms Involved in ...
    Sep 22, 2018 · However, to date, the xDLVO theory is yet to be applied for the quantitative analysis of membrane fouling mechanisms involved in MF of humic ...
  181. [181]
    Interaction Mechanisms and Predictions of the Biofouling of Polymer ...
    The XDLVO model was found to predict better the AFM colloidal probe adhesion data and the QCM-D adsorption behavior of BSA onto the polymer films than the DLVO ...
  182. [182]
    Outlining the Roles of Membrane-Foulant and Foulant-Foulant ... - NIH
    In this review, we outline the roles of membrane-foulant and foulant-foulant intermolecular interactions in MF/UF organic fouling.
  183. [183]
    Interaction mechanisms and predictions on membrane fouling in an ...
    Aug 10, 2025 · The Extended Derjaguin–Landau–Verwey–Overbeek (XDLVO) approach was used to predict the mechanisms regarding the influence of humic acid together ...
  184. [184]
    Fouling control strategies for direct membrane ultrafiltration
    May 15, 2022 · The current study addresses a physical cleaning strategy based on membrane module rotation for DMF. The effect of rotational speed (0 – 340 rpm) and relaxation ...Missing: efficacy | Show results with:efficacy
  185. [185]
    Fouling and Chemical Cleaning of Microfiltration Membranes - NIH
    Mar 10, 2021 · Biofouling decreases membrane efficiency by lowering precise membrane flux [20]. The physical cleaning process is not enough to remove ...Missing: loss | Show results with:loss
  186. [186]
    Evaluation of different cleaning strategies for different types of ...
    Feb 15, 2020 · Simple physical cleanings (i.e., hydraulic flushing and osmotic backwashing) are firstly employed to remove membrane fouling. Both methods ...Missing: efficacy | Show results with:efficacy
  187. [187]
    [PDF] Flux Decline in Ultrafiltration Processes - https ://ris.utwen te.nl
    The flux decline phenomena can be generally divided in fouling (irreversible and long term phenomena) and concentration polarization (reversible and directly ...<|separator|>
  188. [188]
    Cost of fouling in full-scale reverse osmosis and nanofiltration ...
    Mar 15, 2021 · We found that the cost of fouling in the RO plants was around 24% of OPEX, while the fouling related costs in NF cases was only around 11% due to the low ...
  189. [189]
    Global energy, costs, and emissions from reverse osmosis ...
    Oct 16, 2025 · We quantify the energy, cost, and emissions trade-offs of reverse osmosis desalination under various climate scenarios and salinity levels.
  190. [190]
    Challenges in capital and operation cost reduction of reverse ...
    The SWRO plants with one type of pressure exchanger consume 3.5–4.6 kWh/m3, while other pressure exchangers exhibit energy usage 3.0–5.3 kWh/m3 at different ...
  191. [191]
    Low energy consumption in the Perth seawater desalination plant
    The plant consumes less than 4.2 kilowatt hours per cubic meter including intake, pretreatment, both RO passes, post-treatment, potable water pumping and all ...
  192. [192]
    UF/MF as RO pre-treatment: the real benefit - ScienceDirect.com
    Mar 1, 2008 · Ultrafiltration (UF) and microfiltration (MF) are theoretically the best pre-treatment upstream reverse osmosis, removing from the feed water most of the ...
  193. [193]
    Evaluating the efficiency of different microfiltration and ultrafiltration ...
    Evaluating the efficiency of different microfiltration and ultrafiltration membranes used as pretreatment for Red Sea water reverse osmosis desalination.
  194. [194]
    [PDF] Ultrafiltration as a pretreatment for seawater desalination: A review
    To tackle this problem, ultrafiltration (UF) membrane is employed prior to SWRO membrane and acts as pretreatment in reducing fouling, silt density index (SDI) ...
  195. [195]
    Reverse osmosis: Membranes help beat the drought - ScienceDirect
    The first-pass SWRO component has a specific energy consumption (SEC) of 2.4 kWh/m3, the lowest amount on record for a large-scale SWRO desalination plant.
  196. [196]
    How much energy does desalinisation use? Is it “absurdly cheap”?
    Sep 18, 2024 · Reverse osmosis uses around 2.5 to 3.5 kWh per cubic metre. Thermal technologies can use somewhere in the range of 13 kWh. 3. The average ...
  197. [197]
    Practical minimum energy use of seawater reverse osmosis
    The TM energy consumption band is based on Lin36 calculations, which provide a baseline to estimate the minimum amount of energy that depends on the feed water.<|separator|>
  198. [198]
    Characteristics of Desalination Brine and Its Impacts on Marine ...
    Apr 25, 2022 · Desalination brine disposal raises the salinity of seawater and introduces associated chemicals into the receiving water body, resulting in the ...<|control11|><|separator|>
  199. [199]
    Environmental impacts of desalination and brine treatment
    Desalination does have various environmental impacts: brine discharge, high energy consumption, GHGs emissions, intensive use of chemicals and water intake ...
  200. [200]
    [PDF] Life cycle assessment of reverse osmosis for high-salinity seawater ...
    This study aims to assess the life cycle environmental impact of SWRO desalination for potable and industrial water production. This objective can provide a ...
  201. [201]
    Environmental Impacts of the Brine Disposal and Their Control
    This review examines the M&CSs related to the several environmental impacts (EIs) of desalination engineering and focuses on brine disposal.
  202. [202]
    Membrane Gas‐Separation: Applications - ResearchGate
    Glassy polymers including polyimide (PI), cellulose acetate (CA), and polysulfone (PSF) are traditional gas separation membranes that have been widely used over ...
  203. [203]
    Module permeance of carbon dioxide, nitrogen, and oxygen in Air...
    The permeances of CO 2 and CH 4 , pure and mixed, in the PRISM PA1020-P1 Air Products module with a polysulfone-based membrane, were determined in our previous ...
  204. [204]
    Technoeconomic analysis of oxygen-nitrogen separation for oxygen ...
    Aug 1, 2021 · These methods are generally less energy intensive than cryogenic distillation and are best operated at mid-scale (20–100 tons/day).
  205. [205]
    [PDF] A review of air separation technologies and their integration with ...
    Cryogenic air separation is currently the most efficient and cost-effective technology for producing large quantities of oxygen, nitrogen, and argon as gaseous ...
  206. [206]
    [PDF] Polysulfone membranes for CO2/CH4 separation: State of the art
    Polysulfone pure- and mixed-gas permeation properties have been extensively explored for gas separation due to its low price, chemical stability, and mechanical ...
  207. [207]
    Membrane-Based Technologies for Post-Combustion CO2 Capture ...
    Dec 2, 2023 · This study reviews post-combustion CO 2 capture technologies and the latest progress in membrane processes for CO 2 separation.
  208. [208]
    Recent Advances of Polymeric Membranes in Tackling Plasticization ...
    This review addresses the key challenges in the application of polymeric technology for CO2 separation, focusing on plasticization and aging. A brief ...Missing: separators | Show results with:separators
  209. [209]
    An empirical correlation of gas permeability and permselectivity in ...
    Sep 30, 2009 · This paper considers the current and future potential of polymeric membranes in acidic gas removal, heavy hydrocarbon recovery, water ...Missing: maxima | Show results with:maxima
  210. [210]
    Cu (II)-phenolic complex incorporated hemodialysis membranes for ...
    In the process of simulation dialysis, the PSf/PEI-TA/FeCu displayed above 90 % of urea removal rate, in comparison with 10 % of the pristine membrane.
  211. [211]
    Biocompatibility of Surface-Modified Membranes for Chronic ...
    Apr 3, 2022 · Membrane properties influence solute removal and the interactions with blood components that define the membrane's biocompatibility.
  212. [212]
    Clinical relevance of cytokine production in hemodialysis
    Bioincompatible dialytic treatment may induce an inappropriate monocyte activation and cytokine production, which, in turn, may mediate some of the immune and ...
  213. [213]
    Functionalized Hemodialysis Polysulfone Membranes with Improved ...
    After silanization, the adsorption capacity was improved, with the composite membrane being able to remove urea and also to filter larger lysozyme molecules.
  214. [214]
  215. [215]
    Control of Protein Particle Formation During Ultrafiltration ...
    In bench-scale UF/DF concentration experiments, approximately 5 g mAb was concentrated from roughly 5 to 40 mg/mL, diafiltration was conducted at 40 mg/mL for ...
  216. [216]
    Membrane-based continuous fermentation with cell recycling ... - NIH
    Mar 4, 2023 · This study uses a membrane-based system to recycle cells for continuous propionic acid production from glycerol, achieving a productivity of 1. ...
  217. [217]
    Biodegradable Chitosan-Based Membranes for Highly Effective ...
    The study also demonstrated excellent antifouling properties of the modified membrane that could achieve near 100% water flux recovery. Introduction. Water ...Missing: bioprocessing | Show results with:bioprocessing
  218. [218]
    Case studies of clinical hemodialysis membranes - Nature
    Sep 9, 2020 · ... membranes induce more cytokines and inflammation factors, confirming the better biocompatibility of CTA. This decrease in activation with ...
  219. [219]
    Membrane filtration technologies for sustainable industrial ...
    Compared to organic membranes, they provide better separation efficiency, longer service life, and greater resistance to fouling. Since the 1990s, inorganic ...
  220. [220]
    Pilot-scale evaluation of forward osmosis membranes for volume ...
    Jun 1, 2022 · FO-RO has been applied for the treatment of raw produced water; the system achieved more than 99% rejection of almost all the ions, but fouling ...
  221. [221]
    [PDF] Field demonstration of intensified membrane distillation for treating ...
    In this article, we report the design and development of a pilot scale MD for desalination of real produced water resulting from unconventional oil and gas ...
  222. [222]
    Use of Membrane Technologies in Dairy Industry: An Overview - PMC
    Nov 11, 2021 · UF can be applied as a standardization process of milk; however, the breakthrough use of UF was to convert milk whey into refined proteins for ...Missing: m3/ | Show results with:m3/
  223. [223]
    Whey protein concentrate production by continuous ultrafiltration
    Mar 1, 2007 · The operability studies demonstrate that a standard 12 loop commercial ultrafiltration plant is able to produce the desired specifications of the concentrates.Missing: m3/ | Show results with:m3/
  224. [224]
    Unlocking the Benefits of Hybrid and Standalone Pervaporation for ...
    Jul 26, 2025 · Advantages of HybSi® AR membranes include superior chemical and thermal stability, and up to 50% potential savings in energy. The reader can ...
  225. [225]
    [PDF] Semi-pilot Tests of Ethanol Dehydration using Commercial Ceramic ...
    In this regard, previous technoeconomic studies have shown that a hybrid distillation-pervaporation schemes are superior in terms of energy efficiency and more ...<|separator|>
  226. [226]
    Membrane Fouling: Common Causes, Types, and Remediation
    Oct 12, 2022 · Membrane fouling occurs when contaminants deposit on the membrane surface or feed spacer, limiting liquid flow. Main causes include suspended ...
  227. [227]
    Reverse osmosis membrane compaction and embossing at ultra ...
    Sep 1, 2022 · Davenport et al. [12] observed a 35% permeability decrease and 60% cross-sectional thickness reduction after compacting seawater RO membranes ...Missing: percentage | Show results with:percentage
  228. [228]
    [PDF] FilmTec™ Reverse Osmosis Membranes Technical Manual - DuPont
    Sep 18, 2025 · The only DuPont-approved and valid version of this document is the most current English version provided by DuPont at the time of sale.<|separator|>
  229. [229]
    Microbial removal and integrity monitoring of ro and NF Membranes
    Dec 1, 2003 · These removals indicated that <1% of the permeate flux passed through the pinhole of the compromised element and that changes in conductivity ...
  230. [230]
    Impact of intermittency on membrane integrity - ScienceDirect.com
    Jul 1, 2024 · The second reason is associated with defects on the membrane surface, such as the formation of pinholes (breaches), increase in pore size, and ...
  231. [231]
  232. [232]
    Membrane ageing in full-scale water treatment plants - ScienceDirect
    Feb 1, 2020 · Impacts of ultrafiltration membrane ageing are observed after 5 years of operation. Ageing is evident in both membrane performance and chemical characteristics.
  233. [233]
    Evaluation of the Specific Energy Consumption of Sea Water ... - NIH
    Apr 16, 2022 · Reverse osmosis (RO) is currently the most common membrane process for desalination since high-performance RO membranes and modules have been ...
  234. [234]
    Seawater RO Operating Costs Analysis: A Comprehensive Guide
    Aug 28, 2024 · Interested in understanding seawater RO operating costs analysis and its implications? This in-depth guide is for you. Contact us today!
  235. [235]
    (PDF) Review: Water Desalination Cost - ResearchGate
    Jan 27, 2023 · The reverse osmosis (RO) membrane technology is considered the lowest operating cost in terms of energy relative to the multi-stage flash evaporation (MSF) ...
  236. [236]
    [PDF] Seawater Desalination Costs - WateReuse Association
    By comparison, the range for brackish water membrane desalinating processes (BWRO) is US $0.40/kgal to $4.00/kgal.
  237. [237]
    [PDF] Trends and future outlooks in circularity of desalination membrane ...
    Apr 6, 2023 · Presently, nearly 840,000 RO modules (>14,000 tonnes/year waste of membrane materials) are discarded every year worldwide and by 2025, this ...
  238. [238]
    [PDF] The environmental impacts of plastics and micro-plastics use, waste ...
    Most of the plastic waste ends up in landfills and, ultimately, in the environment. Most plastics do not degrade. Instead, they slowly fragment into smaller ...
  239. [239]
    Recycling of end-of-life polymeric membranes for water treatment
    This brief frontier review aims to serve as a valuable reference for recycling end-of-life water treatment polymeric membranes.
  240. [240]
    Long-term, basin-scale salinity impacts from desalination in ... - Nature
    Nov 29, 2022 · Local impacts include the impingement and entrainment of marine organisms at the water intake and the discharge of heated, hypersaline, ...
  241. [241]
    The perils of building big: Desalination sustainability and brine ...
    Sustainable desalination requires addressing environmental impacts including damage to ecosystems from the high volumes of brine in the Gulf. This paper ...Missing: hypersalinity | Show results with:hypersalinity
  242. [242]
    Impact of Brine Discharge from Seawater Desalination Plants on ...
    Sep 28, 2019 · Because energy for desalination increases with seawater salinity, buildup of salt in brine endangers potable water supply there. Brine also ...Missing: harm hypersalinity
  243. [243]
    Carbon Footprint of Seawater Desalination Technologies: A Review
    May 3, 2024 · Beery and Repke [47] reported a CFw range of 1.4–4.32 kg CO2e/m3 for the SWRO scenarios based on fossil fuel electricity and 0.09–0.12 kg CO2e/m ...
  244. [244]
    Study of the Ecological Footprint and Carbon Footprint in a Reverse ...
    Water desalination in the Canary Islands requires more than 770,000 MWh, equivalent to more than 180,000 tons of fossil fuels each year (3600 barrels per day), ...
  245. [245]
    Comparing the environmental impacts of using bio-renewable and ...
    Bio-renewable solvents have lower production impacts, but membrane fabrication using them has higher environmental impact due to lower yield.
  246. [246]
    Sustainable Polymeric Membranes: Green Chemistry and Circular ...
    Jul 3, 2025 · (6) The typical lifespan of RO and NF membranes ranges from 3 to 7 years, after which they are often disposed of via landfills or incineration.
  247. [247]
    Bio-Based Polymeric Membranes: Development and Environmental ...
    Bio-based polymeric membranes emerge as a greener alternative to counteract the environmental problems caused by synthetic materials. This membrane type is ...Missing: limitations | Show results with:limitations
  248. [248]
    Recent advances in membrane distillation hybrids for energy ...
    This work presents a critical review on recent advances in membrane distillation hybrids and their technological developments towards resource recovery.
  249. [249]
    Machine Learning in Membrane Design: From Property Prediction to ...
    Mar 4, 2024 · By using ML models for rapid screening and optimization, researchers can efficiently evaluate a wide array of membranes, substantially reducing ...
  250. [250]
    Graphene oxide-based membranes for water desalination and ...
    Mar 27, 2024 · This review, while studying these challenges, aims to outline the most recent ground-breaking developments in graphene-based membrane technology.Missing: breakthroughs | Show results with:breakthroughs
  251. [251]
    Engineering High‐Flux 2D Separation Membranes
    Jul 29, 2025 · Last, we propose the future research directions of 2D lamellar membranes from the aspects of material design, stability optimization, scalable ...
  252. [252]
    Anti-fouling membranes for membrane distillation: A critical review of ...
    Oct 18, 2025 · Reviews advanced strategies: bioinspired, Janus, and stimuli-responsive antifouling membranes. Highlights dynamic interfaces for self-cleaning ...
  253. [253]
    Chelation-directed interface engineering of in-place self-cleaning ...
    Mar 4, 2024 · Advanced self-cleaning antifouling membranes would represent a breakthrough in this sector. Herein, a chelation-directed interface ...<|separator|>
  254. [254]
    A review on hybrid membrane-adsorption systems for intensified ...
    Jun 1, 2023 · This review introduces the state-of-the-art on the capabilities of the combined adsorption-membrane filtration systems for water and wastewater treatment ...Missing: prospects | Show results with:prospects
  255. [255]
    Challenges and Perspectives on Photocatalytic Membrane Reactors ...
    Apr 17, 2025 · Photocatalytic membrane reactors (PMRs) have emerged as a promising technology for air purification due to their ability to integrate ...1 Introduction · 2 Photocatalytic Membranes · 3 Photocatalytic Membrane...
  256. [256]
    Hybrid adsorption–photocatalysis composites: a sustainable route ...
    Jul 11, 2025 · This review presents a comprehensive analysis of recent advances in adsorption–photocatalysis hybrid materials, focusing on the synergistic ...Photocatalytic Degradation · Carbon-Based Composites · Metal Oxide Composites