Membrane gas separation

Last updated

Gas mixtures can be effectively separated by synthetic membranes made from polymers such as polyamide or cellulose acetate, or from ceramic materials. [1]

Contents

Membrane cartridge Flux distribution inside the fiber.jpg
Membrane cartridge

While polymeric membranes are economical and technologically useful, they are bounded by their performance, known as the Robeson limit (permeability must be sacrificed for selectivity and vice versa). [2] This limit affects polymeric membrane use for CO2 separation from flue gas streams, since mass transport becomes limiting and CO2 separation becomes very expensive due to low permeabilities. Membrane materials have expanded into the realm of silica, zeolites, metal-organic frameworks, and perovskites due to their strong thermal and chemical resistance as well as high tunability (ability to be modified and functionalized), leading to increased permeability and selectivity. Membranes can be used for separating gas mixtures where they act as a permeable barrier through which different compounds move across at different rates or not move at all. The membranes can be nanoporous, polymer, etc. and the gas molecules penetrate according to their size, diffusivity, or solubility.

Basic process

Gas separation across a membrane is a pressure-driven process, where the driving force is the difference in pressure between inlet of raw material and outlet of product. The membrane used in the process is a generally non-porous layer, so there will not be a severe leakage of gas through the membrane. The performance of the membrane depends on permeability and selectivity. Permeability is affected by the penetrant size. Larger gas molecules have a lower diffusion coefficient. The polymer chain flexibility and free volume in the polymer of the membrane material influence the diffusion coefficient, as the space within the permeable membrane must be large enough for the gas molecules to diffuse across. The solubility is expressed as the ratio of the concentration of the gas in the polymer to the pressure of the gas in contact with it. Permeability is the ability of the membrane to allow the permeating gas to diffuse through the material of the membrane as a consequence of the pressure difference over the membrane, and can be measured in terms of the permeate flow rate, membrane thickness and area and the pressure difference across the membrane. The selectivity of a membrane is a measure of the ratio of permeability of the relevant gases for the membrane. It can be calculated as the ratio of permeability of two gases in binary separation. [3]

The membrane gas separation equipment typically pumps gas into the membrane module and the targeted gases are separated based on difference in diffusivity and solubility. For example, oxygen will be separated from the ambient air and collected at the upstream side, and nitrogen at the downstream side. As of 2016, membrane technology was reported as capable of producing 10 to 25 tonnes of 25 to 40% oxygen per day. [3]

Membrane governing methodology

(a) Bulk flow through pores; (b) Knudsen diffusion through pores; (c) molecular sieving; (d) solution diffusion through dense membranes. Mechanisms of transport in membranes.jpg
(a) Bulk flow through pores; (b) Knudsen diffusion through pores; (c) molecular sieving; (d) solution diffusion through dense membranes.

There are three main diffusion mechanisms. The first (b), Knudsen diffusion holds at very low pressures where lighter molecules can move across a membrane faster than heavy ones, in a material with reasonably large pores. [4] The second (c), molecular sieving, is the case where the pores of the membrane are too small to let one component pass, a process which is typically not practical in gas applications, as the molecules are too small to design relevant pores. In these cases the movement of molecules is best described by pressure-driven convective flow through capillaries, which is quantified by Darcy's law. However, the more general model in gas applications is the solution-diffusion (d) where particles are first dissolved onto the membrane and then diffuse through it both at different rates. This model is employed when the pores in the polymer membrane appear and disappear faster relative to the movement of the particles. [5]

In a typical membrane system the incoming feed stream is separated into two components: permeant and retentate. Permeant is the gas that travels across the membrane and the retentate is what is left of the feed. On both sides of the membrane, a gradient of chemical potential is maintained by a pressure difference which is the driving force for the gas molecules to pass through. The ease of transport of each species is quantified by the permeability, Pi. With the assumptions of ideal mixing on both sides of the membrane, ideal gas law, constant diffusion coefficient and Henry's law, the flux of a species can be related to the pressure difference by Fick's law: [4]

where, (Ji) is the molar flux of species i across the membrane, (l) is membrane thickness, (Pi) is permeability of species i, (Di) is diffusivity, (Ki) is the Henry coefficient, and (pi') and (pi") represent the partial pressures of the species i at the feed and permeant side respectively. The product of DiKi is often expressed as the permeability of the species i, on the specific membrane being used.

The flow of a second species, j, can be defined as:

A simplified design schematic of a membrane separation process Membrane separation process.jpg
A simplified design schematic of a membrane separation process

With the expression above, a membrane system for a binary mixture can be sufficiently defined. it can be seen that the total flow across the membrane is strongly dependent on the relation between the feed and permeate pressures. The ratio of feed pressure (p') over permeate pressure (p") is defined as the membrane pressure ratio (θ).

It is clear from the above, that a flow of species i or j across the membrane can only occur when:

In other words, the membrane will experience flow across it when there exists a concentration gradient between feed and permeate. If the gradient is positive, the flow will go from the feed to the permeate and species i will be separated from the feed.

Therefore, the maximum separation of species i results from:

Another important coefficient when choosing the optimum membrane for a separation process is the membrane selectivity αij defined as the ratio of permeability of species i with relation to the species j.

This coefficient is used to indicate the level to which the membrane is able to separates species i from j. It is obvious from the expression above, that a membrane selectivity of 1 indicates the membrane has no potential to separate the two gases, the reason being, both gases will diffuse equally through the membrane.

In the design of a separation process, normally the pressure ratio and the membrane selectivity are prescribed by the pressures of the system and the permeability of the membrane . The level of separation achieved by the membrane (concentration of the species to be separated) needs to be evaluated based on the aforementioned design parameters in order to evaluate the cost-effectiveness of the system.

Membrane performance

The concentration of species i and j across the membrane can be evaluated based on their respective diffusion flows across it.

In the case of a binary mixture, the concentration of species i across the membrane:

This can be further expanded to obtain an expression of the form:

Using the relations:

The expression can be rewritten as:

Then using

[6]

The solution to the above quadratic expression can be expressed as:

Finally, an expression for the permeant concentration is obtained by the following:

Along the separation unit, the feed concentration decays with the diffusion across the membrane causing the concentration at the membrane to drop accordingly. As a result, the total permeant flow (q"out) results from the integration of the diffusion flow across the membrane from the feed inlet (q'in) to feed outlet (q'out). A mass balance across a differential length of the separation unit is therefore:

where:

Because of the binary nature of the mixture, only one species needs to be evaluated. Prescribing a function n'i=n'i(x), the species balance can be rewritten as:

Where:

Lastly, the area required per unit membrane length can be obtained by the following expression:

Membrane materials for carbon capture in flue gas streams

The material of the membrane plays an important role in its ability to provide the desired performance characteristics. It is optimal to have a membrane with a high permeability and sufficient selectivity and it is also important to match the membrane properties to that of the system operating conditions (for example pressures and gas composition).

Synthetic membranes are made from a variety of polymers including polyethylene, polyamides, polyimides, cellulose acetate, polysulphone and polydimethylsiloxane. [7]

Polymer membranes

Polymeric membranes are a common option for use in the capture of CO2 from flue gas because of the maturity of the technology in a variety of industries, namely petrochemicals. The ideal polymer membrane has both a high selectivity and permeability. Polymer membranes are examples of systems that are dominated by the solution-diffusion mechanism. The membrane is considered to have holes which the gas can dissolve (solubility) and the molecules can move from one cavity to the other (diffusion). [4]

It was discovered by Robeson in the early 1990s that polymers with a high selectivity have a low permeability and opposite is true; materials with a low selectivity have a high permeability. This is best illustrated in a Robeson plot where the selectivity is plotted as a function of the CO2 permeation. In this plot, the upper bound of selectivity is approximately a linear function of the permeability. It was found that the solubility in polymers is mostly constant but the diffusion coefficients vary significantly and this is where the engineering of the material occurs. Somewhat intuitively, the materials with the highest diffusion coefficients have a more open pore structure, thus losing selectivity. [8] [9] There are two methods that researchers are using to break the Robeson limit, one of these is the use of glassy polymers whose phase transition and changes in mechanical properties make it appear that the material is absorbing molecules and thus surpasses the upper limit. The second method of pushing the boundaries of the Robeson limit is by the facilitated transport method. As previously stated, the solubility of polymers is typically fairly constant but the facilitated transport method uses a chemical reaction to enhance the permeability of one component without changing the selectivity. [10]

Nanoporous membranes

Microscopic model of a nanoporous membrane. The white open area represents the area the molecule can pass through and the dark blue areas represent the membrane walls. The membrane channels consists of cavities and windows. The energy of the molecules in the cavity is Uc and the energy of a particle in the window is Uw. Microscopic model of a nanoporous membrane.jpg
Microscopic model of a nanoporous membrane. The white open area represents the area the molecule can pass through and the dark blue areas represent the membrane walls. The membrane channels consists of cavities and windows. The energy of the molecules in the cavity is Uc and the energy of a particle in the window is Uw.

Nanoporous membranes are fundamentally different from polymer-based membranes in that their chemistry is different and that they do not follow the Robeson limit for a variety of reasons. The simplified figure of a nanoporous membrane shows a small portion of an example membrane structure with cavities and windows. The white portion represents the area where the molecule can move and the blue shaded areas represent the walls of the structure. In the engineering of these membranes, the size of the cavity (Lcy x Lcz) and window region (Lwy x Lwz) can be modified so that the desired permeation is achieved. It has been shown that the permeability of a membrane is the production of adsorption and diffusion. In low loading conditions, the adsorption can be computed by the Henry coefficient. [4]

If the assumption is made that the energy of a particle does not change when moving through this structure, only the entropy of the molecules changes based on the size of the openings. If we first consider changes the cavity geometry, the larger the cavity, the larger the entropy of the absorbed molecules which thus makes the Henry coefficient larger. For diffusion, an increase in entropy will lead to a decrease in free energy which in turn leads to a decrease in the diffusion coefficient. Conversely, changing the window geometry will primarily effect the diffusion of the molecules and not the Henry coefficient.

In summary, by using the above simplified analysis, it is possible to understand why the upper limit of the Robeson line does not hold for nanostructures. In the analysis, both the diffusion and Henry coefficients can be modified without affecting the permeability of the material which thus can exceed the upper limit for polymer membranes. [4]

Silica membranes

Silica membranes are mesoporous and can be made with high uniformity (the same structure throughout the membrane). The high porosity of these membranes gives them very high permeabilities. Synthesized membranes have smooth surfaces and can be modified on the surface to drastically improve selectivity. Functionalizing silica membrane surfaces with amine containing molecules (on the surface silanol groups) allows the membranes to separate CO2 from flue gas streams more effectively. [2] Surface functionalization (and thus chemistry) can be tuned to be more efficient for wet flue gas streams as compared to dry flue gas streams. [11] While previously, silica membranes were impractical due to their technical scalability and cost (they are very difficult to produce in an economical manner on a large scale), there have been demonstrations of a simple method of producing silica membranes on hollow polymeric supports. These demonstrations indicate that economical materials and methods can effectively separate CO2 and N2. [12] Ordered mesoporous silica membranes have shown considerable potential for surface modification that allows for ease of CO2 separation. Surface functionalization with amines leads to the reversible formation of carbamates (during CO2 flow), increasing CO2 selectivity significantly. [12]

Zeolite membranes

A typical zeolite. Thin layers of this crystalline zeolite structure can act as a membrane, since CO2 can adsorb inside of the pores. Zeolite-ZSM-5-vdW.png
A typical zeolite. Thin layers of this crystalline zeolite structure can act as a membrane, since CO2 can adsorb inside of the pores.

Zeolites are crystalline aluminosilicates with a regular repeating structure of molecular-sized pores. Zeolite membranes selectively separate molecules based on pore size and polarity and are thus highly tunable to specific gas separation processes. In general, smaller molecules and those with stronger zeolite-adsorption properties are adsorbed onto zeolite membranes with larger selectivity. The capacity to discriminate based on both molecular size and adsorption affinity makes zeolite membranes an attractive candidate for CO2 separation from N2, CH4, and H2.

Scientists have found that the gas-phase enthalpy (heat) of adsorption on zeolites increases as follows: H2 < CH4 < N2 < CO2. [13] It is generally accepted that CO2 has the largest adsorption energy because it has the largest quadrupole moment, thereby increasing its affinity for charged or polar zeolite pores. At low temperatures, zeolite adsorption-capacity is large and the high concentration of adsorbed CO2 molecules blocks the flow of other gases. Therefore, at lower temperatures, CO2 selectively permeates through zeolite pores. Several recent research efforts have focused on developing new zeolite membranes that maximize the CO2 selectivity by taking advantage of the low-temperature blocking phenomena.

Researchers have synthesized Y-type (Si:Al>3) zeolite membranes which achieve room-temperature separation factors of 100 and 21 for CO2/N2 and CO2/CH4 mixtures respectively. [14] DDR-type and SAPO-34 membranes have also shown promise in separating CO2 and CH4 at a variety of pressures and feed compositions. [15] [16] The SAPO-34 membranes, being Nitrogen selective, are also strong contender for natural gas sweetening process. [17] [18] [19]

Researchers have also made an effort to utilize zeolite membranes for the separation of H2 from hydrocarbons. Hydrogen can be separated from larger hydrocarbons such as C4H10 with high selectivity. This is due to the molecular sieving effect since zeolites have pores much larger than H2, but smaller than these large hydrocarbons. Smaller hydrocarbons such as CH4, C2H6, and C3H8 are small enough to not be separated by molecular sieving. Researchers achieved a higher selectivity of Hydrogen when performing the separation at high temperatures, likely as a result of a decrease in the competitive adsorption effect. [20]

Metal-organic framework (MOF) membranes

There have been advances in zeolitic-imidazolate frameworks (ZIFs), a subclass of metal-organic frameworks (MOFs), that have allowed them to be useful for carbon dioxide separation from flue gas streams. Extensive modeling has been performed to demonstrate the value of using MOFs as membranes. [21] [22] MOF materials are adsorption-based, and thus can be tuned to achieve selectivity. [23] The drawback to MOF systems is stability in water and other compounds present in flue gas streams. Select materials, such as ZIF-8, have demonstrated stability in water and benzene, contents often present in flue gas mixtures. ZIF-8 can be synthesized as a membrane on a porous alumina support and has proven to be effective at separating CO2 from flue gas streams. At similar CO2/CH4 selectivity to Y-type zeolite membranes, ZIF-8 membranes achieve unprecedented CO2 permeance, two orders of magnitude above the previous standard. [24]

Structure of a perovskite. A membrane would consist of a thin layer of the perovskite structure. CMR1.PNG
Structure of a perovskite. A membrane would consist of a thin layer of the perovskite structure.

Perovskite membranes

Perovskite are mixed metal oxide with a well-defined cubic structure and a general formula of ABO3, where A is an alkaline earth or lanthanide element and B is a transition metal. These materials are attractive for CO2 separation because of the tunability of the metal sites as well as their stabilities at elevated temperatures.

The separation of CO2 from N2 was investigated with an α-alumina membrane impregnated with BaTiO3. [25] It was found that adsorption of CO2 was favorable at high temperatures due to an endothermic interaction between CO2 and the material, promoting mobile CO2 that enhanced CO2 adsorption-desorption rate and surface diffusion. The experimental separation factor of CO2 to N2 was found to be 1.1-1.2 at 100 °C to 500 °C, which is higher than the separation factor limit of 0.8 predicted by Knudsen diffusion. Though the separation factor was low due to pinholes observed in the membrane, this demonstrates the potential of perovskite materials in their selective surface chemistry for CO2 separation.

Other membrane technologies

In special cases other materials can be utilized; for example, palladium membranes permit transport solely of hydrogen. [26] In addition to palladium membranes (which are typically palladium silver alloys to stop embrittlement of the alloy at lower temperature) there is also a significant research effort looking into finding non-precious metal alternatives. Although slow kinetics of exchange on the surface of the membrane and tendency for the membranes to crack or disintegrate after a number of duty cycles or during cooling are problems yet to be fully solved. [27]

Construction

Membranes are typically contained in one of three modules: [7]

Uses

Membranes are employed in: [1]

Air separation

Oxygen-enriched air is in high demanded for a range of medical and industrial applications including chemical and combustion processes. Cryogenic distillation is the mature technology for commercial air separation for the production of large quantities of high purity oxygen and nitrogen. However, it is a complex process, is energy-intensive, and is generally not suitable for small-scale production. Pressure swing adsorption is also commonly used for air separation and can also produce high purity oxygen at medium production rates, but it still requires considerable space, high investment and high energy consumption. The membrane gas separation method is a relatively low environmental impact and sustainable process providing continuous production, simple operation, lower pressure/temperature requirements, and compact space requirements. [28] [3]

Current status of CO2 capture with membranes

A great deal of research has been undertaken to utilize membranes instead of absorption or adsorption for carbon capture from flue gas streams, however, no current[ when? ] projects exist that utilize membranes. Process engineering along with new developments in materials have shown that membranes have the greatest potential for low energy penalty and cost compared to competing technologies. [4] [10] [29]

Background

Today, membranes are used for commercial separations involving: N2 from air, H2 from ammonia in the Haber-Bosch process, natural gas purification, and tertiary-level enhanced oil recovery supply. [30]

Single-stage membrane operations involve a single membrane with one selectivity value. Single-stage membranes were first used in natural gas purification, separating CO2 from methane. [30] A disadvantage of single-stage membranes is the loss of product in the permeate due to the constraints imposed by the single selectivity value. Increasing the selectivity reduces the amount of product lost in the permeate, but comes at the cost of requiring a larger pressure difference to process an equivalent amount of a flue stream. In practice, the maximum pressure ratio economically possible is around 5:1. [31]

To combat the loss of product in the membrane permeate, engineers use “cascade processes” in which the permeate is recompressed and interfaced with additional, higher selectivity membranes. [30] The retentate streams can be recycled, which achieves a better yield of product.

Need for multi-stage process

Single-stage membranes devices are not feasible for obtaining a high concentration of separated material in the permeate stream. This is due to the pressure ratio limit that is economically unrealistic to exceed. Therefore, the use of multi-stage membranes is required to concentrate the permeate stream. The use of a second stage allows for less membrane area and power to be used. This is because of the higher concentration that passes the second stage, as well as the lower volume of gas for the pump to process. [31] [10] Other factors, such as adding another stage that uses air to concentrate the stream further reduces cost by increasing concentration within the feed stream. [10] Additional methods such as combining multiple types of separation methods allow for variation in creating economical process designs.

Membrane use in hybrid processes

Hybrid processes have long-standing history with gas separation. [32] Typically, membranes are integrated into already existing processes such that they can be retrofitted into already existing carbon capture systems.

MTR, Membrane Technology and Research Inc., and UT Austin have worked to create hybrid processes, utilizing both absorption and membranes, for CO2 capture. First, an absorption column using piperazine as a solvent absorbs about half the carbon dioxide in the flue gas, then the use of a membrane results in 90% capture. [33] A parallel setup is also, with the membrane and absorption processes occurring simultaneously. Generally, these processes are most effective when the highest content of carbon dioxide enters the amine absorption column. Incorporating hybrid design processes allows for retrofitting into fossil fuel power plants. [33]

Hybrid processes can also use cryogenic distillation and membranes. [34] For example, hydrogen and carbon dioxide can be separated, first using cryogenic gas separation, whereby most of the carbon dioxide exits first, then using a membrane process to separate the remaining carbon dioxide, after which it is recycled for further attempts at cryogenic separation. [34]

Cost analysis

Cost limits the pressure ratio in a membrane CO2 separation stage to a value of 5; higher pressure ratios eliminate any economic viability for CO2 capture using membrane processes. [10] [35] Recent studies have demonstrated that multi-stage CO2 capture/separation processes using membranes can be economically competitive with older and more common technologies such as amine-based absorption. [10] [34] Currently, both membrane and amine-based absorption processes can be designed to yield a 90% CO2 capture rate. [29] [10] [35] [36] [33] [34] For carbon capture at an average 600 MW coal-fired power plant, the cost of CO2 capture using amine-based absorption is in the $40–100 per ton of CO2 range, while the cost of CO2 capture using current membrane technology (including current process design schemes) is about $23 per ton of CO2. [10] Additionally, running an amine-based absorption process at an average 600 MW coal-fired power plant consumes about 30% of the energy generated by the power plant, while running a membrane process requires about 16% of the energy generated. [10] CO2 transport (e.g. to geologic sequestration sites, or to be used for EOR) costs about $2–5 per ton of CO2. [10] This cost is the same for all types of CO2 capture/separation processes such as membrane separation and absorption. [10] In terms of dollars per ton of captured CO2, the least expensive membrane processes being studied at this time are multi-step counter-current flow/sweep processes. [29] [10] [35] [36] [33] [34]

See also

Related Research Articles

<span class="mw-page-title-main">Fick's laws of diffusion</span> Mathematical descriptions of molecular diffusion

Fick's laws of diffusion describe diffusion and were first posited by Adolf Fick in 1855 on the basis of largely experimental results. They can be used to solve for the diffusion coefficient, D. Fick's first law can be used to derive his second law which in turn is identical to the diffusion equation.

<span class="mw-page-title-main">Haber process</span> Main process of ammonia production

The Haber process, also called the Haber–Bosch process, is the main industrial procedure for the production of ammonia. The German chemists Fritz Haber and Carl Bosch developed it in the first decade of the 20th century. The process converts atmospheric nitrogen (N2) to ammonia (NH3) by a reaction with hydrogen (H2) using a metal catalyst under high temperatures and pressures. This reaction is slightly exothermic (i.e. it releases energy), meaning that the reaction is favoured at lower temperatures and higher pressures. It decreases entropy, complicating the process. Hydrogen is produced via steam reforming, followed by an iterative closed cycle to react hydrogen with nitrogen to produce ammonia.

<span class="mw-page-title-main">Adsorption</span> Phenomenon of surface adhesion

Adsorption is the adhesion of atoms, ions or molecules from a gas, liquid or dissolved solid to a surface. This process creates a film of the adsorbate on the surface of the adsorbent. This process differs from absorption, in which a fluid is dissolved by or permeates a liquid or solid. Adsorption is a surface phenomenon and the adsorbate does not penetrate through the surface and into the bulk of the adsorbent, while absorption involves transfer of the absorbate into the volume of the material, although adsorption does often precede absorption. The term sorption encompasses both adsorption and absorption, and desorption is the reverse of sorption.

Darcy's law is an equation that describes the flow of a fluid through a porous medium. The law was formulated by Henry Darcy based on results of experiments on the flow of water through beds of sand, forming the basis of hydrogeology, a branch of earth sciences. It is analogous to Ohm's law in electrostatics, linearly relating the volume flow rate of the fluid to the hydraulic head difference via the hydraulic conductivity.

<span class="mw-page-title-main">Forward osmosis</span> Water purification process

Forward osmosis (FO) is an osmotic process that, like reverse osmosis (RO), uses a semi-permeable membrane to effect separation of water from dissolved solutes. The driving force for this separation is an osmotic pressure gradient, such that a "draw" solution of high concentration, is used to induce a net flow of water through the membrane into the draw solution, thus effectively separating the feed water from its solutes. In contrast, the reverse osmosis process uses hydraulic pressure as the driving force for separation, which serves to counteract the osmotic pressure gradient that would otherwise favor water flux from the permeate to the feed. Hence significantly more energy is required for reverse osmosis compared to forward osmosis.

An oxygen concentrator is a device that concentrates the oxygen from a gas supply by selectively removing nitrogen to supply an oxygen-enriched product gas stream. They are used industrially, to provide supplemental oxygen at high altitudes, and as medical devices for oxygen therapy.

<span class="mw-page-title-main">Pressure swing adsorption</span> Method of gases separation using selective adsorption under pressure

Pressure swing adsorption (PSA) is a technique used to separate some gas species from a mixture of gases under pressure according to the species' molecular characteristics and affinity for an adsorbent material. It operates at near-ambient temperature and significantly differs from the cryogenic distillation commonly used to separate gases. Selective adsorbent materials are used as trapping material, preferentially adsorbing the target gas species at high pressure. The process then swings to low pressure to desorb the adsorbed gas.

<span class="mw-page-title-main">Flory–Huggins solution theory</span> Lattice model of polymer solutions

Flory–Huggins solution theory is a lattice model of the thermodynamics of polymer solutions which takes account of the great dissimilarity in molecular sizes in adapting the usual expression for the entropy of mixing. The result is an equation for the Gibbs free energy change for mixing a polymer with a solvent. Although it makes simplifying assumptions, it generates useful results for interpreting experiments.

Barrer is a non-SI unit of gas permeability used in the membrane technology and contact lens industry. It is named after Richard Barrer.

In physics and engineering, permeation is the penetration of a permeate through a solid. It is directly related to the concentration gradient of the permeate, a material's intrinsic permeability, and the materials' mass diffusivity. Permeation is modeled by equations such as Fick's laws of diffusion, and can be measured using tools such as a minipermeameter.

<span class="mw-page-title-main">Membrane reactor</span>

A membrane reactor is a physical device that combines a chemical conversion process with a membrane separation process to add reactants or remove products of the reaction.

<span class="mw-page-title-main">Zeolitic imidazolate framework</span>

Zeolitic imidazolate frameworks (ZIFs) are a class of metal-organic frameworks (MOFs) that are topologically isomorphic with zeolites. ZIF glasses can be synthesized by the melt-quench method, and the first melt-quenched ZIF glass was firstly made and reported by Bennett et al. back in 2015. ZIFs are composed of tetrahedrally-coordinated transition metal ions connected by imidazolate linkers. Since the metal-imidazole-metal angle is similar to the 145° Si-O-Si angle in zeolites, ZIFs have zeolite-like topologies. As of 2010, 105 ZIF topologies have been reported in the literature. Due to their robust porosity, resistance to thermal changes, and chemical stability, ZIFs are being investigated for applications such as carbon dioxide capture.

<span class="mw-page-title-main">Diffusion</span> Transport of dissolved species from the highest to the lowest concentration region

Diffusion is the net movement of anything generally from a region of higher concentration to a region of lower concentration. Diffusion is driven by a gradient in Gibbs free energy or chemical potential. It is possible to diffuse "uphill" from a region of lower concentration to a region of higher concentration, as in spinodal decomposition. Diffusion is a stochastic process due to the inherent randomness of the diffusing entity and can be used to model many real-life stochastic scenarios. Therefore, diffusion and the corresponding mathematical models are used in several fields beyond physics, such as statistics, probability theory, information theory, neural networks, finance, and marketing.

A biogas upgrader is a facility that is used to concentrate the methane in biogas to natural gas standards. The system removes carbon dioxide, hydrogen sulphide, water and contaminants from the biogas. One technique for doing this uses amine gas treating. This purified biogas is also called biomethane. It can be used interchangeably with natural gas.

Nuclear magnetic resonance (NMR) in porous materials covers the application of using NMR as a tool to study the structure of porous media and various processes occurring in them. This technique allows the determination of characteristics such as the porosity and pore size distribution, the permeability, the water saturation, the wettability, etc.

<span class="mw-page-title-main">Membrane</span> Thin, film-like structure separating two fluids, acting as a selective barrier

A membrane is a selective barrier; it allows some things to pass through but stops others. Such things may be molecules, ions, or other small particles. Membranes can be generally classified into synthetic membranes and biological membranes. Biological membranes include cell membranes ; nuclear membranes, which cover a cell nucleus; and tissue membranes, such as mucosae and serosae. Synthetic membranes are made by humans for use in laboratories and industry.

Membrane technology encompasses the scientific processes used in the construction and application of membranes. Membranes are used to facilitate the transport or rejection of substances between mediums, and the mechanical separation of gas and liquid streams. In the simplest case, filtration is achieved when the pores of the membrane are smaller than the diameter of the undesired substance, such as a harmful microorganism. Membrane technology is commonly used in industries such as water treatment, chemical and metal processing, pharmaceuticals, biotechnology, the food industry, as well as the removal of environmental pollutants.

In fluid mechanics, fluid flow through porous media is the manner in which fluids behave when flowing through a porous medium, for example sponge or wood, or when filtering water using sand or another porous material. As commonly observed, some fluid flows through the media while some mass of the fluid is stored in the pores present in the media.

Diffusion-limited escape occurs when the rate of atmospheric escape to space is limited by the upward diffusion of escaping gases through the upper atmosphere, and not by escape mechanisms at the top of the atmosphere. The escape of any atmospheric gas can be diffusion-limited, but only diffusion-limited escape of hydrogen has been observed in our solar system, on Earth, Mars, Venus and Titan. Diffusion-limited hydrogen escape was likely important for the rise of oxygen in Earth's atmosphere and can be used to estimate the oxygen and hydrogen content of Earth's prebiotic atmosphere.

A zeolite membrane is a synthetic membrane made of crystalline aluminosilicate materials, typically aluminum, silicon, and oxygen with positive counterions such as Na+ and Ca2+ within the structure. Zeolite membranes serve as a low energy separation method. They have recently drawn interest due to their high chemical and thermal stability, and their high selectivity. Currently zeolites have seen applications in gas separation, membrane reactors, water desalination, and solid state batteries. Currently zeolite membranes have yet to be widely implemented commercially due to key issues including low flux, high cost of production, and defects in the crystal structure.

References

  1. 1 2 Kerry, Frank (2007). Industrial Gas Handbook: Gas Separation and Purification. CRC Press. pp. 275–280. ISBN   9780849390050.
  2. 1 2 Jang, Kwang-Suk; Kim, Hyung-Ju; Johnson, J. R.; Kim, Wun-gwi; Koros, William J.; Jones, Christopher W.; Nair, Sankar (2011-06-28). "Modified Mesoporous Silica Gas Separation Membranes on Polymeric Hollow Fibers". Chemistry of Materials. 23 (12): 3025–3028. doi:10.1021/cm200939d. ISSN   0897-4756.
  3. 1 2 3 Chong, K. C.; Lai, S. O.; Thiam, H. S.; Teoh, H. C.; Heng, S. L. (2016). "Recent progress of oxygen/nitrogen separation using membrane technology" (PDF). Journal of Engineering Science and Technology. 11 (7): 1016–1030.
  4. 1 2 3 4 5 6 Berend Smit; Jeffrey A. Reimer; Curtis M. Oldenburg; Ian C. Bourg (2014). Introduction to Carbon Capture and Sequestration. Imperial College Press. pp. 281–354. ISBN   978-1-78326-328-8.
  5. Richard W. Baker (2004). Membrane Technology and Applications. John Wiley & Sons Ltd. pp. 15–21. ISBN   978-0-470-85445-7.
  6. Wilcox, Jennifer (2014-04-16). Carbon Capture. Springer. ISBN   978-1-4939-0125-8.
  7. 1 2 Isalski, W. H. (1989). Separation of Gases. Monograph on Cryogenics. Vol. 5. New York: Oxford University Press. pp. 228–233.
  8. Robeson, L.M. (1991). "Correlation of separation factor versus permeability for polymeric membranes". Journal of Membrane Science. 62 (165): 165–185. doi:10.1016/0376-7388(91)80060-j.
  9. Robeson, L.M. (2008). "The upper bound revisited". Journal of Membrane Science. 320 (390): 390–400. doi:10.1016/j.memsci.2008.04.030.
  10. 1 2 3 4 5 6 7 8 9 10 11 12 Merkel, Tim C.; Lin, Haiqing; Wei, Xiaotong; Baker, Richard (2010-09-01). "Power plant post-combustion carbon dioxide capture: An opportunity for membranes". Journal of Membrane Science. Membranes and CO2 Separation. 359 (1–2): 126–139. doi:10.1016/j.memsci.2009.10.041.
  11. Chew, Thiam-Leng; Ahmad, Abdul L.; Bhatia, Subhash (2010-01-15). "Ordered mesoporous silica (OMS) as an adsorbent and membrane for separation of carbon dioxide (CO2)". Advances in Colloid and Interface Science. 153 (1–2): 43–57. doi:10.1016/j.cis.2009.12.001. PMID   20060956.
  12. 1 2 Kim, Hyung-Ju; Chaikittisilp, Watcharop; Jang, Kwang-Suk; Didas, Stephanie A.; Johnson, Justin R.; Koros, William J.; Nair, Sankar; Jones, Christopher W. (2015-04-29). "Aziridine-Functionalized Mesoporous Silica Membranes on Polymeric Hollow Fibers: Synthesis and Single-Component CO2 and N2 Permeation Properties". Industrial & Engineering Chemistry Research. 54 (16): 4407–4413. doi:10.1021/ie503781u. ISSN   0888-5885.
  13. Poshusta, Joseph C; Noble, Richard D; Falconer, John L (2001-05-15). "Characterization of SAPO-34 membranes by water adsorption". Journal of Membrane Science. 186 (1): 25–40. doi:10.1016/S0376-7388(00)00666-9.
  14. Kusakabe, Katsuki; Kuroda, Takahiro; Murata, Atsushi; Morooka, Shigeharu (1997-03-01). "Formation of a Y-Type Zeolite Membrane on a Porous α-Alumina Tube for Gas Separation". Industrial & Engineering Chemistry Research. 36 (3): 649–655. doi:10.1021/ie960519x. ISSN   0888-5885.
  15. Himeno, Shuji; Tomita, Toshihiro; Suzuki, Kenji; Nakayama, Kunio; Yajima, Kenji; Yoshida, Shuichi (2007-10-01). "Synthesis and Permeation Properties of a DDR-Type Zeolite Membrane for Separation of CO2/CH4 Gaseous Mixtures". Industrial & Engineering Chemistry Research. 46 (21): 6989–6997. doi:10.1021/ie061682n. ISSN   0888-5885.
  16. Li, S.; Falconer, J. L.; Noble, R. D. (2006-10-04). "Improved SAPO-34 Membranes for CO2/CH4 Separations". Advanced Materials. 18 (19): 2601–2603. Bibcode:2006AdM....18.2601L. doi:10.1002/adma.200601147. ISSN   1521-4095. S2CID   96222879.
  17. Alam, Syed Fakhar; Kim, Min-Zy; Kim, Young Jin; Rehman, Aafaq ur; Devipriyanka, Arepalli; Sharma, Pankaj; Yeo, Jeong-Gu; Lee, Jin-Seok; Kim, Hyunuk; Cho, Churl-Hee (2020-05-01). "A new seeding method, dry rolling applied to synthesize SAPO-34 zeolite membrane for nitrogen/methane separation". Journal of Membrane Science. 602: 117825. doi:10.1016/j.memsci.2020.117825. ISSN   0376-7388. S2CID   213870612.
  18. Wu, Ting; Diaz, Merritt C.; Zheng, Yihong; Zhou, Rongfei; Funke, Hans H.; Falconer, John L.; Noble, Richard D. (2015-01-01). "Influence of propane on CO2/CH4 and N2/CH4 separations in CHA zeolite membranes". Journal of Membrane Science. 473: 201–209. doi:10.1016/j.memsci.2014.09.021. ISSN   0376-7388.
  19. Li, Shiguang; Zong, Zhaowang; Zhou, Shaojun James; Huang, Yi; Song, Zhuonan; Feng, Xuhui; Zhou, Rongfei; Meyer, Howard S.; Yu, Miao; Carreon, Moises A. (2015-08-01). "SAPO-34 Membranes for N2/CH4 separation: Preparation, characterization, separation performance and economic evaluation". Journal of Membrane Science. 487: 141–151. doi:10.1016/j.memsci.2015.03.078. ISSN   0376-7388.
  20. Cao, Zishu; Anjikar, Ninad D.; Yang, Shaowei (February 2022). "Small-Pore Zeolite Membranes: A Review of Gas Separation Applications and Membrane Preparation". Separations. 9 (2): 47. doi: 10.3390/separations9020047 . ISSN   2297-8739.
  21. Gurdal, Yeliz; Keskin, Seda (2012-05-30). "Atomically Detailed Modeling of Metal Organic Frameworks for Adsorption, Diffusion, and Separation of Noble Gas Mixtures". Industrial & Engineering Chemistry Research. 51 (21): 7373–7382. doi:10.1021/ie300766s. ISSN   0888-5885.
  22. Keskin, Seda; Sholl, David S. (2009-01-21). "Assessment of a Metal−Organic Framework Membrane for Gas Separations Using Atomically Detailed Calculations: CO2, CH4, N2, H2 Mixtures in MOF-5". Industrial & Engineering Chemistry Research. 48 (2): 914–922. doi:10.1021/ie8010885. ISSN   0888-5885.
  23. Zornoza, Beatriz; Martinez-Joaristi, Alberto; Serra-Crespo, Pablo; Tellez, Carlos; Coronas, Joaquin; Gascon, Jorge; Kapteijn, Freek (2011-09-07). "Functionalized flexible MOFs as fillers in mixed matrix membranes for highly selective separation of CO2 from CH4 at elevated pressures". Chemical Communications. 47 (33): 9522–9524. doi:10.1039/c1cc13431k. ISSN   1364-548X. PMID   21769350.
  24. Venna, Surendar R.; Carreon, Moises A. (2010-01-13). "Highly Permeable Zeolite Imidazolate Framework-8 Membranes for CO2/CH4 Separation". Journal of the American Chemical Society. 132 (1): 76–78. doi:10.1021/ja909263x. ISSN   0002-7863. PMID   20014839.
  25. Kusakabe, Katsuki (1994-10-24). "Separation of CO2 with BaTiO3 membrane prepared by the sol—gel method". Journal of Membrane Science. 95 (2): 171–177. doi:10.1016/0376-7388(94)00109-X.
  26. Yun, S.; Ted Oyama, S. (2011). "Correlations in palladium membranes for hydrogen separation: A review". Journal of Membrane Science. 375 (1–2): 28–45. doi:10.1016/j.memsci.2011.03.057.
  27. Dolan, Michael D.; Kochanek, Mark A.; Munnings, Christopher N.; McLennan, Keith G.; Viano, David M. (February 2015). "Hydride phase equilibria in V–Ti–Ni alloy membranes". Journal of Alloys and Compounds. 622: 276–281. doi:10.1016/j.jallcom.2014.10.081.
  28. Han, Jiuli; Bai, Lu; Yang, Bingbing; Bai, Yinge; Luo, Shuangjiang; Zeng, Shaojuan; Gao, Hongshuai; Nie, Yi; Ji, Xiaoyan; Zhang, Suojiang; Zhang, Xiangping (3 September 2019). "Highly Selective Oxygen/Nitrogen Separation Membrane Engineered Using a Porphyrin-Based Oxygen Carrier". Membranes. 9 (115): 115. doi: 10.3390/membranes9090115 . PMC   6780238 . PMID   31484439.
  29. 1 2 3 Brunetti, A.; Scura, F.; Barbieri, G.; Drioli, E. (2010-09-01). "Membrane technologies for CO2 separation". Journal of Membrane Science. Membranes and CO2 Separation. 359 (1–2): 115–125. doi:10.1016/j.memsci.2009.11.040.
  30. 1 2 3 Bernardo P., Clarizia G. (2013). "30 Years of Membrane Technology for Gas Separation" (PDF). The Italian Association of Chemical Engineering. 32. S2CID   6607842. Archived from the original (PDF) on 2017-09-01.
  31. 1 2 Baker, Richard W. (2002-03-01). "Future Directions of Membrane Gas Separation Technology". Industrial & Engineering Chemistry Research. 41 (6): 1393–1411. doi:10.1021/ie0108088. ISSN   0888-5885.
  32. Bernardo P., Clarizia G (2013). "30 Years of Membrane Technology for Gas Separation" (PDF). The Italian Association of Chemical Engineering. 32. S2CID   6607842. Archived from the original (PDF) on 2017-09-01.
  33. 1 2 3 4 Brice Freeman, Pingjiao Hao, Richard Baker, Jay Kniep, Eric Chen, Junyuan Ding, Yue Zhang Gary T. Rochelle. (January 2014). "Hybrid Membrane-absorption CO2 Capture Process". Energy Procedia. 63: 605–613. doi: 10.1016/j.egypro.2014.11.065 .{{cite journal}}: CS1 maint: multiple names: authors list (link)
  34. 1 2 3 4 5 Lin, Haiqing; He, Zhenjie; Sun, Zhen; Kniep, Jay; Ng, Alvin; Baker, Richard W.; Merkel, Timothy C. (2015-11-01). "CO2-selective membranes for hydrogen production and CO2 capture – Part II: Techno-economic analysis". Journal of Membrane Science. 493: 794–806. doi: 10.1016/j.memsci.2015.02.042 .
  35. 1 2 3 Huang, Yu; Merkel, Tim C.; Baker, Richard W. (2014-08-01). "Pressure ratio and its impact on membrane gas separation processes". Journal of Membrane Science. 463: 33–40. doi:10.1016/j.memsci.2014.03.016.
  36. 1 2 Hao, Pingjiao; Wijmans, J. G.; Kniep, Jay; Baker, Richard W. (2014-07-15). "Gas/gas membrane contactors – An emerging membrane unit operation". Journal of Membrane Science. 462: 131–138. doi:10.1016/j.memsci.2014.03.039.