Sie sind auf Seite 1von 20

Membrane Filtration

Membrane filtration (MF) is a pressure-driven separation process that employs a


membrane for both mechanical and chemical sieving of particles and macromole-
cules (Benjamin and Lawler, 2013).

From: Advances in Chemical Mechanical Planarization (CMP), 2016

Related terms:

Protein, Wastewater, Adsorption, Pore Size, Wastewater Treatment, Filtration, Re-


verse Osmosis, Coagulation, Purification

View all Topics

26th European Symposium on Comput-


er Aided Process Engineering
M. Bahadir Saltik, ... Albert van der Padt, in Computer Aided Chemical Engineering,
2016

Abstract
Membrane filtration systems are preferred unit operations in industrial applications
due to their mild operating conditions. However the performance of a membrane
stack drops over time because of the membrane fouling. This decrease is overcomed
by introducing clean membrane stacks. The associated scheduling problem, when
to include new membrane stacks to the operation, is the main topic of this paper.
We construct a dynamic optimization problem to find the optimal time instants of
introducing new membrane stacks. Furthermore, the optimal operating pressure
profile for the optimal scheduling strategy is constructed from the desired output
specifications. The result of the simulation study indicates that the optimal schedul-
ing strategy improves the operation by slowing down the accumulation of fouling.

> Read full chapter


MEMBRANE SEPARATIONS | Filtra-
tion
R. Sahai, in Encyclopedia of Separation Science, 2000

Membrane Filtration
Membrane filtration through a very thin filter medium is also known as ‘surface
filtration’. The solid particles to be separated are usually large compared to the
pore size characteristic of the membrane. The pores on the surface are of irregular
shapes. The rejection of particles is dependent on several factors affecting the
transport through these pores into the tortuous channels. The separation is based
on exclusion discrimination by physical size, charge or affinity or a combination of
these properties. Large particles are rejected on the surface and do not accumulate
on the surface and do not get a chance to enter into the interior of the filter.

Other types of membrane filters are screen filters and here the pores do not lead
into tortuous capillary paths. The pore size is uniform but the distribution of the
pores is random on the filter surface. The filter is made by bombarding a thin
polycarbonate film with neutrons in a reactor. The film is then placed in a bath
of etching solution which preferentially attacks the polymer along the track of the
neutrons. The pore size is regulated by selecting the appropriate reagent, exposure
time and temperature.

Membrane filtration can be dead end or cross-flow. In dead-end filtration all the so-
lution is forced through the membrane. Retained particles collect on the membrane
surface and in the filter greatly reducing flow. A current application of dead-end
filtration is in bacterial testing where the liquid to be tested is passed through
the filter retaining all bacteria on the surface. Most chromatographic filtration
applications are of this type. In cross-flow membrane filtration, the feed liquid flows
tangentially to the membrane surface, which prevents the build up of cake on the
membrane. Both types of filtration use similar membranes.

By convention, membrane filtration or microfiltration is limited to membranes


used to remove particles larger than 0.1 μm in diameter. Membranes able to
remove smaller particles are called ‘ultrafiltration membranes’ and microsolutes can
be removed by reverse osmosis. Ultrafiltration and reverse osmosis are discussed
elsewhere. This article is limited to the process of microfiltration.

The filtration thresholds of common membrane-filtration processes are shown in


Table 1.

Table 1. Filtration threshold of common membrane-filtration processes


Type of filtration Impermeability of membrane
Reverse osmosis <0.001 μm
Ultrafiltration  0.001–0.1 μm
Microfiltration  0.1–10 μm

Microfiltration
Microfiltration is used to separate suspended solids or colloidal particles between 0.1
and 10 μm in diameter from solution. Most of the chromatography applications are
microfiltration based. The same type of membrane with different pore size is used for
these applications. The membrane acts like a physical sieve. The fluid passes through
tortuous channels while the particles are rejected on the surface of the filter. It can be
easily understood as a mechanical sieve with pores leading into a capillary forming
a tortuous path; within this tortuous path, there could be mechanical entrapment
and adsorption (Figure 1).

Figure 1. Tortuous path of micro- and ultrafiltration.

Microfiltration membranes can be subjected to harsher conditions compared to


ultrafiltration membranes. Membranes of different polymers in varying pore sizes
are available. Even nominally the same pore-size membranes of a polymer may differ
from each other in filtration characteristics because they may have different pore-size
distributions, i.e. varying pore size all across the membrane. To aid in wetting, many
membranes have some surfactant pretreatment and their effective pore size may be
different from the real pore size. Often a membrane filter becomes more efficient
as small particles are entrapped within the pores. The large particles captured on
the filter can also alter the effective particle-size rejection in subsequent filtration.
Filter capacity may vary depending on the solute particle-size variation in the feed.
Uniform size particles result in faster clogging of filters.

> Read full chapter


Isolation and Purification of Industrial
Enzymes: Advances in Enzyme Technol-
ogy
Soumya Mukherjee, in Advances in Enzyme Technology, 2019

Membrane Filtration
Membrane filtration involves the use of membrane technology for the separa-
tion of biomolecules and particles for the concentration of process fluids. During
separation, a semi-permeable membrane acts as a selective barrier retaining the
molecules/particles bigger than the pore size, while allowing the smaller molecules
to permeate through the pores (Fig. 2.4). Membrane filtration processes can be dis-
tinguished based on the type of force driving the transport through the membrane,
and are named by the size of the pores in the filter.

Fig. 2.4. Membrane process based on particle retention.

The three major types of membrane filtration process based membrane pore size
and the particle size are microfiltration, ultrafiltration, and reverse osmosis:

• Microfiltration is used for the separation of particles 0.02–10 μm in diameter,


primarily during the primary recovery stages of DSP. Microfilters are available
in materials such as ceramic and steel that can be aggressively cleaned and
sterilized.
• Ultrafiltration largely depends on effective fluid-management techniques. By
using hydrodynamic considerations, polarized solutes can be sheared from
the membrane surface, thereby increasing the back diffusion and reducing
the decline in performance.The application of the two-stage UF strategy has
been successfully implied on separation of different types of proteins, such
as -lactalbumin and -lactoglobulin, from whey protein isolate [12, 13],
surfactin from fermentation broth [14], and for bromelain separation [15].
Nor et al. 2016 [15] introduced ceramic membranes using zirconium oxide as
polymeric material, and reported an increment of 2.5-fold purity of bromelain
by applying the two-stage UF system, which consisted of membranes with
MWCO bigger (in the UF stage 1) and smaller (in the UF stage 2) from the
bromelain molecular weight, when bromelain has a molecular weight ranging
from 23.4 to 35.73 kDa. Most of the previous studies have used polymeric
membranes, such as polysulfone, polyvinyl fluorite, and cellulose acetate. Ex- •
amples of ceramic membranes include alumina ( -Al2O3 and -Al2O3), zirconia
(ZrO2), titania (TiO2), glass (SiO2), and silicon carbide (SiC). A second variation
of ultrafiltration is diafiltration, in which water or other liquids are filtered to
remove unwanted low molecular weight contaminants. This can be used as an
alternative to gel filtration or dialysis for removing ammonium sulfate from a
protein preparation precipitated by the salt (de-salting), for changing a buffer,
or in water purification.
Reverse osmosis (RO) is completely the opposite of the process of osmosis. Os-
mosis occurs naturally without requiring external energy, whereas RO requires
applying external energy/pressure using a high-pressure pump on the side
of the highly concentrated solution. At osmotic pressure, the flow between
two solutions will cease, whereas applying pressure or energy greater than
osmotic pressure reverses the flow from a highly concentrated solution toward
a less concentrated solution. The amount of pressure required depends on
the salt concentration of the feed water. The higher the concentration of the
feed water, the more pressure is required to overcome the osmotic pressure.
There are three major types of reverse osmosis (RO) membranes (Fig. 2.5).
The pore size for an RO membrane is around 0.0001 μm. Cellulosic Fully
aromatic polyamide Thin film compositeFig. 2.5. (A) Cellulose filters exhibit
particle retention levels down to 2.5 μm; (B) glass microfilters are made from
borosilicated glass; (C) multigrade GMF 150 combines two filters in one for
fast effective multilayered filtration; (D) diatomaceous earth also known as
kieselgarh is the fused skeletal remains of diatoms composed of silica; and
(E) perlite is naturally occurring glassy volcanic rock.

Currently, many strategies are being developed to circumvent the technical bottle-
neck associated with DSP separation challenges. Some processes, such as (1) fusing
target proteins to transgenic plants, and (2) heat precipitation of host cell proteins
(HCPs), [16] are routinely used in industries.

Transgenic plants are routinely used for the production of proteins, which includes
several steps designed to result in the high expression of foreign proteins, from gene
manipulation to breeding. Examples of such common proteins are -glucuronidase,
avidin, laccase, and trypsin. These proteins represent a range of molecular weights,
activities, and localizations, demonstrating the versatility of the system. The advan-
tages of using transgenic plant technology for industrial enzyme production include
replacement of chemicals that cause environmental pollution. A study reported the
heterologous expression of amylopullulanase in maize seeds, and (APU) derived
from the bacteria Thermoanaerobacter thermohydrosulfuricus. Subsequent fermenta-
tion of the accumulated starch showed directly enhanced bioethanol production
from maize grains. This study showed simplification of starch-based bioethanol
production from maize grains using a biotechnology approach [17]. The fusion of
transgenic proteins to elastin-like peptides (ELPs) is known to enhance the accumu-
lation of transgenic proteins in plants, resulting in higher yield and simpler recovery.
Studies showed that the presence of the ELP tag enhanced the accumulation of
the heterologous proteins in tobacco leaves. The centrifugation-based, or mem-
brane-based inverse transition cycling method (cITC) further helped in the precip-
itation of ELPylated proteins by a combination of salting, heating, centrifugation,
and resolubilization in the absence of salt at a lower temperature [18]. This method
has been further improved by the use of microfiltration to isolate the precipitate in a
more purified form (“the good”). Plant-produced industrial enzymes and transgenic
antigens have been purified from the plant matrix using either Protein A-based
affinity chromatography via Fc fusions [19], Ni-column chromatography and anion
exchange chromatography [20], or a combination of two phase separations, includ-
ing several membrane filtration steps and gel filtration [21].

Thus, by systematic characterization of the host cell protein criteria, an efficient


DSP is possible, with better quality of the product. This categorization can be used
for an USP-DSP integration approach toward an efficient production process by
circumventing the generation or accumulation of “bad” and “ugly” impurities.

> Read full chapter

Disposal of metalworking fluids


M. Graff, in Metalworking Fluids (MWFs) for Cutting and Grinding, 2012

10.5 Physical water treatment processes: membrane filtration


Membrane filtration is also known as cross-flow filtration. It is a gentle, physical
separation process where the emulsion is not thermally or chemically modified (Fig.
10.3). In membrane filtration the filtrate is called permeate, the concentrate is named
retentate and the build-up of dry sludge during membrane filtration is called the
coating. The range of membrane filtration processes is summarized in Fig. 10.4.
10.3. Membrane separation.

10.4. Range of filtration processes.

All membrane processes are concentration-dependent processes. Membranes have


the capacity to hold specific substances back and let others pass selectively. Criteria
for separation of compounds include particle size and polarity. The difference from
classic filtration is that the separation effect goes down to the molecular level so that
higher density molecular substances can be separated from lower density molecular
substances.

A typical process scheme for wastewater treatment using membrane processes is


shown in Fig. 10.5. The driving forces behind the mass flow through the mem-
brane are the gradients (e.g. pressure or concentration) between both sides of the
membrane: An important criterion for dimensioning in membrane filtration is the
quantity (e.g. the volume of a medium or substance which is transported through a
membrane per unit of time).
10.5. Flow chart for a membrane separation process.

There are a number of membrane separation techniques:

• microfiltration

• ultrafiltration

• nanofiltration

• reverse osmosis.

These techniques differ from each other according to their different separation
barriers and operating parameters. The most important of these techniques are
discussed below.

10.5.1 Micro-and ultrafiltration


Micro- and ultrafiltration are two membrane processes that have been established
as standard processes for oil-water separation of used emulsions from metal pro-
cessing for sewage water treatment. Microfiltration is commonly used for low oil
concentrations, for example for the treatment of compression condensates. The oil
concentration of the retentate to aim for is around 10%. Beyond this point the oil
concentration of the permeate increases significantly. Ultra filtration is particularly
suitable for wastewater treatment of used emulsions from the metal processing
industry. Most water-mixed coolant emulsions are treated using this technology. The
driving force in both micro- and ultrafiltration is the differential pressure between
the feed and drain of the membrane, which can lie between 1 and 10 bar. With these
operating pressures, the processes can achieve a surface density flux from 50 1 (m2
h−1) to ca. 500 1 (m2 h−1).

Microfiltration membranes are open-pored symmetric membranes where the width


of the pores inside the body of the membrane is not variegated. Calculation of the
retention is based on the filter effect, meaning that all molecules smaller than the
size of the pores pass through the membrane and larger molecules are held back
by the membrane. Microfiltration operates along the principle of mechanical size
elimination. It is important to mention that the steric composition of the molecules
can play a role in filtration through the membrane.

Ultrafiltration membranes, in contrast, have an asymmetric structure. The body


of the membrane is made of a separation active layer with asymmetric pores in
the submicrometre group and an open-pore support layer. In ultrafiltration, in
contrast to microfiltration, the polarity of the membrane influences the retention
and the permeate flow, rather than the size of particles. Owing to the polarity of
the membrane, diffusion processes also occur at the surface of the membrane. The
relative importance of the diffusion and filter effects differs, depending on individual
membrane materials and their polarity.

Significantly higher permeate flows have been achieved with hydrophilic mem-
branes than using lipophilic membranes. Hydrophilic membranes are characterized
by smaller contact angles between the membrane and water droplets (see Fig.
10.6). As a result, hydrophilic membranes show a higher flow volume through the
membrane than lipophilic membranes or, put in different words, a higher permeate
capacity.

10.6. Contact angle for a hydrophilic membrane.

A common characteristic of micro- and ultrafiltration is the tangential flow of the


emulsion relative to the membrane. This flow minimizes the build-up of a coating
on the membrane through tangential gravitational forces. However, coating layers,
which reduce the effectiveness of the membrane, inevitably build up over time and
with increasing concentration. Periodic permeate rinsing can be used to minimize
the coating layers.

Two basic types of membranes are used:

• organic membranes with membrane material made from polyamides, poly-


sulphones and cellulose acetates
• inorganic membranes with membrane material made from metal oxides,
titanium and aluminium oxides, zirconium oxide, carbon silicides on a ceramic
carrier or carbon materials.

Ceramic materials, in contrast to organic materials, have the advantage of greater


chemical, thermal and mechanical durability. Micro- and ultrafiltration membranes
are manufactured in the form of flat or tube membranes or hollow fibres that fit into
pressure pipes (Fig. 10.7). A unit made of a pressure pipe and a membrane is called
a module. For sewage water treatment, pipe and spiral wound modules are used.
Pipe modules have a pipe diameter of 3–24mm. It is important to pay attention to
a suitable pre-filTtration mechanism to avoid blockages on the feed side.

10.7. Spiral wound (a) and ceramic (b) membranes.

Permeate from membrane separation typically has a mineral oil carbon hydrogen
concentration of <10 mg 1−1. The concentration of heavy metal and salt depends on
the composition of the wastewater and usually exceeds critical values. Low molecular
organic substances are often retained in the permeate, which means it requires
further treatment. Possible re-treatment methods depend on the condition of the
watery phase. The oil content of retentates ranges between 30 and 45%. The oil
component can either be thermally recycled or sent to a re-refining process.

10.5.2 Combined wastewater treatment processes


Depending on the composition of used emulsions, a single process is usually not
enough to ensure compliance with European and other regulations governing the
disposal of wastewater. The following processes can be used:

• precipitation/flocculation

• nanofiltration and/or reverse osmosis

• evaporation

• biological processes (under certain conditions).

A common process combination used in metal processing consists of ultrafiltration


followed by chemical precipitation/coagulation. During the first step, ultrafiltration
is used to separate the oil from the used emulsion. Then, during chemical precipita-
tion/coagulation, heavy metals in the permeate settle in the form of sludge and can
be drawn off. The quality of filtrates from chemical precipitation/coagulation usually
meets the relevant standards for wastewater disposal. In exceptional cases drawing
off the sludge can turn out to be problematic.

Less common is the arrangement of an evaporator and a membrane system. The


type of membrane system used depends on the intended quality of the distillate.
Usually an ultrafiltration or microfiltration system is used. The permeate can still be
contaminated with a few short-chained organic substances, so it is more appropriate
to re-use it within the manufacturing site in some way.

Even less commonly used is a relatively new and innovative process combination
of ultrafiltration and nanofiltration. Since this is a combination of two membrane
systems, the level of automation can be significantly increased. Permeate resulting
from nanofiltration is reliable enough to meet the critical values required by Euro-
pean standards.

10.5.3 Nanofiltration and reverse osmosis


In contrast to micro- and ultrafiltration, which are based on blown membranes,
nanofiltration and reverse osmosis use semi-permeable or diffusion membranes.
The separating effect is no longer dependent on the size of particles but on the
dissolving potential of the materials that the membranes are made from. The
resulting osmotic pressure generated has to be overcome to enable a flux through
the membrane. The operating pressures of this process range from 10–50 bar. For
special applications, such as water treatment of percolating water on landfill sites,
the process can be run at operational pressures of up to 120 bar.

The molecular separation barrier of nanofiltration ranges from 200–2000 Dalton. A


dalton is an American measurement unit for molecular weight. Nanofiltration mem-
branes carry an electric charge which allows separation of molecules of the same size
but different charge. Because of its separation and charge properties nanofiltration
is able to separate lower molecular compounds and polyvalent ions like heavy metals
and sulphates successfully. Nanofiltration membranes currently available on the
market are organic. Like all organic membrane materials, nanofiltration membranes
have a limited temperature and chemical stability. In recent years efforts have been
made to develop ceramic-based nanofiltration membranes. Products currently on
the market have not yet achieved the necessary separation. An additional challenge
is the high investment costs.

The separation barrier in reverse osmosis is still significantly below that in nanof-
iltration. Reverse osmosis is a well-known membrane process which is mostly used
for desalination of drinking and well water without the application of chemical
substances and thermal energy. The membranes used are exclusively organic mem-
branes.

> Read full chapter

Environmental aspects of planarization


processes
D.E. Speed, in Advances in Chemical Mechanical Planarization (CMP), 2016

10.8.1 Membrane filtration


Membrane filtration (MF) is a pressure-driven separation process that employs a
membrane for both mechanical and chemical sieving of particles and macromole-
cules (Benjamin and Lawler, 2013). Membrane filters can be operated in dead-end
configuration, but more commonly are operated in a cross-flow configuration where
the pressurized feed water is recirculated at a high velocity across the face of the
membrane. Water that passes through the membrane is referred to as permeate,
whereas the water, solids, and solute that are rejected by the membrane are referred
to as either concentrate or reject. The feed is typically recirculated at high velocity in
order to impart a shearing action that helps mitigate the accumulation of solids on
the membrane. The pressure difference across the membrane is referred to as the
transmembrane pressure (TMP), and serves as the driving force for water migration
across the membrane. There are alternative ways to express the transmembrane
potential but one common way is to calculate it as the difference between the feed
and permeate pressure (Pp), where the feed pressure is taken as the average between
the module inlet (Pi) and outlet (Po) pressure, as expressed by Eqn (10.7) (Benjamin
and Lawler, 2013):

(10.7)

The membrane flux is the flow rate of permeate per unit area of membrane surface
and typically proportional to the TMP. The flux for a new membrane, operating with
water only, is referred to as the clean water flux, and serves as a useful benchmark.
Clean water flux rates for membranes that are used for wastewater treatment may
be of the order of 3–4 m3/m2/day.

As wastewater permeates through the membrane there is a tendency for a cake


of solids to form on the feed side of the membrane. The thickness of the cake is
typically limited by the shearing action of the cross-flow, such that the flux undergoes
an initial decay, and then stabilizes to a nearly steady-state value. With continued
use, the cake may begin to densify, at which point the membrane typically requires
cleaning to prevent further decline in the flux rate. If, however, some of the solids
or dissolved chemicals in the feed water load the pore structure of the filters, the
flux may decline in an unrecoverable manner. In addition to the tendency for the
formation of a cake on the feed side of the membrane concentration polarization
may also limit the water flux through the membrane. If the flux is held at a constant
value, the transmembrane pressure will increase as the membrane becomes more
fouled. Conversely, if the TMP is held constant, the flux will decrease as the mem-
brane becomes fouled. For many systems it is desirable to maintain a constant flux
and allow the TMP to increase until cleaning is required.

There are a wide variety of commercially available membranes that can be selected
in consideration of the physical and chemical characteristics of the wastewater
and the treatment objectives. Membranes are often categorized according to the
size of particles or molecules removed, and the operating TMP, as summarized
in Table 10.5 (Koros et al., 1996; Benjamin and Lawler, 2013). Important factors in
membrane selection include mechanical strength, pore size and surface charge, and
resilience to the cleaning chemicals that are almost inevitably required to keep the
membrane clean. A typical membrane filter installation is depicted in Figure 10.10,
where wastewater is introduced to a feed tank and recirculated through the filter
at a high velocity, generating a permeate and a concentrate. This system can be
operated continuously, with the net feed rate equal to the combined permeate and
concentrate flow rates.

Table 10.5. Pressure-driven membrane separation processes

Basis of separa- Typical TMP (kPa) System recovery Regime


tion (%)
Microfilter Particles and dis- 10–100 90–99+ Pore size exclusion
solved macromol- at membrane in-
ecules larger than terface
100 nm are reject-
ed
Ultrafilter Particles and dis- 50–300 85–95+
solved macromolecules
smaller than 100 nm
and larger than 2 nm
are rejected
Nanofilter Particles and dis- 200–1500 75–90+ Solution, diffusion
solved macro- through mem-
molecules smaller brane
than 2 nm are
rejected
Reverse osmosis 500–8000 60–90
Figure 10.10. Process flow diagram for a membrane filter.

MF has been effectively employed in numerous wastewater treatment systems and


can provide a very effective and efficient means of solid–liquid separation. How-
ever, the evaluation of a membrane separation process for any particular wastewater
requires careful evaluation of the susceptibility of the membrane to fouling, which
can be the Achilles heel of membrane separations. Important performance variables
include the rejection of solids, the TMP, the flux rate, and the volume concentration
factor; all of which are best evaluated over a prolonged piloting period. Typically the
flux rate decays over time as solids build up on, and possibly within, the membrane,
and the filter must undergo some form of periodic regeneration. Membrane regen-
eration is generally accomplished by some combination of pressure backpulsing,
membrane movement, and/or chemical cleaning. In evaluating the applicability
of a membrane process to a particular wastewater it is important to distinguish
reversible fouling, which can be recovered by periodic membrane cleaning, and
irreversible fouling, which can lead to a progressive irrecoverable loss of filtration
capacity. Reversible fouling can be mitigated by the introduction of periodic cleaning
steps into the filter’s operational cycle, but requires application-specific process
development that must be accommodated into the filter system design, capacity
statement, and operating plan.

MF is also sometimes employed for recycling of CMP wastewater. In recycling


applications where the objective is to recover the water, the MF permeate may
undergo additional treatment by reverse osmosis to recover a purified water recycle
stream. Alternatively, the reject from a membrane filter can be employed as a means
of recovering and recycling the slurry itself (Testa et al., 2011, 2014).

Laboratory- and pilot-scale evaluations of the application of MF to CMP wastewater


treatment have been reported by Chang et al. (2006), Huang et al. (2004, 2011a,b),
Juang et al. (2008), Kim et al. (2002), Lo and Lo (2004), Pan et al. (2005), Springer
et al. (2013), Su et al. (2014), Testa et al. (2011), Wu et al. (2004), and Yang and Yang
(2004), Yang et al. (2012), among others. In the work by Springer et al. (2013), 10
and 100 kDa MW cut-off filters of regenerated cellulose membranes were evaluated
for the treatment of a Klebosol 30R50 slurry, with mean particle size of 28 nm,
diluted to 1.4 g/L as SiO2. The 100 kDa membrane provided a steady state flux
of 0.37 m3/m2/day at a TMP of 80 kPa, and a 0.41 m3/m2/day flux at 200 kPa.
The 10 kDa membrane provided a flux of 0.35 and 0.51 m3/m2/day at 80 and
200 kPa, respectively. However, with repeat use, the flux rate for the 100 kPa was
nonrecoverable, whereas the flux rate on the 10 kDa membrane was recoverable.
Su et al. (2014) evaluated the use of a 30 kDa GE Osmonics polyvinyldiene fluoride
membrane for the treatment of Klebosol 1501-50 silica slurry with a mean particle
size of 75 nm. They measured the steady-state flux at concentrations of 500, 1000,
1500, and 2000 mg/L SiO2, and found that the flux decreased systematically with
increasing concentration. At a feed concentration of 1500 mg/L the steady-state flux
at the end of 2 h was 1.36 m3/m2/day at 275 kPa. Similarly, Su et al. (2014) obtained
1.05 m3/m2/day flux at 275 kPa with a Cabot SS-25 silica slurry, which had 150 nm
mean particle size. Yang et al. (2012) evaluated the treatment of a semiconductor
backgrind wastewater (1366 mg/L total solids (TS); 546 mg/L total dissolved solids
(TDS); 1366 Nephelometric Turbidity Units (NTU); pH 7.9) with a GE Osmonics spiral
wound polysulfone membrane with a mean pore diameter of 40 nm. They obtained
1.92 m3/m2/day at 40 kPa with untreated wastewater, and 2.16 m3/m2/day when
the wastewater feed was precoagulated with PACl. The precoagulation changed the
mean particle size from 1680 to 17,800 nm in diameter. Huang et al. (2004) found
that the addition of 0.2 mg/L cationic PAA increased the filterability, but increased
the effluent Si concentration. However, increasing the PAA to 1 mg/L PAA resulted in
an overdose that caused rapid filter fouling. Applications of MF to CMP wastewater
containing high concentrations of silica must be evaluated carefully in consideration
of the slow kinetics of silica precipitation and/or formation of polymeric silica.

> Read full chapter

Membranes for Microbial Fuel Cells


Sangeetha Dharmalingam, ... Moogambigai Sugumar, in Microbial Electrochemical
Technology, 2019

1.7.9.4.2 Microfiltration Membrane Filtration Membranes


Microfiltration Membrane filtration membranes (MFMs) are extensively used as
sludge separators for wastewater treatment in view of their high durability and
filtration efficiencies. Thus, their use has been comprehensive to MFCs [126,127].
These membranes as other membranes discussed earlier act as a barrier isolating
the anodic and cathodic solutions allowing various charged and neutral species to
pass through them depending on their pore size. They are cheaper than most IEMs
and are mainly used for wastewater treatment. In addition, many other cost-effective
polymer MFMs including nylon mesh, cellulose filters, and polycarbonate filters
have also been widely reported to function as separators [64]. In spite of these
advantages, MFM faces certain limitations such as higher permeation of oxygen,
substrate, and resistance. Mass transport and ohmic resistance are two contentious
issues in most cases. Substrate and oxygen permeation will result in high internal
resistance and lower power density. In order to reduce these effects, large electrode
spacing is essential [77,125]. Hence, to justify the use of MFMs, these issues should
be addressed before using them for MFC applications [71].

> Read full chapter

Fundamentals of Quorum Sensing, An-


alytical Methods and Applications in
Membrane Bioreactors
Hanife Sari Erkan, ... Güleda Önkal Engin, in Comprehensive Analytical Chemistry,
2018

2.3.1 Biomass Parameters

2.3.1.1 MLSS Concentration


As membrane filtration is a kind of solid–liquid separation, the MLSS in MBR should
be considered as one of the important foulants. There are contradictory results on
the effect of MLSS concentration on membrane fouling. It was reported that the
increase in MLSS has negative impact on the hydraulic performance of MBR, if the
other operational parameters are not taken into account (Chang and Kim, 2005;
Cicek et al., 1999). In some studies it was found out that an increase in MLSS
concentrations showed positive impact (Brookes et al., 2006; Defrance and Jaffrin,
1999), whereas, in some cases, insignificant impact (Hong et al., 2002; Le-Clech et
al., 2003) on membrane performance was reported. It was also reported that negative
impact was found, when the MLSS concentration is above 30 g/L (Lubbecke et al.,
1995). According to Rosenberger et al. (2005), membrane fouling decreases with
an increase of MLSS concentration at low MLSS concentration (< 6 g/L); however,
membrane fouling increase with an increase of MLSS above 15 g/L. The MLSS
concentration does not have a significant effect on membrane fouling, when it
varies between 8 and 12 g/L. On the other hand, some authors reported that the
concentration of MLSS might not significantly affect fouling propensity at low flux
operations in MBR. The MLSS in MBR is mainly responsible for the resistance of cake
layer (Itonaga et al., 2004). Generally, submerged MBRs are operated at constant
flux mode where cake layer resistance is limited. The concentration of MLSS can
be said to be a poor indicator of membrane fouling propensity, as there is no clear
correlation between MLSS concentration and other foulant properties (Jefferson et
al., 2004).

> Read full chapter

Fundamentals
In The MBR Book, 2006

2.1.4.4 Concentration polarisation


For membrane filtration processes, the overall resistance at the membrane:solution
interface is increased by a number of factors which each place a constraint on the
design and operation of membrane process plant:

(a) the concentration of rejected solute near the membrane surface,

(b) the precipitation of sparingly soluble macromolecular species (gel layer forma-
tion) at the membrane surface,
(c) the accumulation of retained solids on the membrane (cake layer formation).

All of the above contribute to membrane fouling, and (a) and (b) are promoted by
CP. CP describes the tendency of the solute to accumulate at membrane:solution
interface within a concentration boundary layer, or liquid film, during crossflow
operation (Fig. 2.12). This layer contains near-stagnant liquid, since at the membrane
surface itself the liquid velocity must be zero. This implies that the only mode of
transport within this layer is diffusion, which is around two orders of magnitude
slower than convective transport in the bulk liquid region. However, it has been
demonstrated (Romero and Davis, 1991) that transport away from the membrane
surface is much greater than that governed by Brownian diffusion and is actually
determined by the amount of shear imparted at the boundary layer; such transport
is referred to as “shear-induced diffusion”.

Figure 2.12. Concentration polarisation


Rejected materials nonetheless build up in the region adjacent to membrane, in-
creasing their concentration over the bulk value, at a rate which increases expo-
nentially with increasing flux. The thickness of the boundary layer, on the other
hand, is determined entirely by the system hydrodynamics, decreasing in thickness
when turbulence is promoted. For crossflow processes, the greater the flux, the
greater the build-up of solute at the interface; the greater the solute build-up, the
higher the concentration gradient; the steeper the concentration gradient, the faster
the diffusion. Under normal steady-state operating conditions, there is a balance
between those forces transporting the water and constituents within the boundary
layer towards, through and away from the membrane. This balance is determined
by CP.

> Read full chapter

Fundamentals of Quorum Sensing, An-


alytical Methods and Applications in
Membrane Bioreactors
Reham M. Abu Shmeis, in Comprehensive Analytical Chemistry, 2018

2.11.1 Membrane Filtration Method


In the membrane filtration (MF) method, a measured volume of sample is passed
through a membrane filter of known pore size that retains bacteria on or near the
filter surface. The colonies that develop after incubation on a selective medium are
presumed to have originated from individual bacteria, thus representing a direct
count of the number of bacteria in the original sample. The MF method is often the
method of choice in the analysis of most water types (Russell, 2002; Smith et al.,
2013; Tortora et al., 2010).

> Read full chapter

Affinity Separation: Affinity Mem-


branes☆
Karsten Haupt, S.M.A. Bueno, in Reference Module in Chemistry, Molecular Sci-
ences and Chemical Engineering, 2015
Membrane Geometry
Just like filtration membranes in general, affinity membranes can be produced in
different configurations, and membrane modules of various geometries are com-
mercially available or have been manufactured in research laboratories (Figure 1).

Figure 1. Different geometries of affinity membranes: (a) flat sheet, (b) stack of flat
disks, (c) hollow fiber, (d) spiral-wound flat sheet, and (e) continuous rod. The arrows
indicate flow directions.

Flat sheet or disk membranes can be mounted as individual membranes in specially


designed cartridges or in commercial ultrafiltration units for use in dead-end f-
iltration mode. This allows for the production of inexpensive single or multiple-use
devices for the rapid adsorption of a target molecule from dilute samples in batch
or continuous recycling mode. Cartridges are also available that allow for operation
in cross-flow filtration mode.

Stacks of flat membrane disks have been employed for affinity membrane chro-
matography in column-like devices, the main purpose being to increase the adsorp-
tion capacity. Another configuration is continuous rod-type membranes which can
be directly cast in a chromatographic column. Both types of membrane columns
are compatible with conventional HPLC or FPLC systems and have advantages over
columns packed with beaded resins as described above. Being highly porous with
a mean pore diameter of 0.1–10 μm, they allow for efficient separations even at
high flow rates.
If the target molecule is to be recovered from complex feed solutions such as
cell homogenates or blood plasma, or from solutions containing high-molecular
mass additives such as antifoam agents or even particulate material, the use of
membranes in dead-end filtration mode is often impossible due to membrane
fouling. A remedy to this problem is the operation in cross-flow filtration mode
where the build-up of a polarization layer at the membrane surface is avoided or
diminished. Hollow-fiber membranes are well adapted for such applications. They
are usually mounted as bundles in tubular cartridges. Another configuration are flat
sheet membranes that are spiral-wound around a cylindrical core. Both systems have
the advantage of high surface-area/cartridge-volume ratios and high operational
capacities.

> Read full chapter

ScienceDirect is Elsevier’s leading information solution for researchers.


Copyright © 2018 Elsevier B.V. or its licensors or contributors. ScienceDirect ® is a registered trademark of Elsevier B.V. Terms and conditions apply.

Das könnte Ihnen auch gefallen