Ceramic Membranes: Fabrication & Applications
Ceramic Membranes: Fabrication & Applications
11 Ceramic Membranes
Minghui Qiu, Xianfu Chen, Yiqun Fan, and Weihong Xing, Nanjing Tech University, Nanjing, China
r 2017 Elsevier B.V. All rights reserved.
1.11.1 Introduction
Ceramic membranes are typically made of aluminum oxide, zirconium oxide, or titanium oxide, silicon carbide, which all belong
to the group of inorganic materials. They are resistant to mechanical, chemical, and thermal stress and have high porosity and a
hydrophilic surface.1 Since the 1980s, ceramic membranes were gradually being used in civilian areas and developed rapidly.2
With their excellent properties, they are not only used in the field of water treatment, but also suitable for applications under harsh
environments, offering reliable performance over long periods of time.
The structure of membranes dictates the appropriate separation process. Membrane separation applications and the corre-
sponding sizes are given in Fig. 1. Ceramic membranes include pores with a defined average size, ranging from macrometers to
nanometers, and they can be used for liquid filtration, gas separation, and pervaporation. But ceramic membranes suitable for the
separation of gases are either microporous or dense, such as microporous silica membranes or zeolite membranes. The driving
force for gas transport is a concentration of chemical potential gradient across the membrane. In this article, we focus on ceramic
membranes for filtration of liquid from microfiltration (MF; 50 nm–1 mm), ultrafiltration (UF; 2–50 nm) to nanofiltration (NF;
o2 nm), which is the main application of ceramic membranes.
In general, ceramic membranes tend to have asymmetric structures composed of a thin, selective layer that fulfills the
separation requirements, and a permeable support structure comprising one to five layers, as shown in Fig. 2. Based on what cutoff
rate shall be achieved, several membrane layers are fixed on the support, starting from coarse intermediate layers to fine separation
layers, with pore sizes decreasing until the designated pore sizes are reached. The aperture ratio of the intermediate layer is less than
that of the support layer, and the purpose of the layer is to prevent particles from permeating the porous support layer while
preparing the function layer.3 Typically, the higher the membrane selectivity required, the more support layers are needed.
Moreover, the thinner top layer of asymmetric membranes enables greater permeability.
The support layer aperture ranges from 1 to 20 mm, with porosity of 30–65%, and its major role is to ensure the mechanical
strength of the overall membrane. Supports define the geometries of the ceramic membrane elements, and they can be flat sheet,
hollow fiber, tubular, or multichannel. Multichannel ceramic membranes have been widely used in industrial applications. While
hollow-fiber, tubular, and flat-sheet ceramic membranes are mostly used in laboratory studies, their applications in industrial
settings are in the initial stages. A picture of a ceramic membrane with a multichannel design is displayed in Fig. 3.
As already mentioned, according to the pore size of the separation layer, ceramic membranes for liquid filtration can be
classified into MF (pore size 450 nm), UF (pore size 2–50 nm), and NF (pore size o2 nm). The fabrication methods for ceramic
membranes strongly depend on their targeted structures, especially the range of the pore size. Generally, solid-state sintering and
sol–gel are the two common techniques for the preparation of microfiltration and ultrafiltration membranes in the large-scale
continuous process. These two methods have a similar process, starting from the well-dispersed suspension or sol coating on the
top of a membrane substrate, followed by heat treatment to achieve the desired final selectivity for MF, UF, or NF.
272 Ceramic Membranes
The ceramic MF membrane is the first to be used successfully in industrial manufacture and commercial applications. Due to
their various advantages, which are important for the treatment and filtration of foods and beverages, most ceramic membranes
were used for the MF of milk, as well as prefiltration of wines and juices. After successful application in the dairy and wine
industries, the application of ceramic MF membranes gradually extended to the food industry, environmental engineering,
biological engineering, electronics industry, purification, and other fields. Commonly used materials for the separation layer
include Al2O3, ZrO2. and TiO2, supported on a porous a-Al2O3 substrate or multilayer composite asymmetric support. There are
currently dozens of vendors in the world that provide such ceramic membrane components, mainly in the United States, France,
Germany, Japan, and China (Table 1).
The ceramic UF membrane has a pore size of generally between 2 and 50 nm. They provide enhanced separation performance,
including high-separation precision and resistance to fouling, due to their smaller pore size and narrow distribution compared to
ceramic MF membranes. In the early stage, the UF membranes were also fabricated by a solid-state sintering process, but using
nanoparticles or nanofibers with smaller particle sizes. The typical applications for ceramic UF membranes are oily wastewater
treatment, beverage and juice purification, biological fermentation treatment and organic solvents filtration. To fabricate UF
membranes with ultrafine pore sizes (2–10 nm), the sol–gel technique starting from the colloidal or polymeric sol was introduced
and developed. Leenaars et al.4 applied the sol–gel technique successfully to prepare the Al2O3 membrane, and the AlOOH sol
synthesis processes and membrane permeation and retention performance were reviewed in detail in a series of papers. Subse-
quently, researchers in the United States, Japan, Germany, France, China, and other countries have been carrying out studies about
Ceramic Membranes 273
this technology. The sol–gel approach has a lot of advantages over other methods of forming UF membranes. By means of sol–gel
processing, a wide variety of chemical compositions and large range of pore sizes (depend on the particle size of sol) can be
achieved. Also, the precursor materials are inexpensive and the processing costs are low as well. Currently, UF g-alumina, titania, or
zirconia membranes are already commercialized, basically by the sol–gel method, and are used for applications such as separation
of dye particles from water and to create intermediate or support layers for other membranes.
Ceramic NF has a molecular weight cutoff (MWCO) between 200 and 1000 Da. There are two types of solutes separated by
different mechanisms: ions based chiefly on their valence state in water (if solutes charged) and sieving based on molecular weight
(if solutes uncharged).5 With the development of nanotechnology, NF membranes have gradually assumed an important position
in a variety of processing applications, including the petrochemical, food processing, wastewater treatment, energy, and medical
industries, taking the place of traditional small molecule separation methods and achieving sustainable development of generic
technology. Ceramic NF membranes are fabricated in a large number of different ceramic materials, such as g-Al2O3, TiO2, ZrO2,
SiO2, TiO2–ZrO2, and SiO2–ZrO2. In recent years, ceramic NF membranes are moving from the laboratory research to industrial-
scale products, though only a small amount of ceramic NF membranes were applied in the industrial projects. The development of
preparation methods and optimization of membrane structures and performance for ceramic NF membranes were reviewed in the
latter of this article.
Critical to advancement in ceramic membrane technology are high membrane permeability and selectivity, and the use of
inexpensive and controllable fabrication methods. During the membrane processes, the permeability and selectivity are related to
the materials; microstructures such as pore size and distribution, porosity, and membrane thickness; and other surface properties.
By research into the quantitative relationships between controlling parameters and membrane microstructure, some mathematical
models for the process of membrane preparation have been found and the quantity control of membrane preparation can be
achieved.6 On the other hand, the surface properties of ceramic membranes play a significant role in their separation performance
and fouling tendency.7,8 It has been well known that a modified membrane with special surface properties promotes the ceramic
membrane applications in separating water-in-oil emulsions, oil purification, solvent recovery and protein purification.
Combining high performance in terms of flux stability, separation efficiency, and long lifetime with the continued decrease in
fabrication cost as market volumes increase, ceramic membranes are becoming competitive now and could even cause a radical
shift in the membrane market over the next decade. BlueTech Research has modeled ceramic membrane UF adoption for water and
wastewater applications over the previous 20 years, as illustrated in Fig. 4. Adoption of ceramic membranes has been fairly modest,
with an estimated 200 þ treatment plants representing a population equivalent of roughly 5 million people utilizing the tech-
nology over the past 20 years. The past 5 years have seen a small uptick in the application of ceramic membranes for treatment,
with two large-scale plants developed by PWN and Metawater (4100,000 m3/d) opening in 2014. Up to now, more than 10,000
items of ceramic membrane equipment have been established in the world, including about 2000 by Jiangsu Jiuwu High-Tech
Co. Ltd. in China.
An overview of membrane structure, preparation methods, and current applications is given in the following sections of
this article.
Fig. 4 Ceramic membrane UF adoption for water and wastewater applications over the previous 20 years.
274 Ceramic Membranes
Membrane preparation is a core issue in the fields of membrane science and technology. Many scientists focused their attention on
this subject, while numerous engineers have developed new equipment to improve the technology of membrane preparation. In
the past several decades, fabrication technologies of ceramic membrane have made significant progress. Some new methods, such
as the pore-forming agent method,9 template method,10 fiber construction method,11 sol–gel method,12 and chemical deposition
method, as well as low-cost fabrication methods, have been successfully applied for the preparation of ceramic membranes to
enhance the permeability, selectivity, and separation precision, as well as antifouling properties. For example, to obtain highly
permeable ceramic membranes, pore forming and fiber construction methods are proposed. The pore forming approach is to
enlarge the number of pores and to increase the porosity of ceramic membranes by adding pore-forming agents. The template
method is a special type of the pore-forming agent method, in which the agent can create pores with specific sizes, as well as
ordered shapes. The fiber construction method uses ceramic fiber as raw material set up through the layers of fiber channels to
diversify the hole shapes, with the goal of improving porosity.
To develop ceramic membranes with smaller pores and higher separation precision for the NF process, the sol–gel technique is
currently the principal method used to prepare mesoporous and ceramic membrane. It usually employs metal alkoxides as a
precursor to forming polymeric or colloidal sols, which contain numerous nanosized particles. In addition, some chemical
deposition methods are also applied to improve the separation properties of ceramic membranes by decreasing pore size, patching
the defects, and changing the surface properties. Last but not least, to satisfy the requirements for commercial applications of
ceramic membranes, some fabrication methods are used that employ low-cost raw materials, a low-cost sintering process, and
some advanced fabrication technology. In this article, these methods will be discussed extensively.
distribution. The template method has attractive characteristics, such as variability in the selection of template agents and
flexibility in the adjusting of pore structure. Hence, the template method has attracted a good deal of interest from researchers in
recent years.
Fig. 5 SEM images (A, top view; B, cross-sectional view) of 3D ordered macroporous membranes.
Fig. 6 SEM images of 3D ordered macroporous Al2O3 membranes, using PMMA as the template.
276 Ceramic Membranes
membrane on porous substrate via a spin-coating process. As prepared, the composite membrane showed an average pore size of
about 200–300 nm.
The sol–gel method controls the separation precision of membrane layers, mainly by adjusting the particle size of colloids in
sols. The colloid size could be controlled to be as small as several nanometers, so as to prepare UF and NF ceramic membranes
with small pore size, narrow pore size distribution, and high separation precision.
Fig. 8 Pure water permeability (a) and PEG retention (b) of the YSZ NF membrane.
Fig. 9 Field-emission SEM images of TiO2–ZrO2 NF membrane: (a) surface and (b) cross section.
Subsequently, the crack-free thin g-Al2O3,53 YSZ,54 ZrO2,55 and TiO2–ZrO256 NF membranes were prepared by a colloidal sol–gel
route. The YSZ NF membrane exhibited great water permeability of 28 L/(m h bar) and a low MWCO of 800 Da (Fig. 8).54 When the
filtration was operated at 601C, the fluorescent brightener recovery value of 99% was obtained, while NaCl removal was greater than
98%, indicating that ceramic NF membranes were the technology of choice for the treatment of high-salinity dye wastewater.
The combined utilization of different metal oxides to fabricate composite ceramic membranes has attracted more and more
interest in increasing the diversity and performance of NF membranes. Tsuru et al.57 prepared a series of SiO2–ZrO2 colloidal
composite sols with various particle size distributions. Based on these composite sols, they obtained SiO2–ZrO2 composite NF
membranes with pore sizes in the order of 2.9 nm, 1.6 nm, and 1.0 nm, which decreased with the particle size of colloidal
composite sols. Lu et al.56 fabricated TiO2–ZrO2 composite NF membranes via a modified colloidal sol–gel process using
a mixture of Zr-inorganic salts and Ti-alkoxides. The FESEM images of the TiO2–ZrO2 NF membrane is shown in Fig. 9.
Ceramic Membranes 279
It demonstrated that a well-defined and crack-free top layer with a thickness of about 200 nm was formed. The composite NF
membrane exhibited a low MWCO of about 500 Da and a high pure water permeability of 35–40 L/(m h bar) and showed desired
retention rates of 99.6% for 1 ppm Co2 þ , 99.2% for 13 ppm Sr2 þ , and 75.5% for 7 ppm Cs þ in simulative radioactive wastewater
treatment (pH 3, T ¼ 251C), suggesting a great potential for ceramic NF membranes in radioactive effluent disposal.
stable permeability and microstructure after ultrasonic treatment. In addition, during the filtration of the CaCO3 suspension, the
membrane treated by cosintering had a higher steady permeate flux than the membrane prepared by the conventional process and
can withstand a high backflushing pressure.
In a subsequent study,90 a cosintering process was applied to fabricate bilayer ZrO2/Al2O3 membranes on Al2O3 substrates. It
was found that good bonding between the membrane layers and the support after cosintering could be achieved by using thinner
sublayer membrane without defects. However, the top-layer membrane should be not very thin. Otherwise, it would be insuf-
ficient to promote the sintering of the sublayer membrane during the cosintering process due to the compressive stress caused by
fast shrinkage of the top-layer membrane. The thicknesses for the sublayer and the top-layer membranes were optimized as 15 and
10 mm, respectively. The bilayer ZrO2/Al2O3 membrane with an average pore size of about 0.28 mm and a pure water flux of about
2.82 103 L/(m2 h bar) was obtained after being cosintered at 12001C.
To use the cosintering method for the fabrication of ceramic membranes with smaller pores, Qiu et al.36 prepared titania UF
membranes by using sol-coated nanofibers as the material for the sublayer during the cosintering process. In this respect, titania
nanofibers were applied to cover the porous substrate to produce a uniform layer with high porosity and flux. Meanwhile, titania
nanoparticles from sol were used to improve the mechanical strength of the titania nanofiber membrane. The enhance mechanism
could be explained by the formation of sintering neck between nanofibers with colloidal particles (sol) at lower sintering
temperatures. Subsequently, titania colloidal particulate sol is deposited on the top of the titania nanofiber layer and followed by a
cosintering process at a suitable sintering temperature of 4801C to reduce the pore size and achieve high separation efficiency. As
prepared, the UF membrane has a homogeneous surface without obvious defects. Meanwhile, the UF membrane showed high
pure water flux permeability of about 1100 L/(m2 h bar), while the MWCO is 32,000. Dong et al.91 prepared porous cordierite MF
membranes by a cosintering process. Meanwhile, the low-cost, industrial-grade powder was used as the raw material to reduce
fabrication costs and process time. The results demonstrated that the cosintering of two-layer cordierite membranes (45.0 and 15.0
mm thick) was feasible due to their low difference in sintering shrinkage behavior.
Application-oriented consideration for the fabrication of ceramic membranes mainly includes the designs of microstructure,
material properties, and operation parameters for ceramic membranes. The separation performance of ceramic membranes closely
depends on their structure parameters and material properties. The structure parameters include average pore size, pore size
distribution, membrane thickness, porosity, pore shape, and tortuosity, which are the main factors in the permeability of ceramic
membranes. The material properties include chemical stability, thermal stability, mechanical strength, and surface properties,
which not only affect the permeability of the membrane separation performance but also have a close relationship with membrane
lifespan. The operation parameters include flow velocity on membrane surface, TMP, and temperature, which affect the degree of
concentration polarization, membrane pollution, and permeability and rejection properties.
Fig. 10 Effect of pore size on the permeability of ceramic membrane in the filtration of ovalbumin and dextran solutions.
The ceramic membranes with high porosity mean having more open pores and high permeability. Generally, the porosity of
membrane layers is about 20–60%, while the porosity of substrates is much larger than that of membrane layers. For MF
membranes, the porosity is usually larger than 30%.97
Considering the abovementioned phenomenon, some researchers have established theoretical relationships between structures
and performance, so as to build application-oriented structure design methods for ceramic membranes. Xu et al.98–100 proposed a
research framework of application-oriented ceramic membrane design. A new, crossflow MF model of particle suspensions was
built, in which a blocking factor (k) was proposed to describe membrane fouling.98 The correlation between membrane per-
meability and membrane structure parameters was described in this model. Furthermore, the variation of permeate flux with
filtration time was obtained, and the influences of membrane structure parameters such as pore size, thickness, and porosity on
permeate flux could be theoretically predicted under the ideal conditions through this model. The simulated results showed that
there was an optimum pore size in the filtration of suspensions to obtain maximal permeate flux. Meanwhile, the optimum pore
size was influenced by the size and size distribution of the particulates.99 Finally, the novel membrane design and application
methods were well verified in the nanosuspensions system of titania, where the model calculation showed good agreement with
the experimental results.100
Fig. 11 Time dependence of the contact angle for Al2O3 and Al2O3–TiO2 membrane.
Fig. 12 Variation of the permeate flux for Al2O3 and Al2O3–TiO2 membranes versus filtration time.
more stable flux was observed through the Al2O3–TiO2 composite membrane (Fig. 12), owing to more hydrophilic groups, to
reduce the adsorption of oil droplets. Moreover, the composite membrane had the same type of charge (i.e., a negative charge)
with oil droplets, which benefits the control of membrane fouling during filtration of oily emulsion wastewater.
On the other hand, surface hydrophobic modification of ceramic membranes was accomplished with a self-assembled
monolayer of organosilane.107,108 The time dependence of the variation in contact angles for the ZrO2 membrane and the
modified ZrO2 membrane are shown in Fig. 13. The contact angle of the original ZrO2 membrane is less than 17 degrees and
decreased as time passes. However, the modified ZrO2 membrane has a much larger contact angle of about 134 degrees, and the
contact angle stays stable. The result proves that this modification can endow the ZrO2 membrane surface with much better
hydrophobicity, which means a better fouling resistance in the oil system. As shown in Fig. 14, the water rejection of the modified
membranes was always greater than that of the unmodified membranes. For the unmodified membranes, the water rejection was
around 88%, and there was a slight decrease when the TMP increased from 0.04 to 0.13 MPa, while for the modified membranes,
the water rejection was around 98%, and there was no clear difference between the water rejections when the TMP increased.
Fig. 13 Time dependence of contact angles for the ZrO2 membranes and the modified ZrO2 membranes.
Fig. 14 Comparison of water rejections between the ZrO2 membranes and the modified ZrO2 membranes in the filtration of water-in-kerosene
emulsions at different TMP values.
systems always performed at minimum permeability during the separation of proteins, microorganisms, and macromolecules
when the pH value of solutions was around the IEP. However, for the systems with inorganic particles, the permeability of
membranes increased when the pH value of solutions was around the IEP due to aggregation of inorganic particles. In addition,
the change of pH values in solution would also affect the surface charge of membranes. In general, ceramic membranes
are negatively charged in neutral aqueous solution.109 Therefore, the selectivity of proteins for mixtures could be improved
dramatically by adjusting the zeta potential of membranes.110–113 Meanwhile, the change in zeta potential of membranes could
also affect electroviscous effects when fluid passes through membrane pores, consequently affecting the permeability of
membranes.114,115
Inorganic ions in the solution also have a significant impact on membrane separation performance, especially in the treatment
of protein systems. On the one hand, some inorganic salt composites could directly absorb and deposit on the membrane surface
and in membrane pores, or increase the absorbance of protein onto the membrane surface. On the other hand, inorganic salts
affect the ionic strength, which have strong effects on the solvability and structure of proteins and even the density of protein-
fouling cakes. As a result, these changes caused by inorganic ions significantly affect the permeability of ceramic membranes.
Alventosa et al.116 investigated the UF of model solutions containing NaCl and a reactive dye using ceramic membranes.
Higher flux decline and total resistance, together with lower color removal, were observed when adding NaCl to the solution,
which affected the repulsion–attraction phenomenon by changing the membrane charge.
Yoon et al.117 studied the fouling mechanism of inorganic membranes in the system of the membrane-coupled anaerobic
bioreactor (MCAB) by using alcohol distillery wastewater as a digester feed. The results demonstrated that the fouling mechanism
of inorganic membranes was significantly different from that of conventional membrane filtration processes. The inorganic
Ceramic Membranes 285
precipitate, struvite (MgNH4PO4 6H2O), rather than anaerobic microbial floes, was found to be the main foulant. In this process,
the struvite appears to be precipitated not only on the membrane surface, but also inside the membrane pores, and it was not
easily removed by general physical cleaning. Therefore, chemical cleaning by acid was still necessary to recover the membrane flux.
Hoogland et al.118 studied the effect of pH on the permeability of alumina membrane, for both pure water and the filtration of
mineral slurries. The maximum pure water flux was found in a pH value below 2, which equaled the IEP value of membrane. In
contrast, the membrane resistance was greatest at high pH, which was far from the IEP. This could be explained by an electro-
osmosis phenomenon at high pH. However, higher permeabilities for alumina membranes were obtained at both low and high
pH in the filtration of mineral slurries (silica particles). This is because at high pH, both the membrane and the particles exhibit
negative charges, which lead to repulsive forces at the membrane-solution interface and depolarization of the membrane.
Moosemiller et al.119 found that both the supported g-Al2O3 and TiO2 ceramic membranes showed maximum permeate flux
for particle-free electrolyte solution when the pH value was around the IEP. Nazzal et al.120 studied the effects of pH and ionic
strength on the performance of ceramic MF membranes. However, the maximum permeate flux was not found around the IEP.
Meanwhile, the flux decreased as ionic strength increased. The results suggested that counter-ions exert an influence on the
permeate flux at distances substantially greater than a Debye length from the pore walls. Huisman et al.121 found that the increases
in water fluxes of 2–8% for both MF and UF membranes were caused by the increasing of salt concentration from 30 mM to 0.1 M.
These results were explained by electroviscous effects: increased salt concentrations led to lower zeta-potentials and thinner double
layers, offering less resistance to water passage.
Kwon et al.122 studied the effect of particle size and ionic strength of the feed suspension on critical flux. It was demonstrated
that the ionic strength of suspension had a significant effect on the critical flux. For an ionic strength less than 1 101.5 M, there
was a decrease in the critical flux. This could be due to the dense layer of deposit, which is the result of a less diffuse layer thickness
of particles. Above this ionic strength, a significant increase in critical flux was noticed, which may be due to the aggregation of
particles.
Chevereau et al.123 studied the effect of the presence of single salts in solution on the behavior of UF TiO2 membranes. It was
found that the IEP of TiO2 remains stable in the presence of NaCl, while it shifts to a higher pH value in the presence of MgCl2. At
native pH (close to 7), the adsorption of Mg2 þ on TiO2 surface leads to an inversion of the surface charge, as well as to an
important increase of the surface charge density.
Elzo et al.101 studied the effects of pH, ionic strength, and salts on the crossflow MF performance of 0.5 mm silica particles
using an inorganic membrane of 0.2 mm pore size. High permeate fluxes are obtained at high pH and low salt concentrations. In
contrast, low filtration fluxes are measured to have a high salt concentration, low pH, and a CaCl2 electrolyte. These results were
explained by the repulsion between the silica particles, which could be affected by the conditions.
Chiu124 studied the influence of time or aging on the electrokinetic characteristics of ceramic MF and their subsequent impact
on the MF performance of whey suspension. It was found that the zeta potential of the membrane increases as the aging period
increases, and then plateaus toward the end of the aging period. Consequently, a higher permeate flux was found over the aging
period.
Huisman et al.125 investigated the critical flux of MF while filtering a suspension of silica particles. It was found that neither the
zeta potential of the membrane nor the zeta potential of the particles influenced the observed critical flux. The critical flux
increased linearly with the wall shear stress and decreased with increasing particle concentration. Two different particle transport
mechanisms, particle rolling (the torque-balance model) and shear-induced diffusion, were applied for the expression of
critical flux.
Due to the complexity of the solution system, there is still no uniform law for the effect of solution properties on the
performance of membrane processes. There are even some conflicting results. Therefore, further studies should be carried out
focusing on these issues.
measuring it at a high crossflow velocity for a laminar flow. Consequently, they obtained the critical flux at a crossflow velocity of
0.015 m/s, which is slightly less than 185.4 L/(m2 h).
In general, the viscosity of solution decreases with the increase of temperature. Meanwhile, the solubility and mass transfer
coefficient increase with temperature. These events could accelerate the diffusion of solutes from the membrane surface to the body
of fluid, so as to reduce the thickness of the concentration polarization layer. As a result, the membrane permeate flux and
separation efficiency improved. For a dilute solution, the relationship between the permeate flux and temperature could be
predicted by the relationship between viscosity and temperature. Chang et al.130 found that the permeate flux almost doubled over
its original value when the temperature increased from 201C to 601C in the separation of a stable oil-in-water emulsion. Da et al.54
studied the effect of temperature on the performance of ceramic NF membranes in the treatment of dye wastewater. The flux value
was about 44 L/(m2 h) at 301C and increased to 84 L/(m2 h) at 601C. However, higher temperatures inescapably caused a decrease
in solute retention. This observation was explained by two reasons: (i) higher temperatures accelerated the fluorescent brightener
diffusion through the membrane; and (ii) the layer of adsorbed water molecules on the pore walls would be thinner at higher
temperatures, decreasing the hydrodynamic drag forces inside the pores as the effective pore diameter increased with the
temperature.
The unique thermal stability, chemical resistance, and mechanical properties of ceramic membranes give them significant
advantages and make them suitable for application in hash environments. Typical applications of ceramic membranes include
biopharmaceuticals, chemicals, electronics, food and beverage, and effluents. In this article, five industrial applications, including
ceramic membrane reactors, oily wastewater treatments, biological fermentation treatments, plant extracts, and brine refinements,
are introduced.
Fig. 15 The industrialization apparatus of ceramic membrane reactors. Photo courtesy of Jiuwu.
ceramic membrane reactor system could be continually operated for more than 6 months without membrane cleaning. Conse-
quently, some industrial applications of ceramic membrane reactors have ever been realized. Fig. 15 shows an industrialization
apparatus of ceramic membrane reactors. Next, two industrial application examples for heterogeneous catalysis are presented:
cyclohexanone oxime production from the ammoxidation of cyclohexanone over an ultrafine TS-1 catalyst; and p-aminophenol
production from p-nitrophenol hydrogenation over nanosized nickel.
However, the high purity requirements for discharge are hard to reach by using a single membrane process due to the much
stricter requirements of resource conservation and environmental protection. Therefore, the integrated process has been considered
an ideal method for the treatment of oily wastewater.149 Abbasi et al.150 investigated the treatment of oily wastewater by using a
hybrid coagulation-MF process. Four coagulants (i.e., ferrous chloride, ferrous sulfate, aluminum chloride, and aluminum sulfate)
with the addition of calcium hydroxide were evaluated in the coagulation-MF hybrid process at different concentrations. In
addition, they also studied the hybrid MF-powdered activated carbon process.151 In the two hybrid processes, Hermia's models
were applied to study the fouling mechanisms of mullite-alumina ceramic membranes in the treatment of oily wastewater. It was
found that the cake filtration model can be applied to predict permeation flux decline. Nidal et al.152 achieved the treatment of
metal working fluids and assessed the feasibility of permeate reuse by combining UF and NF membrane processes. Maria et al.153
investigated the feasibility of bilge water treatment in the integrated UF and reverse osmosis system. It was found that the permeate
from the first stage of bilge water treatment had oil content below 10 mg/L and was free of suspended solids, and the obtained
reverse osmosis permeate was free of oil. Yu et al.154 investigated the possibility and feasibility of vegetable oily wastewater
treatment with a process that integrated MF and reverse osmosis. The operating conditions for permeate flux and rejection were
optimized in the MF process. The operating temperature should be kept in the range of 30–401C, while the proper crossflow
velocity and the TMP should be kept at about 4 m/s, and 0.15 MPa, respectively. The results demonstrated that the two stage
membranes combining MF and reverse osmosis constitute the most feasible and reliable process for vegetable oily wastewater
treatment.
Fig. 16 Ceramic membrane sets for the treatment of oily wastewater. Photo courtesy of Jiuwu.
Ceramic Membranes 289
To improve the purity of HL, activated-carbon adsorption and ion exchange processes were carried out. Finally, HL was con-
centrated by removing the water through the multieffect evaporation method. In this traditional production process, a mass of
byproduct (gypsum) was produced at a high rate (1 ton per ton of HL).158
To reduce the industry waste residue of gypsum, NaOH or NH4OH is added as a neutralizer to adjust the pH and avoid the
inhibition of cell growth. However, the recovery of high-purity product from the fermentation broth becomes a critical technical
problem by using the traditional production process. Therefore, numerous studies have been carried out to propose new
separation techniques for HL recovery from the fermentation broth, such as extraction,159 adsorption,160 and membrane
separation.161 Over the past two decades, membrane techniques have been extensively investigated because they can be integrated
easily with conventional fermenters.162 The crossflow ceramic MF membrane can retain cells and proteins more efficiently than
plate-and-frame filtration in the conventional process.163
Pal et al.164 reviewed recent developments in the production of monomer-grade HL by using membrane-based processes. They
suggested an integrated production scheme using MF and NF membranes. However, this continuous fermentative HL production
scheme with two-stage integration of membranes cannot ensure the separation of microbial cells for their recycling without
additional provision of cells. To overcome this problem, a promising method is to combine the membrane process with batch
fermentation. Therefore, efforts should be directed toward developing a membrane-integrated process for HL production.
Wang et al.165 investigated the feasibility of an integrated ceramic MF/UF, NF, and bipolar membrane electrodialysis (BMED)
process followed by fermentation broth neutralization with sodium hydroxide. The scheme of the setup is presented in Fig. 17. A
ceramic membrane was used to remove the cells. The permeate from the ceramic membrane was then introduced to an NF process
in which the sugar, protein, and bivalent salts were separated from sodium lactate. The BMED process was applied to the
transformation of sodium lactate to HL and sodium hydroxide. Sodium hydroxide was circulated to the fermentation system to
adjust the pH. Based on this study, a pilot-scale HL-producing process was successfully carried out by Jiangsu Jiuwu High-Tech Co.
Ltd., with an annual production of 1000 t/a.
Fig. 17 Scheme of the integrated membrane process setup: (1) fermenter; (2) UF feed tank; (3) ceramic membrane module; (4) NF feed tank; (5)
NF membrane module; (6) acid tank; (7) BMED stack; (8) caustic tank; (9–13) pump, (P1–P4) pressure gauges, (T1–T4) temperature gauges,
(F1–F4) flow meters.
290 Ceramic Membranes
Fig. 18 Industrial ceramic membrane separation system for the production of cephalosporins. Photo courtesy of Jiuwu.
Table 2 Comparison of ceramic membrane system with plate and frame filter systems in the
treatments of biological fermentation
the benefits of enterprise. Generally, the active ingredients in plant extraction solutions come in low levels. Meanwhile, the
extraction solutions often contain impurities, including starches, proteins, carbohydrates, pigments, and pectin. The molecular
weight of active ingredients usually ranges from several hundred thousand to millions of Daltons.
Nowadays, various types of membrane separation technology has been widely used in the separation and purification,
decolorization, desalination, and concentration of plant extracts.166,167 In these processes, a complete set of integration membrane
technologies was usually applied to remove the macromolecular substances effectively. Many advantages have been shown,
including avoiding the backmixing phenomenon, improving transmittance, decreasing the salt content in the product, relieving
the burden of next stage (decolorization and desalination), and even the replacement of solvents.
Fig. 19 Flow sheet of the production process of anthocyanin by integrated membrane technology.
Fig. 20 Ceramic membrane equipment for an anthocyanin project. Photo courtesy of Jiuwu.
Fig. 21 Ceramic membrane equipment for a liqueur project. Photo courtesy of Jiuwu.
with the cleaning solution. In addition, a dissolution kinetics model associated with both the concentration factor and
the temperature factor was developed, which fitted well with the experimental results. Based on this model, it was found that
the activation energy of BaSO4 dissolution using the cleaning solution consisted of DTPA, oxalic acid, and NaOH was less
than that of using pure DTPA solution, which means the compound solution provided a better cleaning performance than
pure DTPA solution. The flow sheet of the ceramic membrane system for brine purification running in a chloralkali plant is shown
in Fig. 22.
Until now, dozens of industrial devices for the brine refinements have been constructed and successfully been in
smooth operation in the chloralkali enterprises in China. Several industrial devices with different handling capacities are shown
in Figs. 23–25.
Fig. 22 Flow sheet of the ceramic membrane system for brine purification.
Fig. 23 Brine refinement device for the production of caustic soda (10,0000 t/a). Photo courtesy of Jiuwu.
Fig. 24 Brine refinement device for the production of caustic soda (15,0000 t/a). Photo courtesy of Jiuwu.
Ceramic Membranes 293
Fig. 25 Brine refinement device for the production of caustic soda (30,0000 t/a). Photo courtesy of Jiuwu.
The study of the fabrication and application of ceramic membranes has stretched over the past 40 years. The research and
development has been supported by the government and some large companies. Remarkable improvements have been performed
in the development of application-oriented structure design methods that achieve a breakthrough ceramic membrane preparation
technology and promote the industrialization of ceramic membrane products. Although many novel fabrication techniques have
been developed to improve the separation performance and decrease the cost of ceramic membranes, further research is required
to produce membranes with a wide range of unique microstructures and surface properties, which can be adapted for the
application and increase their market volumes.
The emphases of future development regarding ceramic membrane fabrication and application will cover the following aspects:
(1) to develop a low-cost, high-performance ceramic membrane element based on the application-oriented design process and
preparation method; (2) to develop ceramic NF membranes for molecular level separation from lab scale to industrial scale and
open the window to extend their application fields; (3) to enrich surface properties of the ceramic membranes to improve their
fouling resistance and extend their lifetime in practical applications; (4) to strengthen the basic theoretical study of the transport
mechanism in confined nanochannels of ceramic membranes.
Acknowledgments
The authors would like to acknowledge the financial support from the National High Technical Research Program of China
(2007aa030303, 2012AA03A606), the National Basic Research Program of China (2003CB61570700, 2009CB623400), the
National Nature Science Foundation of China (21490580, 91534108), the Project of Priority Academic Program Development
of Jiangsu Higher Education Institutions (PAPD), and the program for Changjiang Scholars and Innovative Research Team in
University (IRT13070). The authors would also like to thank Jiangsu Jiuwu Hi-Tech Corporation and Dr. Wenbo Peng for valuable
contributions on the information of industrial applications.
See also: 1.12 Microstructured Ceramic Hollow Fiber Membranes and Their Applications
References
1. Burggraaf, A. J.; Cot, L. Fundamentals of Inorganic Membrane Science and Technology. Elsevier 1996.
2. Rao, L. J. M. Handbook of Membrane Separations; Wiley, 2009.
3. Van Gestel, T.; Vandecasteele, C.; Buekenhoudt, A.; et al. Alumina and Titania Multilayer Membranes for Nanofiltration: Preparation, Characterization and Chemical
Stability. J. Membr. Sci. 2002, 207 (1), 73–89.
4. Leenaars, A. F. M.; Keizer, K.; Burggraaf, A. J. The Preparation and Characterization of Alumina Membranes With Ultrafine Pores. J. Mater. Sci. 1984, 19,
1077–1088.
5. Hilal, N.; Al-Zoubi, H.; Darwish, N. A.; et al. A Comprehensive Review of Nanofiltration Membranes: Treatment, Pretreatment, Modelling, and Atomic Force Microscopy.
Desalination 2004, 170 (3), 281–308.
6. Drioli, E.; Giorno, L. Comprehensive Membrane Science and Engineering; Elsevier, 2010.
7. Su, C. H.; Xu, Y. Q.; Zhang, W.; et al. Porous Ceramic Membrane With Superhydrophobic and Superoleophilic Surface for Reclaiming Oil From Oily Water. Appl. Surf.
Sci. 2012, 258 (7), 2319–2323.
294 Ceramic Membranes
8. Moritz, T.; Benfer, S.; Arki, P.; et al. Influence of the Surface Charge on the Permeate Flux in the Dead-End Filtration With Ceramic Membranes. Sep. Purif. Technol.
2001, 25 (1–3), 501–508.
9. Liu, S.; Zeng, Y.-P.; Jiang, D. Fabrication and Characterization of Cordierite-Bonded Porous SiC Ceramics. Ceram. Int. 2009, 35 (2), 597–602.
10. Velev, O. D.; Jede, T. A.; Lobo, R. F.; et al. Microstructured Porous Silica Obtained via Colloidal Crystal Templates. Chem. Mater. 1998, 10 (11), 3597–3602.
11. Ke, X.; Zhu, H.; Gao, X.; et al. High-Performance Ceramic Membranes With a Separation Layer of Metal Oxide Nanofibers. Adv. Mater. 2007, 19 (6), 785–790.
12. Das, N.; Maiti, H. S. Ceramic Membrane by Tape Casting and Sol–Gel Coating for Microfiltration and Ultrafiltration Application. J. Phys. Chem. Solids 2009, 70 (11),
1395–1400.
13. Yao, A.; Yu, B.; Yang, K.; et al. Fabrication and Properties of Mullite-Alumina Ceramic Support. Rare Metal Mater. Eng. 2005, 34, 255–258.
14. Dong, G.; Qi, H.; Xu, N. Effect of Active Carbon Doping on Structure and Property of Porous Alumina Support. J. Chin. Ceram. Soc. 2012, 40 (6), 844–850.
15. Gregorová, E.; Pabst, W. Porosity and Pore Size Control in Starch Consolidation Casting of Oxide Ceramics—Achievements and Problems. J. Eur. Ceram. Soc. 2007,
27 (2–3), 669–672.
16. She, J.; Deng, Z.; Daniel-Doni, J.; et al. Oxidation Bonding of Porous Silicon Carbide Ceramics. J. Mater. Sci. 2002, 37 (17), 3615–3622.
17. Collier, A. K.; Liu, W.; Wang, J. G.; et al. Alpha-Alumina Inorganic Membrane Support and Method of Making the Same. US patent 20110045971, 2011.
18. Yang, G. C. C.; Tsai, C.-M. Effects of Starch Addition on Characteristics of Tubular Porous Ceramic Membrane Substrates. Desalination 2008, 233 (1–3), 129–136.
19. Park, S. H.; Xia, Y. N. Macroporous Membranes With Highly Ordered and Three-Dimensionally Interconnected Spherical Pores. Adv. Mater. 1998, 10 (13), 1045–1048.
20. Sadakane, M.; Horiuchi, T.; Kato, N.; et al. Facile Preparation of Three-Dimensionally Ordered Macroporous Alumina, Iron Oxide, Chromium Oxide, Manganese Oxide, and
Their Mixed-Metal Oxides With High Porosity. Chem. Mater. 2007, 19 (23), 5779–5785.
21. Zhao, K.; Fan, Y.; Xu, N. Preparation of Three-Dimensionally Ordered Macroporous SiO2 Membranes With Controllable Pore Size. Chem. Lett. 2007, 36 (3), 464–465.
22. Zhao, K.; Fan, Y.; Wang, R.; et al. Preparation of Closed Macroporous Al2O3 Membranes With a Three-Dimensionally Ordered Structure. Chem. Lett. 2008, 37 (4),
420–421.
23. Xu, J.; Xiang, W.; Hu, F. Preparation of Monodisperse Polystyrene Spheres and Inorganic Porous Films. Rare Metal Mater. Eng. 2008, 37, 196–200.
24. Beck, J. S.; Vartuli, J. C.; Roth, W. J.; et al. A New Family of Mesoporous Molecular-Sieves Prepared With Liquid-Crystal Templates. J. Am. Chem. Soc. 1992,
114 (27), 10834–10843.
25. Ji, H.; Fan, Y.; Jin, W.; et al. Synthesis of Si-MCM-48 Membrane by Solvent Extraction of the Surfactant Template. J. Non-Cryst. Solids 2008, 354 (18), 2010–2016.
26. Xu, D. K.; Fan, Y. Q. Mesoporous Si-MCM-48 Membrane Prepared by Pore-Filling Method. Sci. China Technol. Sci. 2010, 53 (4), 1064–1068.
27. Kumar, P.; Ida, J.; Kim, S.; et al. Ordered Mesoporous Membranes: Effects of Support and Surfactant Removal Conditions on Membrane Quality. J. Membr. Sci. 2006,
279 (1–2), 539–547.
28. Zhang, J.; Li, W.; Meng, X.; et al. Synthesis of Mesoporous Silica Membranes Oriented by Self-Assembles of Surfactants. J. Membr. Sci. 2003, 222 (1–2), 219–224.
29. Choi, H.; Sofranko, A. C.; Dionysiou, D. D. Nanocrystalline TiO2 Photocatalytic Membranes With a Hierarchical Mesoporous Multilayer Structure: Synthesis,
Characterization, and Multifunction. Adv. Funct. Mater. 2006, 16 (8), 1067–1074.
30. Cao, X. P.; Jing, W. H.; Xing, W. H.; et al. Fabrication of a Visible-Light Response Mesoporous TiO2 Membrane With Superior Water Permeability via a Weak Alkaline
Sol–Gel Process. Chem. Commun. 2011, 47 (12), 3457–3459.
31. Lei, W.; Fan, Y. Effect of Sintering Temperature on Microfiltration Membrane Prepared With TiO2 Ceramic-Fiber. Membr. Sci. Technol. 2009, 29 (5), 54–57.
32. Ke, X.; Zheng, Z.; Liu, H.; et al. High-Flux Ceramic Membranes With a Nanomesh of Metal Oxide Nanofibers. J. Phys. Chem. B 2008, 112 (16), 5000–5006.
33. Ke, X.; Zheng, Z.; Zhu, H.; et al. Metal Oxide Nanofibres Membranes Assembled by Spin-Coating Method. Desalination 2009, 236 (1–3), 1–7.
34. Ke, X. B.; Huang, Y. M.; Dargaville, T. R.; et al. Modified Alumina Nanofiber Membranes for Protein Separation. Sep. Purif. Technol. 2013, 120, 239–244.
35. Fernando, J. A.; Chung, D. D. L. Improving an Alumina Fiber Filter Membrane for Hot Gas Filtration Using an Acid Phosphate Binder. J. Mater. Sci. 2001, 36 (21),
5079–5085.
36. Qiu, M.; Fan, S.; Cai, Y.; et al. Co-Sintering Synthesis of Bi-Layer Titania Ultrafiltration Membranes With Intermediate Layer of Sol-Coated Nanofibers. J. Membr. Sci.
2010, 365 (1–2), 225–231.
37. Pastila, P.; Helanti, V.; Nikkila, A. P.; et al. Environmental Effects on Microstructure and Strength of SiC-Based Hot Gas Filters. J. Eur. Ceram. Soc. 2001, 21 (9),
1261–1268.
38. Li, J.; Lin, H.; Li, J. Factors that Influence the Flexural Strength of SiC-Based Porous Ceramics Used for Hot Gas Filter Support. J. Eur. Ceram. Soc. 2011, 31 (5),
825–831.
39. Ding, S.; Zeng, Y.; Jiang, D. In-Situ Reaction Bonding of Porous SiC Ceramics. Mater. Charact. 2008, 59 (2), 140–143.
40. Vacassy, R.; Guizard, C.; Thoraval, V.; et al. Synthesis and Characterization of Microporous Zirconia Powders: Application in Nanofilters and Nanofiltration Characteristics.
J. Membr. Sci. 1997, 132 (1), 109–118.
41. Ju, X.; Huang, P.; Xu, N.; et al. Study of Factors Influencing Pore Size of Zirconia Ultrafiltration Membrane. J. Chem. Eng. Chin. Univ. 2000, 14 (2), 103–105.
42. Manjumol, K. A.; Shajesh, P.; Baiju, K. V.; et al. An ‘Eco-friendly’ All Aqueous Sol Gel Process for Multi Functional Ultrafiltration Membrane on Porous Tubular Alumina
Substrate. J. Membr. Sci. 2011, 375 (1–2), 134–140.
43. Fan, S.; Qiu, M.; Zhou, X.; et al. Preparation of Multichannel TiO2 Ultrafiltration Membrane and Its Application in Dyeing Wastewater. J. Nanjing Univ. Technol. 2010,
33 (1), 44–47.
44. Sakka, S. Handbook of Sol–Gel Science and Technology: Processing, Characterization and Applications; Kluwer Academic Publishers: Holland, 2005.
45. Tsuru, T.; Hironaka, D.; Yoshioka, T.; et al. Titania Membranes for Liquid Phase Separation: Effect of Surface Charge on Flux. Sep. Purif. Technol. 2001, 25 (1–3),
307–314.
46. Tsuru, T.; Ogawa, K.; Kanezashi, M.; et al. Permeation Characteristics of Electrolytes and Neutral Solutes Through Titania Nanofiltration Membranes at High Temperatures.
Langmuir 2010, 26 (13), 10897–10905.
47. Qi, H.; Li, S.; Jiang, X.; et al. Preparation and Ions Retention Properties of TiO2 Nanofiltration Membranes. J. Inorg. Mater. 2011, 26 (3), 305–310.
48. Van Gestel, T.; Sebold, D.; Kruidhof, H.; et al. ZrO2 and TiO2 Membranes for Nanofiltration and Pervaporation. J. Membr. Sci. 2008, 318 (1–2), 413–421.
49. Benfer, S.; Popp, U.; Richter, H.; et al. Development and Characterization of Ceramic Nanofiltration Membranes. Sep. Purif. Technol. 2001, 22–23, 231–237.
50. Spijksma, G. I.; Huiskes, C.; Benes, N. E.; et al. Microporous Zirconia-Titania Composite Membranes Derived From Diethanolamine-Modified Precursors. Adv. Mater.
2006, 18 (16), 2165–2168.
51. Aust, U.; Benfer, S.; Dietze, M.; et al. Development of Microporous Ceramic Membranes in the System TiO2/ZrO2. J. Membr. Sci. 2006, 281 (1–2), 463–471.
52. Cai, Y.; Wang, Y.; Chen, X.; et al. Modified Colloidal Sol–Gel Process for Fabrication of Titania Nanofiltration Membranes With Organic Additives. J. Membr. Sci. 2015,
476, 432–441.
53. Chen, X.; Zhang, W.; Lin, Y.; et al. Preparation of High-Flux g-Alumina Nanofiltration Membranes by Using a Modified Sol–Gel Method. Microporous Mesoporous Mater.
2015, 214, 195–203.
54. Da, X.; Wen, J.; Lu, Y.; et al. An Aqueous Sol–Gel Process for the Fabrication of High-Flux YSZ Nanofiltration Membranes as Applied to the Nanofiltration of Dye
Wastewater. Sep. Purif. Technol. 2015, 152, 37–45.
55. Da, X.; Chen, X.; Sun, B.; et al. Preparation of Zirconia Nanofiltration Membranes Through an Aqueous Sol–Gel Process Modified by Glycerol for the Treatment of
Wastewater With High Salinity. J. Membr. Sci. 2016, 504, 29–39.
56. Lu, Y.; Chen, T.; Chen, X.; et al. Fabrication of TiO2-Doped ZrO2 Nanofiltration Membranes by Using a Modified Colloidal Sol–Gel Process and Its Application in
Simulative Radioactive Effluent. J. Membr. Sci. 2016, 514, 476–486.
Ceramic Membranes 295
57. Tsuru, T.; Wada, S.; Izumi, S.; et al. Silica–Zirconia Membranes for Nanofiltration. J. Membr. Sci. 1998, 149 (1), 127–135.
58. Lin, Y. S. A Theoretical Analysis on Pore Size Change of Porous Ceramic Membranes After Modification. J. Membr. Sci. 1993, 79 (1), 55–64.
59. Xomeritakis, G.; Lin, Y. S. Chemical-Vapor-Deposition of Solid Oxides in Porous-Media for Ceramic Membrane Preparation—Comparison of Experimental Results With
Semianalytical Solutions. Ind. Eng. Chem. Res. 1994, 33 (11), 2607–2617.
60. Labropoulos, A. I.; Romanos, G. E.; Karanikolos, G. N.; et al. Comparative Study of the Rate and Locality of Silica Deposition During the CVD Treatment of Porous
Membranes With TEOS and TMOS. Microporous Mesoporous Mater. 2009, 120 (1–2), 177–185.
61. Lin, C. L.; Flowers, D. L.; Liu, P. K. T. Characterization of Ceramic Membranes. 2. Modified Commercial Membranes With Pore-Size Under 40 Angstrom. J. Membr. Sci.
1994, 92 (1), 45–58.
62. Fernandes, N. E.; Gavalas, G. R. Gas Transport in Porous Vycor Glass Subjected to Gradual Pore Narrowing. Chem. Eng. Sci. 1998, 53 (5), 1049–1058.
63. Sarrade, S.; Guizard, C.; Rios, G. M. New Applications of Supercritical Fluids and Supercritical Fluids Processes in Separation. Sep. Purif. Technol. 2003, 32 (1–3),
57–63.
64. Wang, Z.; Dong, J.; Xu, N.; et al. Pore Modification Using the Supercritical Solution Infiltration Method. AICHE J. 1997, 43 (9), 2359–2367.
65. Brasseur-Tilmant, J.; Chhor, K.; Jestin, P.; et al. Ceramic Membrane Elaboration Using Supercritical Fluid. Mater. Res. Bull. 1999, 34 (12–13), 2013–2025.
66. Tatsuda, N.; Fukushima, Y.; Wakayama, H. Penetration of Titanium Tetraisopropoxide Into Mesoporous Silica Using Supercritical Carbon Dioxide. Chem. Mater. 2004,
16 (9), 1799–1805.
67. Li, F.; Yang, Y.; Fan, Y.; et al. Modification of Ceramic Membranes for Pore Structure Tailoring: The Atomic Layer Deposition Route. J. Membr. Sci. 2012, 397, 17–23.
68. Van Gestel, T.; Van der Bruggen, B.; Buekenhoudt, A.; et al. Surface Modification of g-Al2O3/TiO2 Multilayer Membranes for Applications in Non-Polar Organic Solvents.
J. Membr. Sci. 2003, 224 (1–2), 3–10.
69. Singh, R. P.; Way, J. D.; Dec, S. F. Silane Modified Inorganic Membranes: Effects of Silane Surface Structure. J. Membr. Sci. 2005, 259 (1–2), 34–46.
70. Sah, A.; Castricum, H. L.; Bliek, A.; et al. Hydrophobic Modification of g-Alumina Membranes With Organochlorosilanes. J. Membr. Sci. 2004, 243 (1–2), 125–132.
71. Leger, C.; Lira, H. D. L.; Paterson, R. Preparation and Properties of Surface Modified Ceramic Membranes. Part II. Gas and Liquid Permeabilities of 5 nm Alumina
Membranes Modified by a Monolayer of Bound Polydimethylsiloxane (PDMS) Silicone Oil. J. Membr. Sci. 1996, 120 (1), 135–146.
72. Faibish, R. S.; Cohen, Y. Fouling-Resistant Ceramic-Supported Polymer Membranes for Ultrafiltration of Oil-in-Water Microemulsions. J. Membr. Sci. 2001, 185 (2),
129–143.
73. Rovira-Bru, M.; Giralt, F.; Cohen, Y. Protein Adsorption Onto Zirconia Modified With Terminally Grafted Polyvinylpyrrolidone. J. Colloid Interface Sci. 2001, 235 (1),
70–79.
74. Khemakhem, S.; Ben, Amar R. Grafting of Fluoroalkylsilanes on Microfiltration Tunisian Clay Membrane. Ceram. Int. 2011, 37 (8), 3323–3328.
75. Lin, Y. S.; Burggraaf, A. J. Experimental Studies on Pore Size Change of Porous Ceramic Membranes After Modification. J. Membr. Sci. 1993, 79 (1), 65–82.
76. Hu, J.; Qi, H.; Fan, Y.; et al. Porous Ceramic Support of Coated Alumina Prepared by Low-Temperature Sintering. J. Chin. Ceram. Soc. 2009, 37 (11), 1818–1823.
77. Falamaki, C.; Naimi, M.; Aghaie, A. Dual Behavior of CaCO3 as a Porosifier and Sintering Aid in the Manufacture of Alumina Membrane/Catalyst Supports. J. Eur. Ceram.
Soc. 2004, 24 (10–11), 3195–3201.
78. Wang, Y.; Zhang, Y.; Liu, X.; et al. Microstructure Control of Ceramic Membrane Support From Corundum-Rutile Powder Mixture. Powder Technol. 2006, 168 (3),
125–133.
79. Vasanth, D.; Pugazhenthi, G.; Uppaluri, R. Fabrication and Properties of Low Cost Ceramic Microfiltration Membranes for Separation of Oil and Bacteria From Its
Solution. J. Membr. Sci. 2011, 379 (1–2), 154–163.
80. Dong, Y.; Hampshire, S.; Zhou, J.; et al. Sintering and Characterization of Flyash-Based Mullite With MgO Addition. J. Eur. Ceram. Soc. 2011, 31 (5), 687–695.
81. Fang, J.; Qin, G.; Wei, W.; et al. Preparation and Characterization of Tubular Supported Ceramic Microfiltration Membranes From Fly Ash. Sep. Purif. Technol. 2011,
80 (3), 585–591.
82. Almandoz, M. C.; Marchese, J.; Prádanos, P.; et al. Preparation and Characterization of Non-Supported Microfiltration Membranes From Aluminosilicates. J. Membr. Sci.
2004, 241 (1), 95–103.
83. Dong, Y.; Liu, X.; Ma, Q.; et al. Preparation of Cordierite-Based Porous Ceramic Micro-Filtration Membranes Using Waste Fly Ash as the Main Raw Materials. J. Membr.
Sci. 2006, 285 (1–2), 173–181.
84. Majouli, A.; Younssi, S. A.; Tahiri, S.; et al. Characterization of Flat Membrane Support Elaborated From Local Moroccan Perlite. Desalination 2011, 277 (1–3), 61–66.
85. Kingman, S. W. Recent Developments in Microwave Processing of Minerals. Int. Mater. Rev. 2006, 51 (1), 1–12.
86. Menezes, R. R.; Kiminami, R. Microwave Sintering of Alumina-Zirconia Nanocomposites. J. Mater. Process. Technol. 2008, 203 (1–3), 513–517.
87. Bian, H. M.; Yang, Y.; Wang, Y.; et al. Alumina-Titania Ceramics Prepared by Microwave Sintering and Conventional Pressure-Less Sintering. J. Alloys Compd. 2012,
525, 63–67.
88. Oh, S. T.; Tajima, K.; Ando, M.; et al. Fabrication of Porous Al2O3 by Microwave Sintering and Its Properties. Mater. Lett. 2001, 48 (3-4), 215–218.
89. Feng, J.; Fan, Y.; Qi, H.; et al. Co-Sintering Synthesis of Tubular Bilayer Alpha-Alumina Membrane. J. Membr. Sci. 2007, 288 (1–2), 20–27.
90. Feng, J.; Qiu, M.; Fan, Y.; et al. The Effect of Membrane Thickness on the Co-Sintering Process of Bi-Layer ZrO2/Al2O3 Membrane. J. Membr. Sci. 2007, 305 (1–2),
20–26.
91. Dong, Y.; Lin, B.; Wang, S.; et al. Cost-Effective Tubular Cordierite Micro-Filtration Membranes Processed by Co-Sintering. J. Alloys Compd. 2009, 477 (1–2), L35–L40.
92. Chowdhury, S. R.; Keizer, K.; ten Elshof, J. E.; et al. Effect of Trace Amounts of Water on Organic Solvent Transport Through Gamma-Alumina Membranes With Varying
Pore Sizes. Langmuir 2004, 20 (11), 4548–4552.
93. Zhang, H.; Zhong, Z.; Xing, W. Application of Ceramic Membranes in the Treatment of Oilfield-Produced Water: Effects of Polyacrylamide and Inorganic Salts.
Desalination 2013, 309, 84–90.
94. Zhao, D.; Lau, E.; Huang, S.; et al. The Effect of Apple Cider Characteristics and Membrane Pore Size on Membrane Fouling. LWT Food Sci. Technol. 2015, 64 (2),
974–979.
95. Matsumoto, Y.; Nakao, S.; Kimura, S. Cross-Flow Filtration Solutions of Polymers Using Ceramic Microfiltration. Int. Chem. Eng. 1988, 28, 677–683.
96. Cheng, X.; Li, N.; Zhu, M.; et al. Positively Charged Microporous Ceramic Membrane for the Removal of Titan Yellow Through Electrostatic Adsorption. J. Environ. Sci.
2016, 44, 204–212.
97. Yang, C.; Zhang, G.; Xu, N.; et al. Preparation and Application in Oil–Water Separation of ZrO2/a-Al2O3 MF Membrane. J. Membr. Sci. 1998, 142 (2), 235–243.
98. Xu, N.; Li, W.; Zhao, Y.; et al. Theory and Method of Application-Oriented Ceramic Membranes Design (I). J. Chem. Ind. Eng. (China) 2003, 54 (9), 1284–1289.
99. Li, W.; Zhao, Y.; Liu, F.; et al. Theory and Method of Application-Oriented Ceramic Membranes Design (II). J. Chem. Ind. Eng. (China) 2003, 54 (9), 1290–1294.
100. Zhao, Y.; Li, W.; Zhang, W.; et al. Theory and Method of Application-Oriented Ceramic Membranes Design (III). J. Chem. Ind. Eng. (China) 2003, 54 (9), 1295–1299.
101. Elzo, D.; Huisman, I.; Middelink, E.; et al. Charge Effects on Inorganic Membrane Performance in a Cross-Flow Microfiltration Process. Colloids Surf. A Physicochem.
Eng. Asp. 1998, 138 (2–3), 145–159.
102. Zhang, L.; Li, N.; Zhu, M.; et al. Nano-Structured Surface Modification of Microporous Ceramic Membrane With Positively Charged Nano-Y2O3 Coating for Organic Dyes
Removal. RSC Adv. 2015, 5 (98), 80643–80649.
103. Fievet, P.; Szymczyk, A.; Aoubiza, B.; et al. Evaluation of Three Methods for the Characterisation of the Membrane–Solution Interface: Streaming Potential, Membrane
Potential and Electrolyte Conductivity Inside Pores. J. Membr. Sci. 2000, 168 (1–2), 87–100.
104. Lu, D.; Zhang, T.; Gutierrez, L.; et al. Influence of Surface Properties of Filtration-Layer Metal Oxide on Ceramic Membrane Fouling During Ultrafiltration of Oil/Water
Emulsion. Environ. Sci. Technol. 2016, 50 (9), 4668–4674.
296 Ceramic Membranes
105. Zhang, Q.; Jing, W.; Fan, Y.; et al. An Improved Parks Equation for Prediction of Surface Charge Properties of Composite Ceramic Membranes. J. Membr. Sci. 2008,
318 (1–2), 100–106.
106. Zhang, Q.; Fan, Y.; Xu, N. Effect of the Surface Properties on Filtration Performance of Al2O3–TiO2 Composite Membrane. Sep. Purif. Technol. 2009, 66 (2), 306–312.
107. Gao, N.; Li, M.; Jing, W.; et al. Improving the Filtration Performance of ZrO2 Membrane in Non-Polar Organic Solvents by Surface Hydrophobic Modification. J. Membr.
Sci. 2011, 375 (1–2), 276–283.
108. Gao, N. W.; Fan, Y. Q.; Quan, X. J.; et al. Modified Ceramic Membranes for Low Fouling Separation of Water-in-Oil Emulsions. J. Mater. Sci. 2016, 51 (13),
6379–6388.
109. Hsieh, H. P. Inorganic Membranes for Separation and Reaction; Elsevier Science: Amsterdam, 1996.
110. Almecija, M. C.; Ibanez, R.; Guadix, A.; et al. Effect of pH on the Fractionation of Whey Proteins With a Ceramic Ultrafiltration Membrane. J. Membr. Sci. 2007,
288 (1–2), 28–35.
111. De Angelis, L.; de Cortalezzi, M. M. F. Ceramic Membrane Filtration of Organic Compounds: Effect of Concentration, pH, and Mixtures Interactions on Fouling. Sep.
Purif. Technol. 2013, 118, 762–775.
112. De la Casa, E. J.; Guadix, A.; Ibanez, R.; et al. Influence of pH and Salt Concentration on the Cross-Flow Microfiltration of BSA Through a Ceramic Membrane. Biochem.
Eng. J. 2007, 33 (2), 110–115.
113. Fakhfakh, S.; Baklouti, S.; Baklouti, S.; et al. Preparation, Characterization and Application in BSA Solution of Silica Ceramic Membranes. Desalination 2010, 262 (1–3),
188–195.
114. Farsi, A.; Boffa, V.; Christensen, M. L. Electroviscous Effects in Ceramic Nanofiltration Membranes. ChemPhysChem 2015, 16 (16), 3397–3407.
115. Sbai, M.; Fievet, P.; Szymczyk, A.; et al. Streaming Potential, Electroviscous Effect, Pore Conductivity and Membrane Potential for the Determination of the Surface
Potential of a Ceramic Ultrafiltration Membrane. J. Membr. Sci. 2003, 215 (1–2), 1–9.
116. Alventosa-DeLara, E.; Barredo-Damas, S.; Zuriaga-Agusti, E.; et al. Ultrafiltration Ceramic Membrane Performance During the Treatment of Model Solutions Containing
Dye and Salt. Sep. Purif. Technol. 2014, 129, 96–105.
117. Yoon, S. H.; Kang, I. J.; Lee, C. H. Fouling of Inorganic Membrane and Flux Enhancement in Membrane-Coupled Anaerobic Bioreactor. Sep. Sci. Technol. 1999, 34 (5),
709–724.
118. Hoogland, M. R.; Fane, A. G.; Fell, C. J. D. The Effect of pH on the Cross-Flow Filtration of Mineral Slurries Using Ceramic Membranes. In Proceedings of the First
International Conference on Inorganic Membranes. Montpellier, France, 1989; p. 153.
119. Moosemiller, M. D.; Hill, C. G.; Anderson, M. A. Physicochemical Properties of Supported g-Al2O3 and TiO2 Ceramic Membranes. Sep. Sci. Technol. 1989, 24 (9–10),
641–657.
120. Nazzal, F. F.; Wiesner, M. R. pH and Ionic Strength Effects on the Performance of Ceramic Membranes in Water Filtration. J. Membr. Sci. 1994, 93 (1), 91–103.
121. Huisman, I. H.; Dutré, B.; Persson, K. M.; et al. Water Permeability in Ultrafiltration and Microfiltration: Viscous and Electroviscous Effects. Desalination 1997, 113 (1),
95–103.
122. Grabow, W. O. K.; Dohmann, M.; Haas, C.; et al. Influence of Particle Size and Surface Charge on Critical Flux of Crossflow Microfiltration. Water Quality International
‘98Water Sci. Technol. 1998, 38 (4), 481–488.
123. Chevereau, E.; Zouaoui, N.; Limousy, L.; et al. Surface Properties of Ceramic Ultrafiltration TiO2 Membranes: Effects of Surface Equilibriums on Salt Retention.
Desalination 2010, 255 (1–3), 1–8.
124. Chiu, T. Y. Effect of Ageing on the Microfiltration Performance of Ceramic Membranes. Sep. Purif. Technol. 2011, 83, 106–113.
125. Huisman, I. H.; Vellenga, E.; Trägårdh, G.; et al. The Influence of the Membrane Zeta Potential on the Critical Flux for Crossflow Microfiltration of Particle Suspensions. J.
Membr. Sci. 1999, 156 (1), 153–158.
126. Wang, P.; Xu, N.; Shi, J. A Pilot Study of the Treatment of Waste Rolling Emulsion Using Zirconia Microfiltration Membranes. J. Membr. Sci. 2000, 173 (2), 159–166.
127. Zhao, Y.; Zhong, J.; Li, H.; et al. Fouling and Regeneration of Ceramic Microfiltration Membranes in Processing Acid Wastewater Containing Fine TiO2 Particles. J.
Membr. Sci. 2002, 208 (1–2), 331–341.
128. Alpatova, A.; Kim, E.-S.; Dong, S.; et al. Treatment of Oil Sands Process-Affected Water With Ceramic Ultrafiltration Membrane: Effects of Operating Conditions on
Membrane Performance. Sep. Purif. Technol. 2014, 122, 170–182.
129. Cui, Z.; Peng, W.; Fan, Y.; et al. Effect of Cross-Flow Velocity on the Critical Flux of Ceramic Membrane Filtration as a Pre-treatment for Seawater Desalination. Chin. J.
Chem. Eng. 2013, 21 (4), 341–347.
130. Chang, Q.; Zhou, J.; Wang, Y.; et al. Application of Ceramic Microfiltration Membrane Modified by Nano-TiO2 Coating in Separation of a Stable Oil-in-Water Emulsion. J.
Membr. Sci. 2014, 456, 128–133.
131. Hutchings, G. J. New Approaches to Rate Enhancement in Heterogeneous Catalysis. Chem. Commun. 1999, 4, 301–306.
132. Julbe, A.; Farrusseng, D.; Guizard, C. Porous Ceramic Membranes for Catalytic Reactors—Overview and New Ideas. J. Membr. Sci. 2001, 181 (1), 3–20.
133. Jiang, H.; Meng, L.; Chen, R. Z.; et al. Progress on Porous Ceramic Membrane Reactors for Heterogeneous Catalysis Over Ultrafine and Nano-Sized Catalysts. Chin. J.
Chem. Eng. 2013, 21 (2), 205–215.
134. Zhong, Z.; Xing, W.; Jin, W.; et al. Adhesion of Nanosized Nickel Catalysts in the Nanocatalysis/UF System. AICHE J. 2007, 53 (5), 1204–1210.
135. Chen, R.; Du, Y.; Wang, Q.; et al. Effect of Catalyst Morphology on the Performance of Submerged Nanocatalysis/Membrane Filtration System. Ind. Eng. Chem. Res.
2009, 48 (14), 6600–6607.
136. Jiang, H.; Meng, L.; Chen, R. Z.; et al. A Novel Dual-Membrane Reactor for Continuous Heterogeneous Oxidation Catalysis. Ind. Eng. Chem. Res. 2011, 50 (18),
10458–10464.
137. Kim, M. J.; Kang, O. Y.; Rao, B. S.; et al. Proposing a New Fouling Index in a Membrane Bioreactor (MBR) Based on Mechanistic Fouling Model. Desalin. Water Treat.
2011, 33 (1–3), 209–217.
138. Wu, J.; He, C.; Jiang, X.; et al. Modeling of the Submerged Membrane Bioreactor Fouling by the Combined Pore Constriction, Pore Blockage and Cake Formation
Mechanisms. Desalination 2011, 279 (1–3), 127–134.
139. Chang, I. S.; Le Clech, P.; Jefferson, B.; et al. Membrane Fouling in Membrane Bioreactors for Wastewater Treatment. J. Environ. Eng.-ASCE 2002, 128 (11),
1018–1029.
140. Le-Clech, P.; Chen, V.; Fane, T. A. G. Fouling in Membrane Bioreactors Used in Wastewater Treatment. J. Membr. Sci. 2006, 284 (1–2), 17–53.
141. Zhong, Z.; Xing, W.; Liu, X.; et al. Fouling and Regeneration of Ceramic Membranes Used in Recovering Titanium Silicalite-1 Catalysts. J. Membr. Sci. 2007, 301 (1–2),
67–75.
142. Zhong, Z.; Li, D.; Liu, X.; et al. The Fouling Mechanism of Ceramic Membranes Used for Recovering TS-1 Catalysts. Chin. J. Chem. Eng. 2009, 17 (1), 53–57.
143. Vaidya, M. J.; Kulkarni, S. M.; Chaudhari, R. V. Synthesis of p-Aminophenol by Catalytic Hydrogenation of p-Nitrophenol. Org. Process Res. Dev. 2003, 7 (2), 202–208.
144. Koltuniewicz, A. B.; Field, R. W.; Arnot, T. C. Engineering of Membrane Processes II: Environmental Applications Cross-Flow and Dead-End Microfiltration of Oily-Water
Emulsion. Part I: Experimental Study and Analysis of Flux Decline. J. Membr. Sci. 1995, 102, 193–207.
145. Marchese, J.; Ochoa, N. A.; Pagliero, C.; et al. Pilot-Scale Ultrafiltration of an Emulsified Oil Wastewater. Environ. Sci. Technol. 2000, 34 (14), 2990–2996.
146. Bensadok, K.; Belkacem, M.; Nezzal, G. Treatment of Cutting Oil/Water Emulsion by Coupling Coagulation and Dissolved Air Flotation. Desalination 2007, 206 (1),
440–448.
147. Hua, F. L.; Tsang, Y. F.; Wang, Y. J.; et al. Performance Study of Ceramic Microfiltration Membrane for Oily Wastewater Treatment. Chem. Eng. J. 2007, 128 (2–3),
169–175.
Ceramic Membranes 297
148. Yeom, H. J.; Kim, S. C.; Kim, Y. W.; et al. Processing of Alumina-Coated Clay-Diatomite Composite Membranes for Oily Wastewater Treatment. Ceram. Int. 2016,
42 (4), 5024–5035.
149. Bhave, R. R.; Fleming, H. L. Removal of Oily Contaminants in Wastewater With Microporous Alumina Membranes. AlChE Symp. Ser. 1988, 84, 19–27.
150. Abbasi, M.; Taheri, A. Modeling of Coagulation-Microfiltration Hybrid Process for Treatment of Oily Wastewater Using Ceramic Membranes. J. Water Chem. Technol.
2014, 36 (2), 80–89.
151. Abbasi, M.; Taheri, A. Selecting Model for Treatment of Oily Wastewater by MF-PAC Hybrid Process Using Mullite-Alumina Ceramic Membranes. J. Water Chem.
Technol. 2016, 38 (3), 173–180.
152. Hilal, N.; Busca, G.; Hankins, N.; et al. Desalination Strategies in South Mediterranean Countries: The Use of Ultrafiltration and Nanofiltration Membranes in the Treatment
of Metal-Working Fluids. Desalination 2004, 167, 227–238.
153. Tomaszewska, M.; Orecki, A.; Karakulski, K. Desalination and the Environment Treatment of Bilge Water Using a Combination of Ultrafiltration and Reverse Osmosis.
Desalination 2005, 185 (1), 203–212.
154. Yu, X.; Zhong, Z.; Xing, W. Treatment of Vegetable Oily Wastewater Using an Integrated Microfiltration-Reverse Osmosis System. Water Sci. Technol. 2010, 61 (2),
455–462.
155. Yang, C.; Zhang, G.; Xu, N.; et al. Preparation and Application in Oil7Water Separation of ZrO2/a-Al2O3 MF Membrane. J. Membr. Sci. 1998, 142 (2), 235–243.
156. Zhang, G.; Gu, H.; Xing, W. Treatment of Cool Rolling Emulsion With Inorganic Ceramic Membranes. J. Chem. Eng. Chin. Univ. 1998, 12 (3), 288.
157. Kamble, S. P.; Barve, P. P.; Joshi, J. B.; et al. Purification of Lactic Acid via Esterification of Lactic Acid Using a Packed Column, Followed by Hydrolysis of Methyl
Lactate Using Three Continuously Stirred Tank Reactors (CSTRs) in Series: A Continuous Pilot Plant Study. Ind. Eng. Chem. Res. 2012, 51 (4), 1506–1514.
158. Datta, R.; Henry, M. Lactic Acid: Recent Advances in Products, Processes and Technologies—A Review. J. Chem. Technol. Biotechnol. 2006, 81 (7), 1119–1129.
159. Kyuchoukov, G.; Yankov, D. Lactic Acid Extraction by Means of Long Chain Tertiary Amines: A Comparative Theoretical and Experimental Study. Ind. Eng. Chem. Res.
2012, 51 (26), 9117–9122.
160. Bayazit, S. S.; Inci, I.; Uslu, H. Adsorption of Lactic Acid From Model Fermentation Broth Onto Activated Carbon and Amber Lite IRA-67. J. Chem. Eng. Data 2011,
56 (5), 1751–1754.
161. Bouchoux, A.; Roux-de Balmann, H.; Lutin, F. Investigation of Nanofiltration as a Purification Step for Lactic Acid Production Processes Based on Conventional and
Bipolar Electrodialysis Operations. Sep. Purif. Technol. 2006, 52 (2), 266–273.
162. Li, W.; Xing, W. Advances in Refinement of Lactic Acid From Fermentation Broths. Chem. Ind. Eng. Prog. 2009, 28 (3), 491–495.
163. Persson, A.; Jonsson, A. S.; Zacchi, G. Separation of Lactic Acid-Producing Bacteria From Fermentation Broth Using a Ceramic Microfiltration Membrane With Constant
Permeate Flow. Biotechnol. Bioeng. 2001, 72 (3), 269–277.
164. Pal, P.; Sikder, J.; Roy, S.; et al. Process Intensification in Lactic Acid Production: A Review of Membrane Based Processes. Chem. Eng. Process. 2009, 48 (11–12),
1549–1559.
165. Wang, K.; Li, W.; Fan, Y.; et al. Integrated Membrane Process for the Purification of Lactic Acid From a Fermentation Broth Neutralized With Sodium Hydroxide. Ind. Eng.
Chem. Res. 2013, 52 (6), 2412–2417.
166. Aspelund, M. T.; Glatz, C. E. Purification of Recombinant Plant-Made Proteins From Corn Extracts by Ultrafiltration. J. Membr. Sci. 2010, 353 (1–2), 103–110.
167. Roman, G. P.; Neagu, E.; Moroeanu, V.; et al. The Ultrafiltration Performance of Composite Membranes for the Concentration of Plant Extracts. Rom. Biotech. Lett. 2009,
14 (5), 4620–4624.
168. Xu, N.; Xing, W. Process for Refining Salt-Water by Inorganic-Membrane Filtering. Patent CN1868878, 2006.
169. Shirazi, S.; Lin, C.-J.; Chen, D. Inorganic Fouling of Pressure-Driven Membrane Processes—A Critical Review. Desalination 2010, 250 (1), 236–248.
170. Kim, J.; Yoon, T. I. Direct Observations of Membrane Scale in Membrane Bioreactor for Wastewater Treatment Application. Water Sci. Technol. 2010, 61 (9), 2267–2272.
171. Lee, M.; Kim, J. Membrane Autopsy to Investigate CaCO3 Scale Formation in Pilot-Scale, Submerged Membrane Bioreactor Treating Calcium-Rich Wastewater. J. Chem.
Technol. Biotechnol. 2009, 84 (9), 1397–1404.
172. Gu, J.; Zhang, H.; Zhong, Z.; et al. Conditions Optimization and Kinetics for the Cleaning of Ceramic Membranes Fouled by BaSO4 Crystals in Brine Purification Using a
DTPA Complex Solution. Ind. Eng. Chem. Res. 2011, 50 (19), 11245–11251.