Magalhães2018 Chapter CeramicMembranesTheoryAndEngin
Magalhães2018 Chapter CeramicMembranesTheoryAndEngin
Around the 1970s, in conjunction with the classical processes of separation, a new
class of process, using synthetic membranes with selective barrier, emerged. Such
membranes arise in an attempt to imitate existing natural membranes, with respect
to the characteristics of selectivity and permeability [1].
Membranes are semipermeable and selective barriers, whose function is to
separate phases, restricting, totally or partially, the transport of one or more species
present in a solution.
The membranes are mainly developed from two distinct classes of materials: the
organic, mostly polymers; and inorganic, such as metals and ceramics.
Separation processes using membranes are commonly driven by two forces:
concentration gradient between two phases of a semipermeable membrane and
pressure-driven in the membrane, which occur in microfiltration, ultrafiltration,
nanofiltration and reverse osmosis processes [2].
Filtration, especially those using membranes as a filter mechanism, has presented
good results due to the several characteristics that lead them to have the best cost/
benefit ratio, simplicity of operation, low energy cost, long lifetime and uniformity
of permeate quality throughout the process.
Park et al. [3] report that the efficiency of separation and economic viability in
membrane processes depends on the cost, operating energy, permeate flux and
lifetime of the membrane, which is directly related to the fouling, caused by the
polarization effect by concentration, formation of a gel layer on its surface,
adsorption and blocked pores.
In the literature, several studies have been reported on the development and
application of membranes, for example, the experimental works of Ma et al. [4],
Jiang et al. [5], Cheng et al. [6], Matos et al. [7] and Zhu et al. [8]. In addition, we
can cite the following theoretical works that use computational fluid dynamics as a
tool in the study of membrane separation processes, Darcovich et al. [9] and
Geraldes et al. [10] using flat plate membranes; Serra et al. [11] and Serra and
Wiesner [12] using circular membranes, and Bellhouse et al. [13], De Souza [14]
and Magalhães [15], using tubular membranes.
Particles
Microfiltration
10 μm
Ultrafiltration Macromolecules
0.1 μm
Sugars
Nanofiltration
0.002 μm
Water
ultrafiltration, the pore diameter range from 0.001 to 0.1 lm; for nanofiltration
membranes, the pore diameter range from 0.0005 to 0.002 lm; the choice depends
on the morphology and size of the particles in the solution [16]. Following this
topic will be detailed.
Ultrafiltration is a process indicated for solute particle size 10 times larger than the
particle size of the solvent molecule and is intended for the concentration or
fractionation of macromolecules [2]. In the water supply treatment, in which the
organic contaminants is the main concern, the ultrafiltration process is considered
adequate, presenting lower energy consumption, greater efficiency for the removal
of organic pollutants and natural organic matter as a function of the molecular
weight [19].
114 H. L. F. Magalhães et al.
membranes excel in the treatment of emulsions with particle diameters less than
20 lm, compared to other treatment techniques such as gravity sedimentation,
adsorption and flotation. They also have several benefits compared to polymer
membranes due to their mechanical performance and easy regeneration.
The pore shape, size and distribution are parameters that influence the perme-
ability, in which the fragility is the main disadvantage of the ceramic membrane,
which can be circumvented with a support material [33]. Further, Chen et al. [34]
studying tubular ceramic membranes for nanofiltration, report that although they are
particularly important for processes involving molecular separation under adverse
conditions, they are usually made by the sol-gel process, which often presents
problems of low efficiency and unsatisfactory control of the membrane properties.
In their work, Chen et al. [34], present a strategy for confection of ceramic mem-
brane, based on the atomic layer deposition and calcination, thus reducing this
problem.
In synthesis, the ceramic membranes have high applicability, being commonly
used in several separation processes, with good results.
Hwang and Kammermeyer [35] report that permeation can be defined as the phe-
nomenon that occurs when a particular species or component passes through
another substance, occasionally occurring, but not necessarily by the diffusion
mechanism. Since the term diffusion is specifically used to denote molecular dif-
fusion, the term permeation emerges as a phenomenological definition involving
various transport mechanisms.
Through the definition of the phenomenological permeability, it is possible to
express schematically the profile of concentration through a membrane. Figure 3
graphically shows this behavior. In this figure, we can see the existence of the initial
concentration (concentration of the contaminated fluid) and final concentration
(concentration do fluid after treatment with membrane), respectively, C1 and C2, as
well as the initial and final concentrations on the membrane surface, C1 and C2,
with formation of a boundary layer near the membrane.
When the permeation mechanism is the diffusion, the permeate flux, F, can be
written:
Ceramic Membranes: Theory and Engineering Applications 117
Boundary layer
through a membrane
C1
Boundary layer
Γ2
C2
C1 C2
F ¼ DAB S ð1Þ
e
dC
J C DAB ¼ J Cp ð3Þ
dx
x ¼ 0 ) C ¼ Cm ð4Þ
x ¼ d ) C ¼ Ca ð5Þ
Boundary
layer
Feed flow
Cm
J .C p
J .C
Ca
dC
DAB
dx Cp
x 0
1.2.3 Incrustation
Membrane operations have been widely used today; however, its large-scale and
continuous use in filtration systems has suffered some limitations. The main
problem is the membrane clogging, which provokes reduction of the filtrate flow
and process efficiency as well as the increase in cost due to the energy consumption
and materials for the cleaning process and backwashing.
There are three types of incrustation: deposition, precipitation and biofouling.
When suspended solids deposition occurs, such as colloids, organic, corrosion
products, iron hydroxide, algae and fine particulate matter, blockage of the feed
channel of the membrane modules occur, often being difficult to remove, resulting
in loss of performance.
Precipitation is a type of incrustation that occurs by precipitation of the soluble
compounds present in the feed when its solubility limit is reached. Due to the
concentration polarization, this effect intensifies near the surface of the membrane,
being able to reach the limit of the salts solubility present in the solution causing
precipitation; the most common are calcium carbonate, calcium sulfate, complexes
of silica, barium sulphate, strontium sulfate and calcium phosphate.
Biofouling occurs due to the accumulation of organic material on the membrane
surface, including cellular fragments, extracellular polymer substance and
microorganisms, which result in the formation of biofilms, being equally detri-
mental to the membrane lifetime [39].
During membrane filtration, the formation of the polarized layer and incrusta-
tions result in declining the membrane performance with respect to quality and
flow, reducing its reliability, since in applications with membranes, fouling control
is essential [40].
In pulp and paper industry, ceramic membranes are commonly used as the basis
of advanced treatment processes because of their hydrophilic characteristics, high
durability, high permeate fluxes, and thermal, chemical and mechanical stabilities.
According to Ebrahimi et al. [41], paper processing produces substantial volumes of
wastewater, requiring efficient recovery of waste, impurities and by-products prior
to disposal into environment. Thus, membrane filtration is an effective alternative,
which acts both in wastewater treatment and in the recovery of added value
effluents.
Unlu et al. [42], in his work about the use of membranes for the reuse of waste
water, report that in face of the various methods applicable to the dye wastewater
treatment/recovery in textile effluents, membrane filtration technology is considered
one of the most promising. Silva et al. [43] point out ceramic membranes as an
important tool for the treatment of water from this sector. Meksi et al. [44] notice
that textile wastewater has high staining, biochemical oxygen demand, chemical
oxygen demand and salinity, which makes its treatment essential before its reuse or
disposal in the environment, stressing out the membranes importance.
Lee et al. [45] report the use of ceramic membranes for water treatment. These
authors argue that ceramic membranes are still less used in the treatment of
wastewater to obtain potable water because of a historical conception about the high
manufacturing cost of these membranes in comparison to the polymer membranes
already consolidated in the market for this purpose. However, because the practice
of sustainable development linked to the manufacture of ceramic membranes, it was
possible to make membranes with better ratio cost-benefit, enabling the production
of higher performance and more compact systems, especially in forms tubular,
monolithic and in-plane-aligned, which currently present superior than performance
the hollow fiber membranes and present a large permeation fluxes.
Kumar et al. [46] conducted a research on the treatment of milk wastewater
using a low-cost tubular ceramic membrane. According to the authors, the dairy
industry produces an enormous quantity of wastewater due to the different opera-
tions present in the unit that demand this resource, generating approximately 2.5 L
of wastewater for each 1 L of processed milk. This residue contains a high con-
centration of nutrients, as well as high chemical and biological oxygen demand, as
well as high total suspended solids. Therefore, it is necessary to carry out an
appropriate treatment prior to disposal in the environment. In this context, mem-
branes, especially the ceramic, appear as promising alternative technology, when
compared to conventional treatment techniques, being a more economical process
and requiring less space for operation. Further, we can cite its resistance to cor-
rosive chemical products even in high temperatures.
In oil industry, membrane filtration technology is used in the processes of water
treatment produced from oil and gas exploration and production activities.
According to Jamaly et al. [47] oily wastewater is contaminated water at various
concentrations of oil, and may contain fats, hydrocarbons and petroleum fractions,
such as diesel oil, gasoline and kerosene, thus requiring adequate treatment for
subsequent disposal. According to the specifications regulated by the
Environmental Control Brazilian Agency, the content of oils and grease in effluents
Ceramic Membranes: Theory and Engineering Applications 121
cannot exceed 20 mg/L [48]. Despite of the various technologies applied to the
separation processes, the ceramic membranes have been distinguished by their high
permeate index, acting as one of the final stages of this treatment process.
Damak et al. [36] point out the separation processes using membranes as one of
the most significant developments in chemical and biological process engineering,
evidencing their application in a wide range of industrial operations, in addition to
those previously mentioned, such as: separation of organic molecules, pharma-
ceutical, biological macromolecules, colloids, ions and solvents.
Considered as the law that describes the fluid flow through a porous system,
Darcy’s law demonstrates that the fluid flow rate is directly proportional to the
pressure gradient, as follows:
VJ K DP
¼ ð7Þ
St l e
where VJ is the volume, S represent the area, t is the time, l viscosity, ΔP is the
pressure drop, e is the membrane thickness and K correspond to Darcy’s
permeability.
Hwang and Kammermeyer [35] report that Darcy’s law denotes that the flow
resistance is due to viscous drag, and that the permeability parameter, K, contains all
the properties of the porous system. Fluid viscosity is defined as the internal friction
between fluid sheets flowing at different velocities, where for a fluid in laminar flow,
such friction produces shear forces. When a fluid comes into contact with a solid
surface, it adheres to the surface, resulting in a zero velocity, and as a consequence of
the viscosity and fluid adhesion property, the solid surface experiences a drag force;
the viscous resistance is, thus, a contrary force to the drag force.
Damak et al. [36, 49], Cunha [50], Pak et al. [51], De Souza [14], and others, used
in their research, mathematical models whose the Navier–Stokes equations numer-
ical solutions are connected to Darcy’s law, written as a series resistance model.
122 H. L. F. Magalhães et al.
VJ e3p DP
¼ 2 2 ð8Þ
S t k 1 ep S l e
0
0
e3p
K¼ 2 ð9Þ
k0 1 ep S20
VJ nSpr4 DP
¼ ð10Þ
t 8l e
where n is the number of capillary tubes per area unit, and r is the capillary tube
radius.
Occurring when the fluid moves in a well-defined trajectory, the laminar flow
exhibits layers that individually preserve the characteristics of the fluid, in which
the viscosity is responsible for weakening the tendency for turbulences to appear.
Thus, Damak et al. [36] report a model in which the permeation rate is expressed
as a function of the pressure gradient, viscosity and flow resistance, as follows:
Ceramic Membranes: Theory and Engineering Applications 123
DP
Uw ¼ ð11Þ
l Rm þ Rp
Rp ¼ rp dp ð12Þ
In the Eq. (13), dp is the solute particles average diameter, and ep is the porosity
of the polarization layer per concentration. Equation (13) is valid for disperse
spherical particles (non-deformable) with porosity varying in the range
0.35 ep 0.75.
Damak et al. [36], when studying a separation process using laminar tubular
membrane, mentioned that the concentration polarization layer can be quantified by
the empirical correlation for boundary layer thickness of concentration, dp,
(Eq. 14), valid for Reynolds number varying between 300 < Re < 1000 and
Schmidt number ranging between 600 < Sc < 3200.
0:33
dp z
¼2 ðRe ScÞ0:33 Re0:3
w 1 0:4377 Sc0:0018 Re0:1551
w ð14Þ
di di
where Re is the Reynolds number, Rew represent the permeate Reynolds number, di
is the internal diameter, Sc correspond to Schmidt number, and z is the axial
coordinate.
The Reynolds permeation number is given by Eq. (15), as follows:
qUw Deq
Rew ¼ ð15Þ
l
where u is the membrane porosity, dp is the average diameter of the particles that
constitute the porous material, considering as spherical particles, and Uw is the
permeation velocity.
By using the correlation of the boundary layer thickness it is possible to con-
clude that increasing the number of Reynolds leads to a reduction of the boundary
layer thickness and an increase in the permeate flux, and that an increase in per-
meation Reynolds number and Schmidt number lead to increasing polarization by
concentration [50].
Early discussion is limited to laminar flow through porous media at low Reynolds
number (typically Re < 10) for which a proportionality validity relationships exists
between the fluid volumetric flow rate and pressure gradient.
Unlike the laminar flow regime, turbulent flow (flow at large Reynolds number)
occurs when the fluid does not move along a well-defined path, producing
momentum transfer. In this case, the deviation of the Darcy’s law is clear.
Due to the existence of turbulent flows, the mass conservation and linear
momentum equations only cannot adequately predict the oscillations resulting from
this phenomenon. Therefore, it is necessary to consider other terms to the equation
that predicts the fluid flow behavior. There are several models to predict turbulent
flow, e.g., RNG k-e model.
The RNG (Renormalization Group Theory) model emerged as an alternative of
the k-e standard model to flows with large Reynolds number, since various authors
defend the idea that the k-e turbulent model is inadequate to predict physical
situations where there is rotational flux. For this condition, the model overestimates
the kinetic energy dissipation, which results in a central recirculation region,
smaller than that experimentally observed. Thus, the turbulent kinetic energy and
kinetic energy dissipation rate equations (Eqs. 17 and 18) are defined by:
*
@ lT
ðqkÞ þ r qUk ¼ r lþ rk þ Pk qe ð17Þ
@t rkRNG
and
*
@ lT e
ðqeÞ þ r qUe ¼ r lþ re þ ðCe1RNG Pk Ce2RNG qeÞ ð18Þ
@t reRNG k
where k is the turbulent kinetic energy, e is the turbulent kinetic dissipation rate, l
is the dynamic viscosity, q is the density and lT is the turbulent viscosity, given by:
Ceramic Membranes: Theory and Engineering Applications 125
k2
lT ¼ Cl q ð19Þ
e
where
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Pk
g¼ ð21Þ
qeClRNG
In Eq. (21), the parameter Pk is the production term by the shear effect given by:
** *T
Pk ¼ lT rU rU þ rU þ Pkb ð22Þ
where PKb is the fluctuation production term (Eq. 23) determined as follows:
lT *
Pkb ¼ g rq ð23Þ
qrq
*
where g is the gravity acceleration vector, q the fluid specific mass, lT is the
turbulent viscosity and rl is the turbulent Prandtl number.
The values for the model constants, Ce2RNG, ClRNG, bRNG, are specified on the
Table 2.
and 3 m length, according to the dimensions used in the work of Damak et al. [53],
being, therefore, subjected to a tangential flow of oily water. As the membrane has
angular symmetry, only one cross section was used in the YZ plane.
For the numerical simulation, was generated a computational domain 2D, using
ICEM CFD software, as shown in Fig. 6. The mesh was created with a total of
having a total of 205,056 elements and 180,000 nodes, with concentration of the
elements near the surface of the membrane. Figure 7 shows the boundaries con-
ditions used in the modeling: inlet, concentrate outlet and permeate outlet.
Input
PlanYZ
Exit
(Concentrated)
Permeate outlet
Ceramic
membrane
Input Outlet of
concentrate
Axis of symmetry
From the considerations, the mass and linear momentum conservation equations
can be written as:
*
r qU ¼ 0 ð24Þ
*
where q and U are, respectively, fluid density and velocity vector.
• Momentum equation
*
* * * T
r qU U r lrU ¼ rp þ r l rU ð25Þ
@ *
ðqHÞ þ r qU H r ðCe rTÞ ¼ SH ð26Þ
@t
where Ce is the effective thermal diffusivity, H is enthalpy and SH is the heat source
term. The first term of Eq. (26) is responsible by the energy accumulation; the
second relates to heat transfer by convection and the third one is responsible for
heat transfer by diffusion.
• Mass transfer equation
*
U rC ¼ DAB r2 C ð27Þ
where C, is the solution (oil) concentration, DAB is the mass diffusivity, considered
constant for each Schmidt number (Sc), as follows:
l
DAB ¼ ð28Þ
Scq
where y is the radial coordinate, R is the tubular membrane internal radius and Uz is
the average velocity, determined by:
Re l
Uz ¼ ð30Þ
qR
C ¼ C0 ð31Þ
T ¼ T0 ð32Þ
P ¼ P0 ð33Þ
@C
¼0 ð34Þ
@z
@Uz
¼0 ð35Þ
@y
rT ¼ 0 ð36Þ
@Uz
¼0 ð37Þ
@y
@C
¼0 ð38Þ
@y
Uy ¼ 0 ð39Þ
@T
¼0 ð40Þ
@y
@Uz
¼0 ð41Þ
@x
@C
¼0 ð42Þ
@x
130 H. L. F. Magalhães et al.
@T
¼0 ð43Þ
@x
Uz ¼ 0 ð44Þ
rT ¼ 0 ð45Þ
Through the porous wall, radial velocity Uy is equal to the permeation velocity
Uw, as follows:
DP
Uy ¼ Uw ¼ ð46Þ
l Rm þ Rp
where Rm and Rp are, respectively, the membrane resistance and the resistance from
concentration polarization, and ΔP is the transmembrane pressure.
The equation responsible for the mass transport (Eq. 47) [54] was included in the
model as a source term. This, term is given by:
@C
Uw Rr C ¼ DAB ð47Þ
@y
where Rr is the solute intrinsic retention by the membrane that can be assumed to be
constant for a membrane-solute system.
The transmembrane pressure (Eq. 48) is defined as being the difference between
the average permeate pressure, Pp , and the external pressure to the membrane Pex,
as follows:
DP ¼ Pp Pex ð48Þ
The membrane resistance, Rm, the resistance from the polarization by concen-
tration, Rp, the concentration layer thickness, dp, and the resistance, rp, determined
when the concentration layer is considered homogenous, are given by Eqs. (2),
(12), (13) and (14).
Damak et al. [49] used Eq. 14 to determine local variation of the boundary layer
thickness by polarization, dp, so that the equilibrium between the convective and the
diffusive flows was reached when [(C − C0)/C0] < 0.001.
According to Damak et al. [49], the parameters (Re, Sc, Rew and z/d) from
Eq. (14) correspond to a membrane separation system for liquid ultrafiltration, low
particle concentration and laminar flow in the wall of the porous tube.
Ceramic Membranes: Theory and Engineering Applications 131
Important parameters used in the study of the 2D flow correspond to the thermal
and physico-chemical properties of the fluid and the membrane. The parameters are
given in Table 3.
In the simulations carried out, (Ansys CFX 15 software), the permeation
Reynolds number, the average diameter of the oil droplet, dp, Schmidt number,
initial concentration, permeability, temperature, and intrinsic solute retention by the
membrane, Rr, were kept constant according to Table 4, varying the axial Reynolds
number, Re, as shown in Table 5. All the cases were simulated in isothermal
conditions.
verified that the concentration increases with axial position and that the behavior of
the simulated results is similar to that one reported in the literature.
In Fig. 9, concentration profiles for different Reynolds numbers on the axial
position z = 0.50L are illustrated. It is expected that as the permeation velocity
increases a larger number of particles will be conducted convectively for near the
membrane surface. Further, according to the resistance model, the permeate flux
increases with transmembrane pressure. Therefore, higher Reynolds number pro-
vokes an increase in the transmembrane pressure, and consequently, concentration
in the membrane surface also will increase. From this figure, we can see that results
agreed well with those obtained in the literature.
Fig. 9 Reynolds number effect on the concentration profile inside the membrane (z = 0.50L)
(cases 1, 2 and 3)
Fig. 10 Thickness of the concentration boundary layer versus the non-dimensional variable z/d
for different Reynolds numbers
4 Final Considerations
Ceramic membranes are considered in the literature as one of the most promising
technologies in the present day. They are always present mainly in the stages of
separation and effluents treatment from the most varied industrial sectors. They can
be produced in order to assume various configurations and modules, adapting
134 H. L. F. Magalhães et al.
Acknowledgements The authors are grateful for financial support provided by CNPq, CAPES
and FINEP (Brazilian Research Agencies). We also acknowledge scientific support from the
authors mentioned along this chapter.
References
1. Habert, A.C., Borges, C.P., Nobrega, R.: Membrane separation processes. E-papers, Brazil,
Rio de Janeiro (2006) (in Portuguese)
2. Timoteo, J.F.J.: Anodizing for the production of ceramic membranes, p. 69. Masters
dissertation, Mechanical engineering, UFRN, Rio Grande do Norte, Brazil (2007) (in
Portuguese)
3. Park, J.Y., Jin Choi, S., Reum Park, B.: Effect of N2-back-flushing in multichannels ceramic
microfiltration system for paper wastewater treatment. Desalination 202(1–3), 207–214
(2007)
4. Ma, W., Guo, Z., Zhao, J., Yu, Q., Wang, F., Han, J., Pan, H., Yao, J., Zhang, Q., Samal, S.
K., De Smedt, S.C., Huang, C.: Polyimide/cellulose acetate core/shell electrospun fibrous
membranes for oil-water separation. Sep. Purif. Technol. 177, 71–85 (2017)
5. Jiang, Y., Hou, J., Xu, J., Shan, B.: Switchable oil/water separation with efficient and robust
Janus nanofiber membranes. Carbon 115, 477–485 (2017)
6. Cheng, Q., Ye, D., Chang, C., Zhang, L.: Facile fabrication of superhydrophilic membranes
consisted of fibrous tunicate cellulose nanocrystals for highly efficient oil/water separation.
J. Membr. Sci. 525, 1–8 (2017)
7. Matos, M., Gutiérrez, G., Lobo, A., Coca, J., Pazos, C., Benito, J.M.: Surfactant effect on the
ultrafiltration of oil-in-water emulsions using ceramic membranes. J. Membr. Sci. 520, 749–
759 (2016)
8. Zhu, X., Dudchenko, A., Gu, X., Jassby, D.: Surfactant-stabilized oil separation from water
using ultrafiltration and nanofiltration. J. Membr. Sci. 529, 159–169 (2017)
9. Darcovich, K., Dal-Cin, M.M., Ballevre, S., Wavelet, J.P.: CFD assisted thin channel
membrane characterization module. J. Membr. Sci. 124(2), 181–193 (1997)
Ceramic Membranes: Theory and Engineering Applications 135
10. Geraldes, V., Semião, V., Pinho, M.N.: Numerical modeling of mass transfer in slits with
semi-permeable membrane walls. Eng. Comput. 17(3), 192–217 (2000)
11. Serra, C.A., Wiesner, M.R., Laîné, J.M.: Rotating membrane disk filters: design evaluation
using computational fluid dynamics. Chem. Eng. J. 72(1), 1–17 (1999)
12. Serra, C.A., Wiesner, M.R.: A comparison of rotating and stationary membrane disk filters
using computational fluid dynamics. J. Membr. Sci. 165(1), 19–29 (2000)
13. Bellhouse, B.J., Costigan, G., Abhinava, K., Merry, A.: The performance of helical
screw-thread inserts in tubular membranes. Sep. Purif. Technol. 22–23, 89–113 (2001)
14. De Souza, J.S.: Theoretical study of the microfiltration process in ceramic membranes, p. 134.
Doctoral Thesis, Process Engineering, UFCG, Paraíba, Brazil (2014) (in Portuguese)
15. Magalhães, H.L.F.: Study of the thermofluidodynamics of the treatment of effluents using
ceramic membranes: modeling and simulation, p. 102. Masters dissertation in Mechanical
Engineering, Federal University of Campina Grande, Campina Grande, Brazil (2017) (in
Portuguese)
16. Taketa, T.B., Ferreira, M.Z., Gomes, M.C.S., Curvelo, N.: Production of biodiesel by ethyl
transesterification of vegetable oils and their separation and purification by ceramic
membranes. In: VIII Brazilian Congress of Chemical Engineering in Scientific Initiation,
Uberlândia, Minas Gerais, Brasil (2009) (in Portuguese)
17. De Carvalho, R.B., Cristiano, P.B., Nobrega, R.: Formation of double-spreading cellulosic flat
membranes for nanofiltration and reverse osmosis processes. Polymers 11(2), 65–75 (2001)
(in Portuguese)
18. Lopes, A.C.: Degradation study of polymeric membranes of commercial nanofiltration by
sodium hypochlorite, p. 92. Master dissertation, Federal University of Rio de Janeiro, Rio de
Janeiro, Brazil (2006) (in Portuguese)
19. Mierzwa, J.C., Da Silva, M.C.C., Rodrigues, L.D.B., Hespanhol, I.: Water treatment for
public supply by ultrafiltration: comparative evaluation through the direct costs of
implantation and operation with the conventional and conventional systems with activated
carbon. Sanitary Environ. Eng. 13(1), 78–87 (2008) (in Portuguese)
20. Rosa, D.S., Salvini, V.R., Pandolfelli, V.C.: Processing and evaluation of the properties of
porous ceramic tubes for microfiltration of emulsions. Ceramics 52(322), 167–171 (2006) (in
Portuguese)
21. Makabe, R., Akamatsu, K., Nakao, S.: Classification and diafiltration of polydispersed
particles using cross-flow microfiltration under high flow rate. J. Membr. Sci. 523, 8–14
(2017)
22. Yang, X., Zhou, S., Li, M., Wang, R., Zhao, Y.: Purification of cellulase fermentation broth
via low cost ceramic microfiltration membranes with nanofibers-like attapulgite separation
layers. Sep. Purif. Technol. 175, 435–442 (2017)
23. Wang, X., Wang, C., Tang, C.Y., Hu, T., Li, X., Ren, Y.: Development of a novel anaerobic
membrane bioreactor simultaneously integrating microfiltration and forward osmosis
membranes for low strength wastewater treatment. J. Membr. Sci. 527, 1–7 (2017)
24. Suresh, K., Pugazhenthi, G., Uppaluri, R.: Fly ash based ceramic microfiltration membranes
for oil-water emulsion treatment: parametric optimization using response surface methodol-
ogy. J. Water Process Eng. 13, 27–43 (2016)
25. Suresh, K., Pugazhenthi, G.: Cross flow microfiltration of oil-water emulsions using clay
based ceramic membrane support and TiO2 composite membrane. Egypt. J. Petrol. 14(1), 1–
10 (2016)
26. Becker, C.M.: Obtaining and characterizing sulphonated polyelectrolytes based on styrenic
copolymers for polymer membranes, p. 96. Masters dissertation, Federal University of Rio
Grande do Sul, Porto Alegre, Brazil (2007) (in Portuguese)
27. Leite, A.M.D., Ito, E.N., Araújo, E.M., Lira, H. De L., Barbosa, R.: Obtaining microporous
membranes from nanocomposites of polyamide 6/national clay. Part 1: influence of clay
presence on membrane morphology. Polymers 19(4), 271–277 (2009) (in Portuguese)
136 H. L. F. Magalhães et al.
28. Perles, C.E.: Physical and chemical properties related to the development of Nafion®
membranes for PEMFC fuel cell applications. Polymers 18(4), 281–288 (2008) (in
Portuguese)
29. Silva, A.A., Melo, K.S., Maia, J.B.N.: Study of the water/oil separation potential of alumina
tubular ceramic membranes through analysis of flow and turbidity measurements. In: 2º
Brazilian Congress of P & D in Oil & Gas. Rio de Janeiro: UFRJ, Rio de Janeiro, Brazil
(2003) (in Portuguese)
30. Cui, J., Zhang, X., Liu, H., Liu, S., Yeung, K.L.: Preparation and application of zeolite/
ceramic microfiltration membranes for treatment of oil contaminated water. J. Membr. Sci.
325(1), 420–426 (2008)
31. Silva, F.A., Lira, H.L.: Preparation and characterization of cordierite ceramic membranes.
Ceramics 52(324), 276–282 (2006) (in Portuguese)
32. Zhu, L., Chen, M., Dong, Y., Tang, C.Y., Huang, A., Li, L.: A low-cost mullite-titania
composite ceramic hollow fiber microfiltration membrane for highly efficient separation of
oil-in water emulsion. Water Res. 90, 277–285 (2016)
33. Maia, D.F.: Development of ceramic membranes for oil/water separation, p. 111. Doctorate
Thesis in Process Engineering, Federal University of Campina Grande, Campina Grande,
Brazil (2006) (in Portuguese)
34. Chen, H., Jia, X., Wei, M., Wang, Y.: Ceramic tubular nanofiltration membranes with tunable
performances by atomic layer deposition and calcination. J. Membr. Sci. 528, 95–102 (2017)
35. Hwang, S.T., Kammermeyer, K.: Membranes in Separation. Wiley, Canada (1975)
36. Damak, K., Ayadi, A., Schmitz, P., Zeghmati, B.: Modeling of cross-flow membrane
separation processes under laminar flow conditions in tubular membrane. Desalination 168,
231–239 (2004)
37. Mulder, M.: Basic Principles of Membrane Technology, 1st edn. Kluwer Academic
Publishers, Netherlands (1996)
38. Adams, M.C., Barbano, D.M.: Effect of ceramic membrane channel diameter on limiting
retentate protein concentration during skim milk microfiltration. J. Dairy Sci. 99(1), 167–182
(2016)
39. Oliveira, D.R.: Pre-treatment of the reverse osmosis process using microfiltration and
investigation of membrane cleaning and recovery techniques, p. 129. Master dissertation,
Federal University of Rio de Janeiro, Rio de Janeiro, Brazil (2007) (in Portuguese)
40. Ahmed, S., Seraji, M.T., Jahedi, J., Hashib, M.A.: Application of CFD for simulation of a
baffled tubular membrane. Chem. Eng. Res. Des. 90(5), 600–608 (2012)
41. Ebrahimi, M., Busse, N., Kerker, S., Schmitz, O., Hilpert, M., Czermak, P.: Treatment of the
bleaching effluent from sulfite pulp production by ceramic membrane filtration. MDPI
Membr. 7, 1–15 (2015)
42. Unlu, M., Yukseler, H., Yetis, U.: Indigo dyeing wastewater reclamation by membrane-based
filtration and coagulation processes. Desalination 240, 178–185 (2009)
43. Silva, M.C., Oliveira, R.C., Lira, H.L., Freitas, N.L.: Obtaining a ceramic membrane to treat
effluent from the textile industry. Electron. J. Mater. Process. 9, 81–85 (2014) (in Portuguese)
44. Meksi, N., Ben Ticha, M., Kechida, M., Mhenni, M.F.: Using of ecofriendly
a-hydroxycarbonyls as reducing agents to replace sodium dithionite in indigo dyeing
processes. J. Clean. Prod. 24, 149–158 (2012)
45. Lee, M., Wu, Z., Li, K.: Advances in membrane technologies for water treatment: materials,
processes and applications. In: Advances in Ceramic Membranes for Water Treatment, vol. 1,
pp. 43–82, 1st edn. Woodhead Publishing, England (2015)
46. Kumar, R. V., Goswami, L., Pakshirajanb, K., Pugazhenthi, G.: Dairy wastewater treatment
using a novel low cost tubular ceramic membrane and membrane fouling mechanism using
pore blocking models. J. Water Process Eng. 13, 168–175 (2016)
47. Jamaly, S., Giwa, A., Hasan, S.W.: Recent improvements in oily wastewater treatment:
progress, challenges, and future opportunities. J. Environ. Sci. 37, 15–30 (2015)
48. CONAMA No20/ART.21, RE. Standard CONAMA n°20, de June 18, Brazil (1986) (in
Portuguese)
Ceramic Membranes: Theory and Engineering Applications 137
49. Damak, K., Ayadi, A., Zeghmati, B., Schmitz, P.: Concentration polarisation in tubular
membranes—a numerical approach. Desalination 171(2), 139–153 (2004)
50. Cunha, A.L.: Treatment of effluents from the petroleum industry via ceramic membranes—
modeling and simulation, p. 201. Doctorate Thesis in Process Engineering, Federal University
of Campina Grande, Campina Grande, Brazil (2014) (in Portuguese)
51. Pak, A., Mohammad, T., Hosseinalipour, S.M., Allahdinib, V.: CFD modeling of porous
membranes. Desalination 222(1–3), 482–488 (2008)
52. Ansys: CFX 15, Solver Theory Guide. Ansys, Japan (2015)
53. Damak, K., Ayadi, A., Zeghmati, B., Schmitz, P.: New Navier-Stokes and Darcy’s law
combined model for fluid flow in cross-flow filtration tubular membranes. Desalination 161
(1), 67–77 (2004)
54. Minnikanti, V.S., Dasgupta, S., De, S.: Prediction of mass transfer coefficient with suction for
turbulent flow in cross flow ultrafiltration. J. Membr. Sci. 154(2), 227–239 (1999)