Вы находитесь на странице: 1из 17

Journal of Membrane Science 207 (2002) 73–89

Alumina and titania multilayer membranes for nanofiltration:


preparation, characterization and chemical stability
Tim Van Gestel a,b,∗ , Carlo Vandecasteele a , Anita Buekenhoudt b , Chris Dotremont b ,
Jan Luyten b , Roger Leysen b , Bart Van der Bruggen a , Guido Maes c
a K.U. Leuven, Department of Chemical Engineering, W. de Croylaan 46, B-3001 Heverlee, Belgium
b Vito (Flemish Institute for Technological Research), Process Technology, Boeretang 200, B-2400 Mol, Belgium
c K.U. Leuven, Department of Chemistry, Celestijnenlaan 200, B-3001 Heverlee, Belgium

Received 2 July 2001; received in revised form 11 February 2002; accepted 12 February 2002

Abstract
The preparation and characterization of porous ceramic multilayer nanofiltration (NF) membranes is described. During
preparation, special care was given to each sub-layer that forms a part of the multilayer configuration: the macroporous
substrate, the membrane interlayers and the NF toplayers. High-quality macroporous supports are prepared from ␣-Al2 O3 .
Three types of colloidal sol–gel derived mesoporous interlayers are considered: Al2 O3 , TiO2 and mixed Al2 O3 –TiO2 . The
active NF toplayer is a very thin and fine textured polymeric TiO2 toplayer.
Optimized ␣-Al2 O3 /␥-Al2 O3 /anatase and ␣-Al2 O3 /anatase/anatase multilayer configurations show high retentions for
relatively small organic molecules (molecular weight cut-off <200). Different membrane layers have very different crystal-
lographic properties resulting in a considerably different chemical stability. Corrosion measurements showed that application
of a multilayer configuration including weakly crystallized ␥-Al2 O3 layers is restricted to mild aqueous media (pH 3–11)
or non-aqueous media (organic solvents). For NF applications in aqueous media with a lower or higher pH, the multilayer
membrane composed of anatase on a ␣-Al2 O3 support is to be preferred. © 2002 Elsevier Science B.V. All rights reserved.
Keywords: Ceramic membranes; Sol–gel; Characterization; Chemical stability; Nanofiltration

1. Introduction merous applications in the area of process, waste and


drinking water treatment. However, the major draw-
Nanofiltration (NF) membranes can be gener- back is that the chemical stability of the conventional
ally classified into two major groups according to polymeric membranes is limited with respect to cor-
their material properties: organic polymeric mem- rosive media like strong acids and organic solvents
branes and inorganic ceramic membranes. Polymeric [1,2]. Therefore, much research to date focuses on
NF-membranes are today commercially available in a ceramic NF-membranes. Since it is generally recog-
large number of different polymeric materials (poly- nized that inorganic materials are inherently more
sulfone, cellulose acetate, polyamide, . . . ). Membrane stable than polymers, they are expected to constitute
filtration based on such membranes has found nu- a promising alternative for polymeric membranes [3].
Ceramic NF-membranes with high performance
∗ Corresponding author. Tel.: +32-1433-5615; parameters such as low cut-off values or high fluxes
fax: +32-1432-1186. can only be obtained in an asymmetric multilayer
E-mail address: vgestelt@vito.be (T. Van Gestel). configuration. The development of such a multilayer

0376-7388/02/$ – see front matter © 2002 Elsevier Science B.V. All rights reserved.
PII: S 0 3 7 6 - 7 3 8 8 ( 0 2 ) 0 0 0 5 3 - 4
74 T. Van Gestel et al. / Journal of Membrane Science 207 (2002) 73–89

configuration includes: shaping of an appropriate sup- proved their feasibility for NF applications (MWCO
port material; formation of mesoporous interlayers; <500).
and synthesis of a microporous toplayer with a cut-off All the authors mentioned mainly considered on
value below 1000. Alumina, titania, zirconia or silica the structural properties of the developed membrane
are considered as the main ceramic materials for the materials. Therefore, only limited data on the chemi-
formation of the multilayer structures [4–6]. cal stability of meso- or microporous membranes are
The preparation route starts with the production available. Züter et al. [18] reported acid corrosion
of a high-quality support system, as the effectiveness tests on mesoporous ␥-Al2 O3 , anatase/TiO2 and ZrO2
of the developed NF membrane is greatly influenced membranes. In our group, Schaep et al. [19] tested the
by the structural properties of the membrane support. corrosion properties of mesoporous ␥-Al2 O3 mem-
Based on earlier studies in our group, alumina sup- branes. These contributions have however mainly
ports produced by conventional slip-casting and reac- a descriptive character and the membrane stability
tion bonded Al2 O3 (RBAO) manufacturing routes are was not investigated systematically by taking into
used as substrate material for the development of a account important parameters like the structural char-
multilayer membrane [7–9]. acteristics of the ceramic material (phase behaviour,
The second part deals with the synthesis of several crystallinity, surface area). To our knowledge, only
intermediate membrane layers. The aim of this modifi- Hollstein et al. [20] addressed in detail the corrosion
cation is twofold: fine-grained mesoporous membrane properties of ceramic materials in aqueous media, but
layers are necessary in order to reduce the coarse pore did not describe the stability of porous materials like
structure of the support material and large surface sol–gel membrane materials.
irregularities or defects can only be covered properly The aim of this paper is to provide results on the
by a multiple dip-coating procedure. In this study, chemical stability of different sol–gel derived mem-
colloidal sol–gel derived membrane layers consisting brane materials. The corrosion behaviour of meso-
of alumina or titania are considered. Firstly, ␥-Al2 O3 porous interlayers and microporous toplayers, along
and anatase/TiO2 membrane interlayers were pre- with the membrane characterization in terms of pore
pared based on earlier studies of Burggraaf et al. [10], size, surface area, pore volume and crystallite size is
Zaspalis et al. [11], Kumar et al. [12] and Lin et al. described. For the determination of the corrosion rate,
[13]. Additionally, different combined Al2 O3 –TiO2 preference was given to simple static corrosion exper-
membrane interlayers were developed in order to iments whereby the amount of dissolved membrane
combine the specific properties of both materials, material was determined in acid or alkaline solutions
namely the possibility of obtaining membrane layers with a pH ranging from 1 to 13. With the experimen-
with a high thickness and without defects on one hand tal information, materials could be selected specifi-
and a good chemical stability, on the other hand. cally for each membrane layer in order to develop an
The final step in the multilayer preparation route optimal multilayer configuration for particular pro-
involves the synthesis of a thin separation toplayer. cess conditions (non-aqueous solvent, acid, alkali).
Formation of such a layer with very small pores re- The performance characteristics of the multilayer
quires an appropriate sol consisting of very small membranes are also described to indicate possible
nanometer-sized particles, which is typically obtained application fields of the developed NF-membranes.
by the so-called ‘polymeric’ sol–gel technique. In
this work, TiO2 NF toplayers are formed by a simi-
lar sol–gel procedure as previously described by de 2. Experimental
Lange et al. [14] and Keizer et al. [15]. They syn-
thesized microporous amorphous SiO2 toplayers for 2.1. Membrane preparation
the modification of mesoporous ␥-Al2 O3 membrane
layers in order to obtain a multilayer gas separation 2.1.1. Preparation of support materials
membrane. Recently, Voigt et al. [16,17] prepared For the preparation of the membrane support, two
also microporous toplayers, consisting of amor- routes were followed. In the first route, Al2 O3 supports
phous TiO2 , on anatase/TiO2 membrane layers and were prepared by a slip-casting technique, starting
T. Van Gestel et al. / Journal of Membrane Science 207 (2002) 73–89 75

from an aqueous suspension of commercially avail- produced by pouring the remaining sol–gel solutions
able Al2 O3 powder (AKP-30, Sumitomo) with organic in petri-dishes. In both cases, the obtained gel layers
additives (Vanisperse CNH, Borregaard) [7]. In the were dried at room temperature for 24 h. Finally, the
second route, Al2 O3 supports were prepared by the membranes were calcinated for 1 h at 400 ◦ C in a
reaction bonded Al2 O3 (RBAO) manufacturing pro- temperature programmable furnace (Scandia SK-15).
cess [8]. In this process Al2 O3 powder (HC Starck) The heating and cooling rates for Al2 O3 , TiO2 and
and Al particles (Baudier) are milled in a weight ra- mixed membranes were 60, 10 and 15 ◦ C/h, respec-
tio of 40/60. After shaping of the support material by tively. Then, depending on the type of membrane
isostatic pressing, oxidation is performed at 1100 ◦ C, material, the first calcination step was followed by a
followed by a final thermal treatment at 1300 ◦ C. thermal treatment for 1 h at a temperature between
According to both preparation methods, either 400 and 1200 ◦ C.
disk-shaped or tubular membrane supports can be
produced. The final support disks were 25 mm in di- 2.1.3. Preparation of the membrane toplayers
ameter and 2 mm in thickness. The tubular supports by the polymeric sol–gel process
were 120 mm in length with an i.d. of 7 mm and an A polymeric TiO2 sol was produced starting from
o.d. of 10 mm. the common TTI precursor, which was partially hy-
drolyzed with a less than equivalent amount of water
2.1.2. Preparation of the membrane interlayers by in order to obtain a precipitate free TiO2 sol. In con-
the colloidal sol–gel process trast with the colloidal route, for the preparation of the
Precursors for the colloidal sol–gel process were very fine polymeric membrane layer no high molecular
common aluminium and titanium based organometal- weight organic additives could be added. Therefore,
lic compounds, which were hydrolyzed by the addition the freshly prepared titania sol was aged for 3 days in
of an excess H2 O. The ␥-AlOOH sol was prepared order to increase the viscosity of the dipping solution
by hydrolysis of Al-tri-sec-butoxide (Al(OC4 H9 )3 ) and to prevent penetration of the sol into the pores of
(ASB, Merck). The sol was peptized at 90 ◦ C by addi- the intermediate support layer during dip-coating. At
tion of 0.07 mol HNO3 per mole ASB. The TiO2 sol the same time, unsupported membranes were prepared
was obtained by hydrolysis of Ti-tetra-isopropoxide by evaporating the sol in a petri-dish. After drying at
(Ti(OC3 H7 )4 ) (TTI, Janssen Chimica). Peptization room temperature, thermal treatment was performed
was obtained at 50 ◦ C with 0.5 mol of HNO3 per mole for 1 h with heating and cooling rates of 15 ◦ C/h in the
of TTI. Binary alumina/titania sols were prepared temperature range 200–700 ◦ C.
using two different sol–gel routes. The first route in-
volved mixing of two separately peptized ␥-AlOOH 2.2. Characterization
and TiO2 sols in a certain ratio. In the second method,
homogeneous mixed sols were produced in one step 2.2.1. Static characterization
by co-hydrolysis of ASB–TTI precursor mixtures Several methods were used for the characterization
followed by peptization with HNO3 . of the membrane support. The structure was ana-
For alumina and mixed membranes, a solution lyzed by X-ray diffraction (XRD, Siemens D500) and
of polyvinyl alcohol (PVA, Erkol) with an aver- Hg-porosimetry (Autoscan-33, Porosimeter, Quan-
age molecular weight of 72,000 (3.5 g/100 ml H2 O) tachrome). The pore size and the presence of possible
was added to the sols before membrane formation defects in the supports were determined with field
as drying chemical controlling additive (DCCA). emission scanning electron microscopy (JEOL JSM
For the preparation of titania membranes, a solu- 6340 F). The surface roughness was measured with
tion of hydroxypropyl cellulose (HPC, Aldrich) with a laser profilometer (Micro Focus Measurement Sys-
MW = 100,000 (1 g/100 ml H2 O) and PVA with tem, UBM). The mechanical properties were charac-
MW = 72,000 was used (0.1 g/100 ml H2 O). terized by four-point bend tests (Chevron-notch).
Supported gel layers were formed on the mem- The particle sizes in ␥-AlOOH, TiO2 and mixed sols
brane support by dip-coating the supports with were determined by the dynamic laser beam scattering
freshly prepared sols. Non-supported gel layers were technique (PCS, N4 plus Coulter).
76 T. Van Gestel et al. / Journal of Membrane Science 207 (2002) 73–89

The morphology, surface quality and thickness 2.2.3. Dynamic characterization


of intermediate and toplayer membranes was ex- Filtration experiments were carried out with a
amined with optical microscopy (Zeiss) and field commercial laboratory scale installation (Amafilter
emission scanning electron microscopy (JEOL JSM PSS-2TZ). The molecular weight cut-off (MWCO)
6340 F). The structural properties of the membrane of the membranes was determined by using a solu-
materials were characterized by X-ray diffrac- tion containing polyethylene glycols (PEG, Merck)
tion, N2 -adsorption/desorption (Autosorb 1, Quan- of different molecular weight: PEG 200, PEG 600,
tachrome) and thermogravimetric analysis (TGA, 1500, 5600, 10000 and 15000. To determine the PEG
Setaram). XRD measurements were used for the de- retention as a function of the molecular weight, feed
termination of the phase composition and the average and permeate samples were taken and analyzed by
crystallite size in the membrane material. The crys- gel permeation chromatography (GPC, Waters, ultra-
tallite size was determined with the Debye–Scherrer hydrogel). During the filtration experiments, the flux
equation by using the full-width at half-maximum was measured in order to determine the membrane
value (FWHM) of the main diffraction peaks. permeability. All the experiments were carried out at
N2 -adsorption/desorption measurements were per- a transmembrane pressure of 5 bar.
formed to determine the pore size distribution, BET
surface area and pore volume. All structural measure-
ments were carried out on unsupported membrane in- 3. Results and discussion
terlayers and toplayers, assuming that their properties
were similar to those of supported membrane layers. 3.1. Support materials

2.2.2. Static corrosion tests For the development of high-quality supports the
Static corrosion tests were performed in aque- following properties are of major importance: pore
ous solutions at a pH ranging from 1 to 3. These size distribution; porosity; surface quality with the ab-
solutions were prepared using demineralized water sence of large defects or large pores; mechanical and
(MilliQ, Millipore) and nitric acid (Merck, pro anal- chemical stability.
ysis ([Al] = 0.05 ppm, [Ti] = 0.02 ppm)). Then, Structural characteristics of the final Al2 O3 mem-
100 mg calcinated membrane material was immersed brane supports prepared according to the two produc-
in 100 ml of the acid solution. The solution was shaken tion routes are summarized in Table 1. Basically, the
continuously for 4 days on a shaking table (Gerhardt support material consisted of ␣-Al2 O3 . The pore size
laboshake). After the exposure time, liquid samples depended on the preparation method: the fine-grained
were taken, filtered over a Whatman 42 filter and ana- AKP-30 support had an average pore size around
lyzed for their Al and Ti concentration by inductively 0.1 ␮m, while the RBAO support contained larger
coupled plasma-mass spectroscopy (VG-PlasmaQuad grains and showed also a larger pore size. The poros-
PQ-2 Plus). Additionally, corrosion experiments were ity was typically ca. 40% and surface roughness was
also conducted in NaOH solutions (Merck, pro anal- ca. 0.3 ␮m. Based on four-point bend tests, it is shown
ysis (Al content = 0.0005%)) at a pH ranging from that the RBAO derived support has the advantage of
11 to 13. All corrosion experiments were performed being mechanically much stronger than the conven-
at ambient temperature. The pH measurements were tional AKP-30 support. According to Luyten et al.,
carried out with a digital pH-meter (ATI Orion 310). the higher mechanical stability of the RBAO material

Table 1
Characteristics of the support materials
Type Thermal Phase Pore size Porosity (%) Roughness (␮m) Bend-strength Corrosion pH 1
treatment (◦ C) (␮m) (mPa) (␮g Al/l)
AKP-30 1130 ␣-Al2 O3 0.1 40 0.3 40 571
RBAO 1300 ␣-Al2 O3 0.2 40 0.3 100 –
T. Van Gestel et al. / Journal of Membrane Science 207 (2002) 73–89 77

is due to the presence of fine ␣-Al2 O3 grains—formed presence of relatively large particles these sols were
during the oxidation of the Al powder fraction—in semi-transparent and looked light blue. Binary sols,
the large grained primary ␣-Al2 O3 structure [8]. Long obtained by mixing separately peptized ␥-AlOOH
term corrosion measurements on ␣-Al2 O3 support ma- and TiO2 sols, showed both typical characteristics of
terial in nitric acid solutions showed also a high chem- the individual sols, namely a light blue colour and
ical stability. For a corrosion test at pH 1 only small an average particle size distribution around 40 nm.
amounts of aluminium were found even after an ex- Stable mixed sols, prepared by co-hydrolysis of the
posure time of 6 months. precursors, were completely transparent (colourless)
after peptization. In this case smaller sol particle sizes
3.2. Sols were obtained with a mean diameter of ca. 25 nm.
For the synthesis of the TiO2 toplayer, a precipitate-
Both the colloidal and polymeric sols were prepared free sol containing small nanometer-sized polymeric
by hydrolysis of alkoxide precursors to induce con- structures (fractals) was prepared. The essential
densation and inorganic polymerization reactions. In features of the polymeric sol preparation were a care-
the different routes, the essential parameter to control fully controlled hydrolysis with a less than stoechio-
was the degree of hydrolysis. metric amount of water ([H2 O]/[Ti] < 4) followed by
an appropriate aging. The final aged polymeric TiO2
Al(OC4 H9 )3 + xH2 O → Al(OC4 H9 )3−x (OH)x sol was completely transparent. Average fractal sizes
+xC4 H9 OH were smaller than 5 nm, as no particle sizes above
5 nm were observed by laser scattering experiments.
Ti(OC3 H7 )4 + xH2 O → Ti(OC3 H7 )4−x (OH)x
3.3. Membrane interlayers
+xC3 H7 OH
For the formation of mesoporous membrane inter- 3.3.1. Morphology
layers, suspensions of relatively large nanometer-sized For the development of high-quality multilayer
particles were aimed at. This was obtained by react- NF-membranes, we aimed at the formation of rather
ing the alkoxide precursors with a large excess water thick intermediate layers, which were required to
([H2 O]/[ASB] > 3, [H2 O]/[TTI] > 4) followed by a overcome inevitable surface irregularities of the sup-
refluxing step overnight under the conditions shown port.
in Table 2. Alumina layers could easily be produced with an
The final products after peptization were stable fine average thickness above 2 ␮m, thus properly smooth-
colloidal ␥-AlOOH, TiO2 and mixed sols. According ing the roughness of the support [8]. Titania layers
to PCS measurements, the ␥-AlOOH sols contained obtained by a single dip-coating procedure showed an
plate-shaped particles with an average diameter of average thickness of ca. 0.5 ␮m. Therefore, a second
35 nm. Stable TiO2 sols consisted of cylindrical par- dipping procedure had to be used to increase the mem-
ticles with a mean size of about 37 nm. Due to the brane thickness up to ca. 1 ␮m (Fig. 1a). By applying

Table 2
Experimental conditions for the preparation of stable sols
Stable sols [H2 O]/[M]a [HNO3 ]/[M]a pH Temperature (◦ C) Particle size (nm) Characteristics

␥-AlOOH 100 0.07 4 90 35 Opaque blue


TiO2 b 25 0.50 1 80 37 Opaque blue
␥-AlOOH–TiO2 2 40 Opaque blue
Mixed 150 0.50 1 85 25 Transparent
TiO2 c 1 80 <5 Transparent
a [M] = [ASB] or [TTI].
b Colloidal sol.
c Polymeric sol.
78 T. Van Gestel et al. / Journal of Membrane Science 207 (2002) 73–89

Fig. 1. FESEM cross-sections (20,000×) pictures of titania (a) and mixed (b) colloidal sol–gel membrane layers on ␣-Al2 O3 (a) and
RBAO (b) supports.

an Al2 O3 –TiO2 combination as intermediate layer a appeared in the crystallographic phases ␥-AlOOH
thickness of ca. 1 ␮m was obtained (Fig. 1b). (boehmite) and anatase. X-ray diffraction patterns
In practice, in all the membrane formation proce- of binary alumina/titania membranes prepared by
dures a second dip-coating step was used after calcina- mixing separately peptized sols, show also typical
tion of the first layer. This was done to repair defects ␥-AlOOH and anatase diffraction peaks (Fig. 2). It
such as pinholes or regions with non-uniform layer may be concluded that this route did not yield a new
formation, which often appeared in the first sol–gel membrane material, but rather resulted in a binary
intermediate layer. Each final support thus produced ␥-AlOOH/anatase composite membrane material.
consisted of the main support with two sol–gel derived Pore structure characteristics of this membrane were
membrane layers. comparable with those of ␥-AlOOH membranes. This
suggested that the final calcinated membrane con-
3.3.2. Structural properties sisted of a ␥-alumina-like structure with titania grains
incorporated [11]. Pore size characterization showed
3.3.2.1. Standard thermal treatment. After thermal in each case a narrow pore size distribution in the
treatment at 400 ◦ C, alumina and titania membranes 3–4 nm range. The shape of the adsorption/desorption

Fig. 2. XRD patterns of binary ␥-AlOOH/anatase composite membrane material.


T. Van Gestel et al. / Journal of Membrane Science 207 (2002) 73–89 79

Table 3
Structural characteristics of colloidal alumina, titania and mixed membranes
Phase Firing (◦ C) Mean pore size (nm) Surface area (m2 g−1 ) Pore volume (ml g−1 ) Isotherm (type)

␥-AlOOH 400 3.7 318 0.27 IV


␥-Al2 O3 600 3.8 239 0.27 IV
␥-Al2 O3 800 4.6 154 0.22 IV
␦, ␪-Al2 O3 1000 6.3 87 0.18 IV
␣-Al2 O3 1200 9.2 10 0.09 II
Anatase 400 3.4 136 0.14 IV
Anatase 500 4.8 68 0.11 IV
Anatase/rutile 600 7.8 28 0.05 II, IV
Rutile 700 12.0 12 0.04 II
␥-AlOOH/anatasea 400 3.6 294 0.26 IV
Amorphousb 400 1.6 318 0.17 I
Amorphous/anataseb 800 4.8 120 0.14 IV
Amorphous/anataseb 900 7.6 53 0.14 IV
␣-Al2 O3 /anataseb 920 7.6 63 0.15 IV
␣-Al2 O3 /anatase/rutileb 950 8.0 13 0.07 IV, II
␣-Al2 O3 /anatase/rutileb 1000 8.9 5 0.05 II
a Composite membrane material.
b Mixed membrane material.

isotherms was characterized as type IV, which con- membrane material. In these stages, surface diffusion
firmed the fine mesoporous characteristics of these accompanied with neck-formation activity between
membrane materials. The comparable average sizes the ceramic grains played a major role, so that thermal
of ␥-AlOOH and titania particles in the sol state treatment involved mainly a gradual decrease in sur-
(35–40 nm) (Table 2) resulted in similar final mem- face area, while pore size and pore volume changed
brane pore sizes (Table 3). only little [13].
The structure of mixed membranes prepared from For alumina membranes, collapse of the fine meso-
mixed sols differed considerably from those reported porous structure started as conversion to the stable
above. In this case, the membrane material showed ␣-Al2 O3 phase took place. At 1200 ◦ C, after ␣-Al2 O3
a completely amorphous structure, which was re- transformation, grain size grew up to about 50 nm,
flected in a microstructure with an average pore size which was accompanied by strong surface area reduc-
below 2 nm. Additionally, the corresponding adsorp- tion (Fig. 3a). The shape of the adsorption/desorption
tion/desorption isotherms were found to be type I, isotherm was characterized as type II, indicating the
typical for a porous material with a mainly microp- presence of larger macropores [21].
orous texture. Membrane preparation from a mixed Measurements on titania membranes, showed a sim-
sol in which smaller particles were present (25 nm), ilar behaviour in the temperature region 500–700 ◦ C.
resulted in a pore structure with a smaller mean pore Up to 500 ◦ C, the material was present in the
size. anatase phase and average pore size was 3.5–5 nm.
From 500 ◦ C on, the anatase transformed into rutile
3.3.2.2. Thermal treatment at high temperatures. and the fine membrane structure changed drasti-
The effect of increasing firing temperature during cally. Fig. 3b shows that at 600–700 ◦ C accelerated
thermal treatment on structural properties of the dif- crystallite growth took place causing strong sur-
ferent membrane materials is summarized in Table 3 face area reduction and large changes in the pore
and Fig. 3. Initially, the relatively lower tempera- volume.
ture regions before phase transformations are char- For mixed membranes prepared from a homo-
acterized by small crystallite growth and relatively geneous mixed sol, the anatase/rutile transforma-
minor or gradual changes in the crystallinity of the tion occurred at higher temperature resulting in
80 T. Van Gestel et al. / Journal of Membrane Science 207 (2002) 73–89

Fig. 3. Structural development as a function of temperature for alumina (a), titania (b) and mixed (c) membranes.

an extended temperature region where only minor 3.3.3. Chemical stability


structural changes occur. Between 400 and 800 ◦ C,
the membrane material remained in an amorphous 3.3.3.1. Acid corrosion tests.
structure. Average pore diameters of 1.5–5 nm were
obtained and the resulting pore size distributions 3.3.3.1.1. Standard thermal treatment. Fig. 4
were rather narrow. Measurements from 800 ◦ C on shows the results of acid corrosion tests on colloidal
indicated the formation of an anatase phase. After sol–gel derived unsupported membrane layers ob-
thermal treatment at 920 ◦ C, this phase consisted tained after thermal treatment at 400 ◦ C. The amount
of crystalline grains with an average size of 20 nm. of dissolved aluminium is given as a function of pH
Fig. 3c shows that at higher firing temperatures, for ␥-AlOOH, ␥-AlOOH/anatase and mixed mem-
the formation of larger rutile grains resulted in a branes, for anatase the amount of dissolved titanium
strong increase in crystallite size and decrease of is given.
surface area and pore volume. From 950 ◦ C on, the The results illustrate that all membrane materials
shape of the adsorption/desorption isotherm changed showed a rather low corrosion at a pH above 3. At pH 2
also from type IV to II, indicating serious effects and 1, the membrane materials showed, depending on
on the membrane pore structure during transforma- the type of material, a different behaviour. The titania
tion. Meanwhile, the resulting membranes started to corrosion at pH 2 and 1 was low, so that the anatase
show a broad pore size distribution with considerable membrane material can be considered as chemically
tailing. stable down to pH 1. Contrarely, ␥-AlOOH and
T. Van Gestel et al. / Journal of Membrane Science 207 (2002) 73–89 81

Fig. 4. Corrosion behaviour of ␥-AlOOH, anatase, composite and mixed membranes (*pH 2:Ti concentration = 15 ␮g/l: pH 1:Ti
concentration = 80 ␮g/l).

alumina/titania membranes underwent drastic cor- 3.3.3.1.2. Thermal treatment at high temperatures.
rosion effects after exposure to pH 2 and 1, as the Al2 O3 : experimental results of corrosion tests with
nitric acid solutions contained very large amounts of alumina samples treated at different firing tempera-
dissolved Al. tures are summarized in Fig. 5 along with the phase
For ␥-AlOOH membranes, the results are in agree- structure.
ment with earlier findings of Schaep et al. [19], who It is clear that corrosion of alumina was very sen-
already demonstrated the poor chemical stability of sitive to the applied firing temperature. Between 400
␥-Al2 O3 after exposure to HCl solutions with a pH and 600 ◦ C, increase of the calcination temperature
below 3. The chemical stability of the different alu- resulted in a significant increase of the corrosion. In
mina/titania membrane materials prepared from bi- this temperature range, the membrane microstruc-
nary sols (␥-AlOOH/anatase) or from homogeneously ture changed as a result of the first alumina phase
mixed sols is even worse than those of pure ␥-alumina transformation involving the transition from an alu-
and therefore membrane applications are restricted to mina phase with semi-crystallinity (␥-AlOOH) to a
a pH above 3. Obviously, below pH 3, anatase is the mainly amorphous phase containing ␥-Al2 O3 . The
best membrane material. increase of corrosion with increasing firing temper-
From these measurements, it appeared that the ature can be explained as a result of the change in
crystalline properties of the ceramic base material microstructure and the decrease of crystallinity in
were determing for the membrane corrosion be- the membrane material. Further increase of the firing
haviour in acid solutions. For titania, the influence temperature between 600 and 1000 ◦ C, resulted in a
of the degree of crystallinity on the properties of gradual decrease of the corrosion. As in this temper-
the membrane material was apparent. Anatase tita- ature region no phase transformations occurred, the
nia showed a high degree of crystallinity and very most important change in the membrane microstruc-
small corrosion effects. Contrarely, membrane ma- ture involved a reduction of the specific surface area.
terials that contain aluminium were characterized as At around 1000 ◦ C, transformation into the crystal-
semi-crystalline or completely amorphous and high lographic transition phases ␦-Al2 O3 and ␪-Al2 O3
amounts of Al were released. The order of degree was observed. Although the corrosion resistance im-
of corrosion was: ␥-AlOOH < ␥-AlOOH/anatase < proved substantially, such a thermal treatment still
mixed Al2 O3 –TiO2 . The degree of crystallinity fol- did not yield a type of alumina membrane material
lowed the opposite order. with satisfactory corrosion resistance. After thermal
82 T. Van Gestel et al. / Journal of Membrane Science 207 (2002) 73–89

Fig. 5. Corrosion of alumina as a function of temperature.

treatment above 1000 ◦ C, transformation to the final Al2 O3 –TiO2 : the mixed preparation method yielded
crystallographic ␣-Al2 O3 phase started and from this amorphous membranes with micropores after heating
point on the material properties changed seriously: at 400 ◦ C and mesoporous semi-amorphous mem-
the mainly amorphous and semi-crystalline alumina branes after heating up to 800 ◦ C. At the same time,
membrane materials were transformed into a strong for all these membranes a low corrosion resistance
crystalline structure. XRD spectra showed very nar- was observed. Between 400 and 800 ◦ C, the most im-
row diffraction peaks after firing at 1200 ◦ C and the portant change was a reduction in surface area, along
estimated degree of crystallinity was very high. At the with a gradual decrease of the corrosion. After thermal
same time, these structural changes were accompa- treatment above 800 ◦ C, transformation to a crystalline
nied by serious improvements in the material stability. structure occurred, resulting in serious changes in the
The corrosion, both at pH 2 and 1, decreased: the structural and corrosion properties of the membrane
concentration of aluminium was 200 and 330 ␮g/l material. This transformation involved the formation
respectively, for a pH 2 and 1 corrosion test. of three chemically stable phases: ␣-Al2 O3 , anatase
As a result, in the case of alumina, pore size and and rutile. Between 900 and 1000 ◦ C, the transfor-
corrosion resistance behave oppositely with increas- mation of the membrane material resulted also in a
ing firing temperatures. Fig. 5 shows that alumina significant change in the corrosion behaviour (Fig. 6).
membranes prepared at temperatures below 1000 ◦ C At 920 ◦ C, anatase and ␣-Al2 O3 were the dominant
(␥-AlOOH, ␥-Al2 O3 , ␦-Al2 O3 , ␪-Al2 O3 ), with meso- phases and rutile occured in small quantities. As
porous properties, showed high corrosion in strong a result of the presence of two chemically stable
acid solutions. After thermal treatment at 1200 ◦ C, a phases, the corrosion decreased significantly and the
very good chemical stability was obtained, but the cor- pore structure remained relatively stable and kept its
responding pore size distribution showed a transition mesoporous properties. With increasing temperature,
of a fine mesoporous texture into a partially macrop- the degree of crystallinity for the different phases in-
orous structure. Therefore, ␣-Al2 O3 can be used e.g. creased sharply and a continuous improvement in the
as macroporous support material, but not as membrane corrosion behaviour was observed. A small increase
interlayer with a desired pore size distribution in the of the firing temperature from 920 to 950 ◦ C led to a
range 5–10 nm. three times lower corrosion rate. Unfortunately, as a
TiO2 : titania membranes were not subjected to result of drastic structural changes in this region, such
higher firing temperatures, because the anatase mem- as macropore formation and pore volume collapse,
branes prepared after thermal treatment at 400 ◦ C membranes treated at temperatures above 920 ◦ C were
already showed a very good corrosion resistance. unsuitable for mesoporous membrane applications.
T. Van Gestel et al. / Journal of Membrane Science 207 (2002) 73–89 83

Fig. 6. Corrosion of mixed membranes as a function of temperature.

3.3.3.2. Alkaline corrosion tests. Additionally, me- 3.4. Membrane toplayers


mbrane materials with mesoporous properties were
tested for their corrosion resistance in alkaline media. 3.4.1. Morphology
Table 4 shows that the corrosion tendencies in NaOH The finally obtained ceramic membranes can be cat-
solutions were largely comparable to those in HNO3 egorized as typical asymmetric multilayer membranes.
solutions. Fig. 7 shows the cross-section of such a membrane
For ␥-Al2 O3 membranes, low amounts of Al were consisting of a macroporous ␣-Al2 O3 support, two
released up to pH 11; above this pH, very large mesoporous colloidal sol–gel derived anatase mem-
amounts of Al were found. In this pH range, mixed brane layers with an overall thickness of ca. 1 ␮m, and
␣-Al2 O3 /anatase membranes showed a much lower a polymeric sol–gel derived anatase toplayer with a
corrosion rate. For this reason we propose these mixed thickness of ca. 100 nm. It appears that uniform and
membranes as a good alternative for the commonly discrete layers were formed and penetration of the
used mesoporous intermediate ␥-Al2 O3 membrane toplayer into the interlayer pores was succesfully pre-
layers. The price to pay for the improved corrosion vented.
resistance is an increase of the average pore diameter,
7.5 nm (Al2 O3 –TiO2 ) compared to 4 nm (␥-Al2 O3 ). 3.4.2. Structural properties
Therefore, when a smaller average pore size is re- Fig. 8 shows the phase behaviour of the toplayer
quired in combination with an improved corrosion material as a function of heating temperature. Ini-
resistance, membrane interlayers of anatase are to be tially, polymeric sol–gel derived titania showed a
preferred. completely amorphous structure. Above 200 ◦ C,

Table 4
Corrosion of mesoporous interlayer membrane materials in NaOH solutions
pH ␥-Al2 O3 (600 ◦ C), ␣-Al2 O3 /anatase (920 ◦ C), Anatase (400 ◦ C),
concentration Al (␮g/l) concentration Al (␮g/l) concentration Ti (␮g/l)
11 930 570 <10
12 48000 630 35
13 201000 730 50
84 T. Van Gestel et al. / Journal of Membrane Science 207 (2002) 73–89

Fig. 7. FESEM cross-section (50,000×) of a multilayer TiO2 membrane.

transformation into a crystalline structure took place. region, the Horvath–Kawazoe model (HK) was se-
This transformation produced the anatase phase, lected. At the same time, the Barrett–Joyner–Hallenda
which was present up to 600 ◦ C. Pore structure char- model (BJH) was applied for pore size analysis in
acterization of fine-grained polymeric sol–gel derived the mesoporous region [22]. Before transformation,
membranes, at the boundary of the meso- and the characterization of the membrane structure indicated
microporous region, appeared difficult because of the mainly a microporous character. Isotherm shape cor-
lack of a reliable method. Therefore, pore size char- responded to type I, representative for microporous
acterization was based on two models. In order to de- structures. With increasing temperature, crystalliza-
termine the pore size distribution in the microporous tion resulted in a decrease of the micropore volume

Fig. 8. XRD patterns of polymeric sol–gel titania.


T. Van Gestel et al. / Journal of Membrane Science 207 (2002) 73–89 85

Fig. 9. Micropore (a) and mesopore (b) distribution of unsupported polymeric sol–gel derived titania membranes.

and a gradual increase of the average pore size preservation of the fine pore properties and limited
(Fig. 9). After thermal treatment at 400 ◦ C, mainly a risk of cracking.
mesoporous anatase structure was formed with simi-
lar properties as the anatase membranes obtained via 3.4.3. Chemical stability
the collodial route (Table 5). Results of corrosion tests with polymeric sol–gel
Structural evolution of the membrane toplayer dur- derived titania membrane materials are summarized
ing calcination was evaluated by thermogravimetric in Fig. 11. After thermal treatment at 200 ◦ C, titania
analysis (TGA). In the polymeric sol–gel method, cal- corrosion was rather low at pH 2, but high amounts of
cination at an appropriate temperature was especially Ti were released during acid exposure at pH 1.
important in order to remove organic side products Between 200 and 300 ◦ C, the microstructure
and to obtain a stable final inorganic structure. In TGA changed significantly as a result of the transformation
studies, anatase membrane formed via a polymeric from an amorphous into a crystalline structure. As the
sol–gel route (Fig. 10a) appeared rather similar to that transformation involved the formation of the chemi-
formed by the colloidal route (Fig. 10b). No weight cally stable anatase phase, this resulted in a significant
losses were detected above 300 ◦ C, as all organics, decrease of the corrosion. With increasing tempera-
nitric acid and water were completely eliminated at ture, at 300 and 350 ◦ C, the degree of crystallinity in-
this temperature. This implies also that processing creased and a good corrosion resistance was obtained
of a stable polymeric sol–gel derived toplayer was over the complete acid range down to pH 1. Additional
perfectly possible at low thermal treatment with tests on the resistance against alkaline corrosion of the

Table 5
Pore structure characteristics of polymeric titania membranes
Phase Firing (◦ C) Mean pore size (nm)a Surface area (m2 g−1 ) Pore volume (ml g−1 ) Isotherm (type)
Amorphous 200 1.6 384 0.21 I
Anatase 300 2.4 307 0.19 I, IV
Anatase 400 3.4 130 0.13 IV
a BJH-model.
86 T. Van Gestel et al. / Journal of Membrane Science 207 (2002) 73–89

Fig. 10. TGA of polymeric (a) and colloidal sol–gel derived titania (b).

Fig. 11. Corrosion of polymeric titania as a function of temperature.

same membrane material proved also a high stability 4. Dynamic characterization of the multilayer
in NaOH solutions with a pH up to 13 (Table 6). membranes
From the results of all characterizations (pore struc-
ture, TGA, chemical stability), it was concluded that The retention behaviour of supported multilayer
an anatase toplayer membrane structure with microp- membranes was expressed by the MWCO, the molec-
ores and ultrafine mesopores, as obtained by a thermal ular weight of the component from the PEG mixture
treatment at 300 ◦ C, appeared to be the best toplayer that was retained for 90%. Membrane permeability
material. was determined according to the Hagen–Poisseuille
equation from flux measurements. As shown in
Table 6 Fig. 12, flux measurements as a function of pressure
Corrosion of polymeric sol–gel derived anatase (300 ◦ C) in HNO3 showed a linear dependency of flux on transmembrane
and NaOH solutions pressure.
pH Table 7 summarizes the MWCO and the mem-
1 2 3 11 12 13
brane permeability determined for membrane support
systems consisting of the main Al2 O3 support and
Ti concentration (␮g/l) 610 16 <10 <10 <10 11 two mesoporous colloidal sol–gel derived interlayers.
T. Van Gestel et al. / Journal of Membrane Science 207 (2002) 73–89 87

Fig. 12. Flux as a function of transmembrane pressure (TiO2 multilayer NF-membrane).

Table 7 rived anatase toplayer yielded multilayer membranes


Performance mesoporous UF-membrane supports with NF-characteristics. Firstly, with very smooth
Multilayer configuration MWCO Permeability high-quality ␥-Al2 O3 supporting layers, high-quality
(support = ␣-Al2 O3 ) (l h−1 m−2 bar−1 ) NF-membranes could be rather easily obtained. It
(1) ␥-Al2 O3 5000 5 appeared that the smooth surface of these support
(2) ␣-Al2 O3 /anatase 15000–20000 27 membrane layers was helpful for modification with a
(3) Anatase 10000–15000 22 high-quality crack-free anatase toplayer. Membrane
system (4) has a potential for applications in the low
NF-region (MWCO <200), but shows a low corro-
Membrane system (1), prepared by coating colloidal
sion resistance compared to the TiO2 membranes.
␥-Al2 O3 layers, showed the properties of a tight
This membrane configuration is especially interest-
UF-membrane with a MWCO of ca. 5000. Membrane
ing for NF-applications in neutral or non-aqueous
systems (2) and (3) showed both typical UF-properties
media. Performance measurements on the multi-
with a MWCO of ca. 15000 in combination with a
layer membrane configuration (5), consisting of an
higher permeability. Type II was developed as an alter-
anatase toplayer on an anatase UF-membrane support,
native for ␥-Al2 O3 membranes, but showed a smaller
showed a MWCO <1000 and a high permeability
thickness and larger pore size which resulted in a
(±20 l h−1 m−2 bar−1 ). This multilayer membrane,
higher permeability and cut-off value. The high per-
consisting of three anatase sol–gel layers, is there-
meability of the anatase support system (3) can also be
fore a convenient configuration for NF-applications
explained by the small layer thickness. The intermedi-
in aqueous corrosive media. A lower MWCO in
ate cut-off is probably due to the (inevitable) presence
the range of 200 is also posssible for a comparable
of small defects in these thin membrane layers.
multilayer membrane (6), consisting of an anatase
According to the results in Table 8, an addi-
intermediate UF-membrane support modified with an
tional modification with a polymeric sol–gel de-
optimized anatase toplayer obtained by a very slow
drying process, but for such a tighter NF-membrane
Table 8
Membrane performance disk-shaped multilayer membranes
the permeability is strongly reduced.

Multilayer configuration MWCO Permeability


(support = ␣-Al2 O3 ) (l h−1 m−2 bar−1 ) 5. Conclusions
(4) ␥-Al2 O3 /anatase <200 4
(5) Anatase/anatase 500–1000 19 The membranes obtained consist of a macrop-
(6) Anatase/anatase <200 2
orous ␣-Al2 O3 support, two mesoporous membrane
88 T. Van Gestel et al. / Journal of Membrane Science 207 (2002) 73–89

layers with an overall thickness ranging from Acknowledgements


1 ␮m for titania to 4 ␮m for alumina, and a fi-
nal polymeric titania toplayer with a thickness of Herman Cooreman is gratefully acknowledged
ca. 100 nm. Depending on the thermal treatment, for performing the ICP-Mass Spectroscopy mea-
␥-AlOOH, ␥-Al2 O3 , anatase, ␥-AlOOH/anatase, surements and Frans Servaes for his kind help and
amorphous mixed and ␣-Al2 O3 /anatase mixed mem- technical support.
brane layers were formed with narrow pore size
distributions ranging from the higher microporous
region (<2 nm) to the lower mesoporous region References
(2–8 nm).
[1] M. Mulder, Basic principles of membrane technology, 2nd
In mild aqueous (pH 3–11) or non-aqueous liquids,
Edition, Kluwer Academic, Dordrecht, 1996.
all membrane materials studied are chemically stable. [2] J.A. Hestekin, C.N. Smothers, D. Bhattacharyya, in: Procee-
In our NF-membrane development, strong preference dings of the 1999 AIChE Annual Meeting,Nanofiltration for
was given to a multilayer configuration, consisting Removal of Organics from Aqueous and Organic Solvent
of mesoporous ␥-Al2 O3 layers and an anatase NF Streams, cd-ROM Edition, Dallas, USA, 31 October–5
November, 1999.
toplayer due to the relatively easy processability of
[3] A.J. Burggraaf, L. Cot, Fundamentals of inorganic membrane
this membrane configuration. science and technology, Elsevier science and Technology
In strong acid (pH <3) or alkaline solutions Series 4, Elsevier, Amsterdam, 1996.
(pH >11), membrane materials consisting of weakly [4] J. Luyten, T. Van Gestel, J. Cooymans, C. Smolders, Proce-
crystallized or amorphous phases have to be avoided ssing of multilayer ceramic nanofiltration membranes, Inno-
vative Processing/Synthesis: Ceramics, Glasses, Composites
in the multilayer configuration. Although the finer
IV, Ceramic Transactions, Vol. 115, 2000.
pore structure and the high-quality layer forma- [5] A. Larbot, S. Alami-Younssi, M. Persin, J. Sarrazin, L.
tion of such materials is very advantageous, tests Cot, Preparation of a ␥-alumina nanofiltration membrane, J.
on the corrosion behaviour indicated that a transi- Membr. Sci. 97 (1994) 167–173.
tion to crystalline membrane structures is strictly [6] R. Vacassy, C. Guizard, V. Thoraval, L. Cot, Synthesis and
characterisation of microporous zirconia powders. Application
needed. For the development of membrane interlay- in nanofiltration characteristics, J. Membr. Sci. 132 (1997)
ers, mesoporous anatase or mixed ␣-Al2 O3 /anatase 109–118.
were considered. With a second dip-coating pro- [7] S. Vercauteren, Clay and pillared clay membranes: synthesis,
cedure included, preference was given to anatase characterization and transport properties, PhD thesis,
as interlayer material due to the higher chemi- University of Antwerp, Belgium, 1997.
[8] J. Luyten, J. Cooymans, C. Smolders, S. Vercauteren,
cal stability, smaller pore size and narrow pore E.F. Vansant, R. Leysen, Shaping of Multilayer Ceramic
size distribution, while a sufficient overall inter- Membranes by Dip-Coating, J. Eur. Ceram. Soc. 17 (1997)
layer thickness of ca. 1 ␮m was obtained. For the 273–279.
formation of the NF toplayer with a very fine mi- [9] J. Luyten, J. Cooymans, W. Adriansen, R. Leysen,
crostructure, only anatase is suitable as membrane in: Proceedings of the 9th CIMTEC—World Ceramic
Congress on Alternative process routes for better ceramic
material. membrane supports, Florence, Italy, 14–19 June, 1999,
The ␣-Al2 O3 support, the mesoporous ␥-Al2 O3 , pp. 449–456.
anatase and ␣-Al2 O3 /anatase interlayers and the [10] A.J. Burggraaf, K. Keizer, in: R.R. Bhave (Ed.), Synthesis of
anatase toplayer were characterized by a MWCO in Inorganic membranes, Inorganic membranes: Characterization
the micro-, ultra- and NF range, respectively. For and Applications, van Nostrand Rheinhold, New York, 1991,
pp. 10–63.
the finally optimized ␣-Al2 O3 /␥-Al2 O3 /anatase and [11] V.T. Zaspalis, W. Van Praag, K. Keizer, J.R.H. Ross, A.J.
␣-Al2 O3 /anatase/anatase NF-membranes, a MWCO Burggraaf, Synthesis and characterization of primary alumina,
below 200 was obtained. Due to their chemical sta- titania and binary membranes, J. Mater. Sci. 27 (1992) 1023–
bility with respect to non-aqueous organic liquids 1035.
or aqueous solutions with acid or alkaline pH, our [12] K.N. Kumar, Nanostructured ceramic membranes: layer
and texture formation, PhD thesis, University of Twente,
membrane development has provided ceramic mem- Enschede, The Netherlands, 1993.
branes applicable in a wide range of NF appli- [13] C.-H. Chang, R. Gopalan, Y.S. Lin, A comparative
cations. study on thermal and hydrothermal stability of alumina,
T. Van Gestel et al. / Journal of Membrane Science 207 (2002) 73–89 89

titania and zirconia membranes, J. Membr. Sci. 91 (1994) [18] J.M. Hofman-Züter, Chemical and thermal stability of
27–45. mesoporous ceramic membranes, PhD thesis, University of
[14] R.S.A. de Lange, J.H.A. Hekkink, K. Keizer, A.J. Burggraaf, Twente, Enschede, The Netherlands, 1995.
Formation and characterisation of supported microporous [19] J. Schaep, C. Vandecasteele, B. Peeters, J. Luyten, C.
ceramic membranes prepared by sol–gel modification Dotremont, D. Roels, Characteristics and retention properties
techniques, J. Membr. Sci. 99 (1995) 57–75. of a mesoporous ␥-Al2 O3 membrane for nanofiltration, J.
[15] S. Vercauteren, K. Keizer, E.F. Vansant, J. Luyten, R. Membr. Sci. 163 (1999) 229–237.
Leysen, Porous ceramic membranes: preparation, transport [20] T. Hollstein, in: Proceedings of the 9th CIMTEC—World
properties and applications, J. Porous Mater. 5 (1998) 241– Ceramic Congress on Corrosion of Advanced Ceramics by
258. Liquid Media: a Review, Florence, Italy, 14–19 June, 1999,
[16] I. Voigt, G. Fischer, P. Puhlfürb, D. Seifert, in: Procee- pp. 433–444.
dings of the 5th International Conference on Inorganic [21] S. Lowell, J.E. Shields, Powder surface area and porosity,
Membranes ICIM, New Filtration Ceramics—From Support 3rd edition, Powder technology series, Delft University of
to NF-Membrane Completely of TiO2 , Nagoya, Japan, 22–26 Technology, The Netherlands, 1991.
June, 1998, pp. 42–45. [22] H. Richter, G. Tomandl, C. Siewert, A. Piorra, in: Proceedings
[17] P. Puhlfürb, A. Voigt, R. Weber, M. Morbé, Microporous of the 5th International Conference on Inorganic Membranes
TiO2 membranes with a cut-off <500 Da, J. Membr. Sci. ICIM, Ceramic nanofiltration membranes made of ZrO2 and
174 (2000) 123–133. TiO2 , Nagoya, Japan, 22–26 June, 1998, pp. 30–33.

Вам также может понравиться